2009 Impoundment Design Manual
2009 Impoundment Design Manual
Second Edition
May 2009
Rev. Aug. 2010
RECORD OF REVISIONS
ENGINEERING AND DESIGN MANUAL
COAL REFUSE DISPOSAL FACILITIES
Second Edition
May 2009
Revision Date Page Comment
August 2010 6-144 Equation 6-18 correction
August 2010 6-151 Equation 6-23 defnition of B for 4 < Cu < 8
August 2010 7-83 Figure 7.22, Note 1 reference to Table 7.7
August 2010 8-51 Equation 8-13 correction
August 2010 9-30 Table 9.4 correction for size classifcation
August 2010 9-34 Typographical error
August 2010 9-37 Table 9.5 addition of WV and TX references
August 2010 R-14 Addition of reference (Howard, 1977)
ii
The engineering guidance presented in this document has been compiled or developed by DAppolonia
Engineering under contract with the Mine Safety and Health Administration (MSHA) from referenced
sources as well as input from MSHA and the coal industry. However, the guidance, recommenda-
tions, and conclusions presented herein may not necessarily represent the offcial policies of MSHA
or the U.S. government.
INSTRUCTIONS FOR USAGE OF THIS DVD MANUAL
This DVD requires Adobe Reader 7.0 or later. Other PDF viewing software may limit functionality.
Adobe Reader can be downloaded at http://get.adobe.com/reader/
Bookmarks are provided for the Manual chapters, appendices and the reference list and are dis-
played at the left of the text.
Links in the Manual text in BLUE will transfer you to another location in the Manual such as a section
header, fgure or table. If you wish to return to the original link location, you can hold down the ALT
key and hit the left arrow () key. On Mac computers hold down the Command key () and hit the
left arrow key. If you move around at the new location in the text, you may have to repeat this action
several times to return to your origination point. At the center of the bottom of most pages there is a
PREVIOUS VIEW link. Clicking on this link will return you to your previous location in the document
IF (and only if) you have not scrolled or otherwise moved away from the original landing location. Hit-
ting the left arrow key without holding down the ALT key will move you back one (1) page only.
The Table of Contents (TOC) listings are linked to the appropriate pages in the Manual text. Click-
ing on a specifc section header in the TOC will land you on the page where the section header is
located in the text. It is recommended that whole pages be displayed when clicking on the TOC links
because the section header that you are looking for may not be displayed on your monitor if the page
is zoomed up so that only part of the page is displayed. When you fnd the section you are looking
for, you can zoom to a comfortable reading size.
Links in GREEN will transfer you to technical references. Some of these are large documents and may
require a few seconds to load. These documents are in PDF format and are stored on the DVD. You
can return to the link in the Manual text by holding down the ALT key and hitting the left arrow () key.
On Mac computers hold down the Command key () and hit the left arrow key. The more you move
around in the reference document, the more times you will have to execute ALT + left arrow or + left
arrow. If you use the reference document extensively, you may not be able to link back to the Manual in
this manner, and then it will be necessary to close the reference document in Adobe Reader and then
go to the File > History list and reopen Manual 2009. This procedure will return you to the Manual title
page and not to the page where you linked from previously. Reference documents can be opened in a
new window by holding down the CTRL key when clicking on reference links, and we recommend this
procedure. On Mac computers hold down the Option key () when clicking the link.
If you load the Manual PDF to your computer, the references should be loaded to the same folder.
Suggestions for changes, corrections, or updates to the Manual should be directed to:
Mine Safety and Health Administration
Pittsburgh Technical Support Center
Mine Waste and Geotechnical Engineering Division
P.O. Box 18233
Pittsburgh, PA 15236
Attention: Dam Safety Offcer
Preface
MAY 2009
This Manual presents guidance on procedures for use in the engineering design, construction moni-
toring, operation and inspection of coal refuse impoundments and embankments in the United States.
It is an update of the original 1975 edition and refects advances in engineering, construction, and fa-
cility monitoring and operations practices. The primary intent of the Manual is to serve as a uniform
guide to safe refuse disposal practices for those concerned with coal mining and preparation. The
Manual serves this purpose in several ways by: (1) providing experienced embankment dam design
engineers with the characteristics of coal refuse and its disposal so that their experience can be ap-
propriately applied; (2) providing specialized technical knowledge concerning embankment design
in a form that can be used by engineers who do not specialize in this feld; (3) updating geotechnical,
structural, hydrologic and hydraulic design criteria for a range of embankment and impoundment
conditions, and spillway and drainage structures; (4) providing guidance on disposal requirements
and limitations for mine operators to include refuse disposal in the overall coal production operation;
and (5) providing guidance on construction, operation, inspection, monitoring and instrumentation,
and emergency action planning associated with the implementation of safe and reliable designs.
The 1975 edition of the Manual was prepared following the failure of a coal waste dam at Bufalo
Creek, West Virginia, that resulted in 125 fatalities. This Manual update was prompted by the recog-
nition that signifcant advances have been made in the felds of coal waste disposal and dam safety
in the 30-plus years since the original Manual was published. Another impetus was an incident that
occurred in Martin County, Kentucky, in 2000 in which over 300 million gallons of water and fne coal
refuse from a slurry impoundment broke into an underground mine. Slurry subsequently fowed out
of two mine openings and impacted streams in two separate watersheds. This incident prompted the
U.S. Congress to provide funding to the National Research Council (NRC) to examine ways to reduce
the potential for similar accidents. The NRCs report, Coal Waste Impoundments: Risks, Responses,
and Alternatives, which was released in 2002, included a number of recommendations for MSHA.
One recommendation was that MSHA continue to adopt and promote the best available technology
and practices with regard to site evaluation, design, construction, and operation of impoundments.
MSHA reported to Congress that one measure to address the NRC recommendations would be this
updating of the original Coal Refuse Design Manual.
The guidance presented in this Manual represents information, methods and procedures that are rec-
ommended for consideration by designers, coal operators, and regulators. The guidance presented in
this Manual is not regulation and cannot be enforced as such. It is not intended to preclude the ap-
plication of other credible methods and procedures or the use of other and new information that will
result in a safe and reliable coal refuse disposal facility. It is the responsibility of the designer to inves-
tigate the requirements of the project, recognize the unique and critical aspects of the site conditions,
and prepare designs that refect actual site conditions, features, loadings, and constraints.
iii
< PREVIOUS VIEW
iv
Preface
MAY 2009
In this update of the Manual, new chapters have been developed on seismic design and on site min-
ing and foundation issues, two topics that can have an impact on the type of disposal facility de-
signed. The long operating life of coal refuse facilities makes monitoring of embankment behavior
and facility maintenance particularly important. The sections on operation, monitoring, and instru-
mentation summarize procedures and devices to aid the designer and operator for defning and
implementing an appropriate feld observation program. In addition, general guidance is provided
for the preparation of emergency action plans. These plans are recommended for certain dams and
impoundments,and are required by some state regulatory agencies and encouraged by MSHA as
part of addressing hazardous conditions under 30 CFR 77.216-3.
In addition to concern for safety, the Manual addresses environmental considerations and controls
that may infuence the design of embankments and impoundments. Executive Order 11514 Protec-
tion and Enhancement of Environmental Quality, dated March 5, 1970, requires all Federal agencies
to Monitor, evaluate and control on a continuing basis their agencies activities so as to protect and
enhance the quality of the environment. Such activities shall include those directed to controlling
pollution and enhancing the environment and those designed to accomplish other program objec-
tives which may afect the quality of the environment. The Manual does not present guidance in es-
tablishing criteria for environmental controls (e.g., hydraulic conductivity of liner systems), because
such guidance is more appropriately lef to other references and regulatory agencies.
This Manual is intended to provide the designer with an important source of information. However,
the text and accompanying fgures, tables and references should not be applied without proper engi-
neering knowledge and judgment. Responsibility for actual design lies with the Professional Engineer
in responsible charge of the work. The use or application of the methods and information contained
herein is strictly the responsibility of the person utilizing the material. Designs should be based on
sound engineering principles and judgment and refect actual site conditions, and they should not
merely be paterned afer a successful design used at another location or possibly portrayed in the
Manual. The designer should be diligent and recognize that advances in approaches, criteria, and
methods will occur that may afect the applicability of portions of any reference or design guide.
This Manual was prepared by engineers and scientists with background and experience in the
subject mater, with input from MSHA personnel who review design plans and conduct investi-
gations at disposal sites. Overall direction and technical content were provided by Mr. Robert E.
Snow of DAppolonia Engineering. Manual coordination was performed by Dr. James L. Withiam,
DAppolonia, and fnal editing of the text was provided by Dr. J. Timothy Onstot. Proper recognition
to the entire Project Team and DAppolonia staf for their devoted eforts is not possible here. The
main contributors to and reviewers of technical chapters are noted in Table 1.
Special recognition is also given to Mr. John W. Fredland, Dam Safety Ofcer for MSHA, as the contract-
ing ofcers technical representative, Mr. Harold L. Owens, Mr. George H. Gardner, and to other MSHA
personnel who provided many valuable comments in suggesting content and reviewing the text for
publication. Finally, grateful appreciation is given to those in industry who provided input and review
comments during the process of preparation of this document. This includes input from the National
Mining Association and its consultants, as well as federal and state agencies and universities.
This Manual is available in hard copy and DVD. The DVD format includes hyperlinks and search capa-
bilities using Adobe Acrobat Reader sofware. Hyperlinks allow the display of the highlighted citation
of a fgure, table, appendix, or selected reference in the text. Selected references (in PDF format) that
are available in the public domain are included on the DVD version. For references not in the public
domain, reasonable eforts were made to obtain copyright permission. No hyperlink is provided for
references of substantial size, lack of availability in the public domain, or where permission for reprint
could not be obtained. The complete citation for all references is provided in the References section.
< PREVIOUS VIEW
v MAY 2009
Engineering and Design Manual Coal Refuse Disposal Facilities
Suggestions for changes, corrections, or updates to this Manual should be directed to:
Mine Safety and Health Administration
Pitsburgh Technical Support Center
Mine Waste and Geotechnical Engineering Division
P.O. Box 18233
Pitsburgh, PA 15236
Atention: Dam Safety Ofcer
For any suggested changes, corrections or updates, commenters should identify specifc pages, para-
graphs, tables, or fgures within the Manual, along with proposed replacement or new material.
Sources of proposed material should be completely cited, along with permission for use in a future
revised edition of this document.
This document is intended solely for noncommercial and educational purposes. The material pre-
sented in this Manual has been prepared in accordance with recognized engineering practices as cited
herein. The guidance should not be used without frst securing competent advice with respect to its
suitability for any application. The publication of the material contained herein is not intended as any
representation or warranty, either expressed or implied, on the part of the individuals, frms or agen-
cies involved, or any other person or organization named herein, that this information is suitable for
TABLE 1 MAJOR AND EXPERT CONTRIBUTORS AND REVIEWERS
Chapters 1 to 5
Introduction, Background,
Planning and Design Components
Mr. Robert E. Snow, DAppolonia Mr. Richard G. Almes
Chapter 6
Geotechnical Exploration,
Materials Testing, Engineering
Analysis and Design
Dr. James L. Withiam, DAppolonia
Mr. William J. Johnson, DAppolonia
Mr. Joseph W. Premozic, DAppolonia
Mr. Christopher J. Lewis, DAppolonia
Mr. Richard G. Almes
Dr. Umesh Dayal
Chapter 7
Seismic Design: Stability and
Deformation Analyses
Mr. Michael Paster, GEI
Mr. William J. Johnson, DAppolonia
Mr. Christopher J. Lewis, DAppolonia
Mr. Richard F. Tobin, GEI
Dr. Gonzalo Castro, GEI
Dr. William F. Marcuson
Dr. Peter K. Robertson
Mr. Blaise E. Genes, CEC
Mr. Richard G. Almes
Chapter 8
Site Mining and Foundation
Issues
Dr. James L. Withiam, DAppolonia
Mr. Robert E. Snow, DAppolonia
Mr. Christopher J. Lewis, DAppolonia
Dr. Vincent A. Scovazzo, J.T. Boyd
Mr. Frederick R. Vass, Alliance
Dr. Kelvin K. Wu
Mr. Richard G. Almes
Chapters 9 and 10
Hydrology and Hydraulics
Environmental Considerations
Ms. Colleen M. Campbell, DAppolonia
Mr. Robert E. Snow, DAppolonia
Dr. J. Timothy Onstott, DAppolonia
Mr. Richard G. Almes
Chapters 11 and 12
Construction and Disposal
Operations
Monitoring, Inspection and
Facility Maintenance
Mr. Robert M. Shusko, DAppolonia
Mr. Robert E. Snow, DAppolonia
Mr. Frederick R. Vass, Alliance
Mr. Richard G. Almes
Chapter 13
Instrumentation
Dr. James L. Withiam, DAppolonia Mr. Richard G. Almes
Chapter 14
Emergency Action Planning
Dr. J. Timothy Onstott, DAppolonia Mr. Richard G. Almes
< PREVIOUS VIEW
vi
Preface
MAY 2009
any general or specifc use or promises freedom from infringement of any patent or patents. Anyone
making use of this information assumes all liability resulting from such use. Reference to commercial
products or frms and use of trade names and trademarks in this document is for descriptive purposes
only, does not constitute endorsement, and may not be used for advertising or promotion purposes.
< PREVIOUS VIEW
Table of Contents
May 2009 vii
Preface ..........................................................................................................................................................................................iii
List of Tables ................................................................................................................................................................... xxvii
List of Figures ............................................................................................................................................................... xxxiii
Acronyms ................................................................................................................................................................................xli
Chapter 1 Introduction
Chapter 2 Background and Characterization
of Coal Refuse
2.1 GEOLOGICAL NATURE OF COAL REFUSE ............................................................................................................2-2
2.1.1 Origin of Coal ..............................................................................................................................................................2-2
2.1.2 Coal-Related Rocks ....................................................................................................................................................2-3
2.2 COAL MINING AND COAL PREPARATION (CLEANING) ..................................................................................2-4
2.3 REFUSE TRANSPORT AND DISPOSAL PLACEMENT ........................................................................................2-6
2.3.1 Coarse Refuse ..............................................................................................................................................................2-6
2.3.2 Fine Refuse and Combined Refuse .....................................................................................................................2-6
2.4 COAL REFUSE CHARACTERIZATION ......................................................................................................................2-7
2.4.1 General Geochemical Characteristics ................................................................................................................2-7
2.4.2 General Geotechnical Characteristics ................................................................................................................2-7
2.4.2.1 Coarse Coal Refuse ....................................................................................................................................................2-8
2.4.2.2 Fine Coal Refuse Slurry ............................................................................................................................................2-8
2.4.2.3 Dewatered Fine Coal Refuse ..................................................................................................................................2-9
2.4.2.4 Combined Coal Refuse ............................................................................................................................................2-9
2.4.2.5 Fine Coal Refuse Paste .......................................................................................................................................... 2-10
2.5 DISPOSAL PRACTICES .............................................................................................................................................. 2-10
2.5.1 Disposal Practices Prior to the Bufalo Creek Failure ................................................................................. 2-10
2.5.2 Disposal Practices Subsequent to the Bufalo Creek Failure ................................................................... 2-11
viii MAY 2009
Table of Contents
Chapter 3 Coal Refuse Disposal Facilities and
Other Impounding Structures
3.1 Hazard Potential for Disposal Facilities and Other Impounding Structures ................................................................3-1
3.2 Coal Refuse Impoundment Facilities ..................................................................................................................3-3
3.2.1 Size and Hazard Considerations ...........................................................................................................................3-4
3.2.1.1 Impoundment Defnition ...................................................................................................................................3-4
3.2.1.2 Impoundments in Series .....................................................................................................................................3-5
3.2.1.3 Incised Impoundments .......................................................................................................................................3-5
3.2.2 Disposal Facility Confguration .............................................................................................................................3-6
3.2.3 Hazard Potential Rating ...........................................................................................................................................3-9
3.3 COAL REFUSE NON-IMPOUNDING FACILITIES (REFUSE PILES) .............................................................. 3-10
3.3.1 Coarse Refuse Embankments............................................................................................................................. 3-10
3.3.2 Combined Refuse Embankments ..................................................................................................................... 3-11
3.3.3 Segregated Refuse Embankments ................................................................................................................... 3-11
3.3.4 Confguration and Development Staging ..................................................................................................... 3-11
3.4 COAL REFUSE SLURRY CELL FACILITIES ............................................................................................................ 3-12
3.4.1 Size and Hazard Classifcation ............................................................................................................................ 3-13
3.4.2 Disposal Facility Confguration and Development .................................................................................... 3-16
3.5 UNDERGROUND INJECTION SITES ..................................................................................................................... 3-17
3.5.1 Siting ........................................................................................................................................................................... 3-17
3.5.2 Injection System Design ...................................................................................................................................... 3-18
3.5.3 Risk Assessment and Response Plan ............................................................................................................... 3-19
3.5.4 Contingency Plan ................................................................................................................................................... 3-20
3.5.5 Monitoring Plan ...................................................................................................................................................... 3-20
3.6 RECOVERY (REMINING) OF COAL REFUSE DISPOSAL FACILITIES .......................................................... 3-21
3.7 OTHER IMPOUNDING STRUCTURES .................................................................................................................. 3-22
3.8 SMALL PONDS AND SIMILAR STRUCTURES ................................................................................................... 3-23
Chapter 4 Project Planning
4.1 UNIQUE ASPECTS OF REFUSE DISPOSAL ...........................................................................................................4-1
4.2 OPERATIONS AND SITE-RELATED FACTORS ......................................................................................................4-2
4.3 SUSTAINABLE MINING PRACTICES ........................................................................................................................4-4
4.4 ECONOMIC CONSIDERATIONS ................................................................................................................................4-4
4.5 ENVIRONMENTAL AND REGULATORY CONSIDERATIONS ...........................................................................4-6
4.6 PLANNING SEQUENCE ................................................................................................................................................4-7
4.7 DESIGN SEQUENCE ......................................................................................................................................................4-7
< PREVIOUS VIEW
ix MAY 2009
Engineering and Design Manual Coal Refuse Disposal Facilities
Chapter 5 Coal Refuse Disposal Facility
Design Components
5.1 DISPOSAL CAPACITY REQUIREMENTS AND SCHEDULING ..........................................................................5-1
5.1.1 Refuse Generation Rates and Design Capacity/Life ......................................................................................5-1
5.1.2 Slurry Impoundment Staging and Scheduling ...............................................................................................5-2
5.1.3 Slurry Cell Staging .....................................................................................................................................................5-3
5.1.4 Non-impounding Refuse Embankment Staging ...........................................................................................5-5
5.1.5 Amendments and Co-disposal of Combustion Waste .................................................................................5-7
5.1.5.1 Amendments for Neutralization and Stabilization of Refuse ................................................................5-7
5.1.5.2 Co-disposal of Combustion Waste ..................................................................................................................5-7
5.2 EROSION AND SEDIMENTATION CONTROL AND DESIGN STORM RUNOFF
MANAGEMENT ...............................................................................................................................................................5-8
5.2.1 Erosion and Sedimentation Control ...................................................................................................................5-8
5.2.2 Slurry Impoundment Infow Design Storm ......................................................................................................5-9
5.2.3 Slurry Cell Facilities ...................................................................................................................................................5-9
5.2.4 Refuse Embankment Design Storm ................................................................................................................. 5-11
5.3 FACILITY CONFIGURATION AND GEOMETRY ................................................................................................. 5-11
5.3.1 Regulatory Criteria ................................................................................................................................................. 5-11
5.3.2 Embankment Design Considerations ............................................................................................................. 5-11
5.3.2.1 Stability under Normal Operating Conditions ......................................................................................... 5-11
5.3.2.2 Stability under Extreme Events ..................................................................................................................... 5-12
5.3.2.3 Settlement and Subsidence............................................................................................................................ 5-12
5.3.2.4 Environmental Issues, Impoundment Elimination and Site Reclamation ...................................... 5-13
5.3.3 Impoundment Design Considerations ........................................................................................................... 5-13
5.3.3.1 Settling and Clarifcation ................................................................................................................................. 5-13
5.3.3.2 Seepage Control and Containment ............................................................................................................. 5-13
5.3.3.3 Spillway and Decant Systems......................................................................................................................... 5-13
5.3.3.4 Subsidence and Breakthrough Potential ................................................................................................... 5-13
5.4 FOUNDATION AND MINE SUBSIDENCE CONSIDERATIONS ..................................................................... 5-13
5.4.1 Foundation Materials ............................................................................................................................................ 5-13
5.4.2 Abutment and Foundation Geometry ............................................................................................................ 5-14
5.4.3 Mine Subsidence Potential ................................................................................................................................. 5-14
5.4.4 Impoundment Breakthrough Potential into Mine Workings .................................................................. 5-15
5.4.5 Mine Entry Barriers and Bulkheads .................................................................................................................. 5-16
5.5 INTERNAL DRAINAGE AND EMBANKMENT SEEPAGE CONTROL............................................................ 5-16
5.6 DECANT AND EMERGENCY SPILLWAY SYSTEMS .......................................................................................... 5-17
5.6.1 Decant Performance Considerations .............................................................................................................. 5-17
5.6.2 Open-Channel Spillway Performance Considerations .............................................................................. 5-18
5.7 SURFACE DRAINAGE CONTROLS ......................................................................................................................... 5-18
< PREVIOUS VIEW
x MAY 2009
Table of Contents
5.7.1 Permanent Drainage Controls ........................................................................................................................... 5-18
5.7.2 Temporary Drainage Controls ............................................................................................................................ 5-19
5.8 INSTRUMENTATION AND PERFORMANCE MONITORING ......................................................................... 5-19
5.8.1 Seepage ..................................................................................................................................................................... 5-19
5.8.2 Piezometric Levels .................................................................................................................................................. 5-19
5.8.3 Pool Levels ................................................................................................................................................................ 5-19
5.8.4 Rainfall Data ............................................................................................................................................................. 5-20
5.8.5 Deformation or Movement ................................................................................................................................. 5-20
5.9 RECLAMATION, ABANDONMENT AND POST-MINING LAND USE ......................................................... 5-20
Chapter 6 Geotechnical Exploration, Material Testing,
Engineering Analysis and Design
6.1 GEOTECHNICAL DESIGN PROCEDURES ..............................................................................................................6-2
6.2 GENERAL CONSIDERATIONS ....................................................................................................................................6-2
6.2.1 Unique Characteristics of Refuse Disposal Facilities .....................................................................................6-2
6.2.2 Site Conditions ...........................................................................................................................................................6-4
6.2.2.1 Topography..............................................................................................................................................................6-4
6.2.2.2 Climate/Weather ....................................................................................................................................................6-6
6.2.2.3 Geology and Surfcial Soils .................................................................................................................................6-6
6.2.2.4 Miscellaneous Site Considerations ..................................................................................................................6-7
6.2.3 Embankment Materials ...........................................................................................................................................6-8
6.2.3.1 Coarse Coal Refuse ................................................................................................................................................6-8
6.2.3.2 Fine Coal Refuse .................................................................................................................................................. 6-10
6.2.3.3 Combined Refuse ............................................................................................................................................... 6-10
6.2.3.4 Borrow Materials ................................................................................................................................................. 6-11
6.2.3.5 Coal Combustion Products from Power Plants ........................................................................................ 6-13
6.2.4 Scheduling ................................................................................................................................................................ 6-18
6.3 DESIGN CONSIDERATIONS .................................................................................................................................... 6-19
6.3.1 Impounding Embankments ............................................................................................................................... 6-19
6.3.1.1 Homogeneous Embankments ........................................................................................................................... 6-21
6.3.1.2 Zoned Embankments ........................................................................................................................................ 6-21
6.3.1.3 Foundation Seepage Control ......................................................................................................................... 6-26
6.3.2 Non-impounding Embankments ..................................................................................................................... 6-30
6.3.3 Slurry Cell Embankments .................................................................................................................................... 6-34
6.3.4 Embankment Construction Staging ................................................................................................................ 6-35
6.3.4.1 Upstream Method .............................................................................................................................................. 6-35
6.3.4.2 Downstream Method ........................................................................................................................................ 6-37
6.3.4.3 Centerline Method ............................................................................................................................................. 6-38
6.3.4.4 Embankments Supplemented with Borrow Material ............................................................................ 6-39
6.3.5 Special Considerations for Existing Embankments .................................................................................... 6-39
6.3.6 Other Impounding Embankments ................................................................................................................... 6-41
< PREVIOUS VIEW
xi MAY 2009
Engineering and Design Manual Coal Refuse Disposal Facilities
6.4 SITE GEOTECHNICAL/GEOLOGICAL EXPLORATION .................................................................................... 6-43
6.4.1 Background Data Sources ................................................................................................................................... 6-44
6.4.1.1 Topographic Maps.............................................................................................................................................. 6-44
6.4.1.2 Geologic Maps and Publications................................................................................................................... 6-46
6.4.1.3 Agricultural Soil Surveys .................................................................................................................................. 6-47
6.4.1.4 Satellite Imagery, Aerial Photographs and Other Imagery.................................................................. 6-47
6.4.1.5 Past Investigations and Area Mining Plans ............................................................................................... 6-48
6.4.1.6 Individual Site Mapping ................................................................................................................................... 6-49
6.4.2 Surfcial Reconnaissance and Geologic Mapping ....................................................................................... 6-49
6.4.3 Subsurface Exploration and In-Situ Test Planning ...................................................................................... 6-50
6.4.3.1 Program Planning ............................................................................................................................................... 6-52
6.4.3.2 Subsurface Exploration and In-situ Test Methods and Applicability ............................................... 6-54
6.4.3.3 Test Pits ................................................................................................................................................................... 6-55
6.4.3.4 Boring Methods and the Standard Penetration Test (SPT) .................................................................. 6-57
6.4.3.5 Undisturbed Soil Sampling ............................................................................................................................. 6-59
6.4.3.6 Rock Coring ........................................................................................................................................................... 6-59
6.4.3.7 Cone Penetrometer Test (CPT) and Piezocone Penetrometer Test (CPTu or PCPT) .................... 6-61
6.4.3.8 Field Vane Shear Test (FVST) ........................................................................................................................... 6-63
6.4.3.9 Directional (Longhole) Drilling ...................................................................................................................... 6-66
6.4.3.10 Field Hydraulic Conductivity Tests ............................................................................................................... 6-67
6.4.3.11 Groundwater-Level Measurements ............................................................................................................. 6-68
6.4.3.12 Water Flow and Quality Tests ......................................................................................................................... 6-69
6.4.3.13 Backflling of Boreholes .................................................................................................................................... 6-69
6.4.4 Geophysical Methods ........................................................................................................................................... 6-70
6.4.4.1 Surfcial Geophysical Techniques .................................................................................................................. 6-71
6.4.4.1.1 Seismic Refraction .......................................................................................................................................... 6-71
6.4.4.1.2 Seismic Refection .......................................................................................................................................... 6-74
6.4.4.1.3 Electrical Resistivity ........................................................................................................................................ 6-76
6.4.4.1.4 Spontaneous Potential (SP) ........................................................................................................................ 6-78
6.4.4.1.5 Electromagnetics (EM) .................................................................................................................................. 6-79
6.4.1.1.6 Ground Penetrating Radar (GPR) .............................................................................................................. 6-80
6.4.4.1.7 Gravity................................................................................................................................................................. 6-82
6.4.4.1.8 Magnetics .......................................................................................................................................................... 6-83
6.4.4.2 Borehole Geophysical Techniques ............................................................................................................... 6-83
6.4.4.2.1 Crosshole and Downhole/Uphole Seismic Surveys ........................................................................... 6-85
6.4.4.2.2 Video and Laser/Sonar Imaging ................................................................................................................ 6-89
6.5 MATERIAL PROPERTY DETERMINATION THROUGH TESTING .................................................................. 6-90
6.5.1 Selection of Samples for Testing ....................................................................................................................... 6-91
6.5.1.1 Bulk and Disturbed Samples .......................................................................................................................... 6-91
6.5.1.2 Undisturbed Samples ....................................................................................................................................... 6-95
6.5.1.3 Reconstituted Samples ..................................................................................................................................... 6-96
< PREVIOUS VIEW
xii MAY 2009
Table of Contents
6.5.2 Classifcation and Index Property Tests .......................................................................................................... 6-96
6.5.2.1 Moisture Content ................................................................................................................................................ 6-97
6.5.2.2 Specifc Gravity ..................................................................................................................................................6-100
6.5.2.3 Atterberg Limits ................................................................................................................................................6-101
6.5.2.4 Particle-size Distribution ................................................................................................................................6-102
6.5.2.5 Chemical Characterization ............................................................................................................................6-104
6.5.3 Compaction and Density Tests ........................................................................................................................6-105
6.5.3.1 Fine-grained Soils and Coal Refuse ............................................................................................................6-105
6.5.3.2 Coarse-Grained Soils and Coal Refuse ......................................................................................................6-108
6.5.4 Hydraulic Conductivity Tests ............................................................................................................................6-109
6.5.5 Geosynthetic Materials Tests ............................................................................................................................6-111
6.5.6 Consolidation Tests ..............................................................................................................................................6-112
6.5.7 Shear Strength and Related Tests ...................................................................................................................6-116
6.5.7.1 Vane-Shear Test .................................................................................................................................................6-120
6.5.7.2 Direct-Shear Test ...............................................................................................................................................6-120
6.5.7.3 Unconfned-Compression Test ....................................................................................................................6-122
6.5.7.4 Triaxial-Compression Test ..............................................................................................................................6-124
6.5.8 Seismic Property Characterization .................................................................................................................6-128
6.5.8.1 Cyclic-Triaxial Test .............................................................................................................................................6-128
6.5.8.2 Cyclic Loading Followed by Monotonic Loading ..................................................................................6-128
6.5.8.3 Resonant-Column Test ....................................................................................................................................6-129
6.5.9 Rock Property Tests ..............................................................................................................................................6-129
6.5.9.1 Point-Load Index Test ......................................................................................................................................6-129
6.5.9.2 Unconfned Compressive Strength Test ...................................................................................................6-132
6.5.9.3 Indirect Tensile Strength Test .......................................................................................................................6-133
6.5.9.4 Rock Durability ..................................................................................................................................................6-133
6.5.10 Structural Material Tests .....................................................................................................................................6-135
6.5.10.1 Concrete ...............................................................................................................................................................6-135
6.5.10.2 Controlled Low-Strength Materials (CLSM) .............................................................................................6-135
6.5.11 Verifcation of Test Results .................................................................................................................................6-137
6.6 GEOTECHNICAL ANALYSES .................................................................................................................................6-138
6.6.1 Analytical and Numerical Methods in Geotechnical Engineering ......................................................6-138
6.6.2 Seepage Analysis ..................................................................................................................................................6-139
6.6.2.1 Basic Assumptions, Conditions Requiring Seepage Analysis ...........................................................6-139
6.6.2.2 Seepage Analysis Methods ...........................................................................................................................6-142
6.6.2.2.1 Graphical (Flow Net) Approach ...............................................................................................................6-143
6.6.2.2.2 Analytical Solutions .....................................................................................................................................6-144
6.6.2.2.3 Numerical Modeling Approach ...............................................................................................................6-145
6.6.2.3 Seepage Control Measures ...............................................................................................................................6-146
6.6.2.3.1 Granular Drain and Filter Requirements...............................................................................................6-146
6.6.2.3.2 Geotextile-Wrapped Drains ......................................................................................................................6-149
6.6.2.3.3 Seepage Control along Conduits ............................................................................................................6-156
< PREVIOUS VIEW
xiii MAY 2009
Engineering and Design Manual Coal Refuse Disposal Facilities
6.6.2.3.4 Relief Wells ......................................................................................................................................................6-160
6.6.2.3.5 Horizontal Drains ..........................................................................................................................................6-161
6.6.2.3.6 Impoundment Liners ..................................................................................................................................6-163
6.6.2.3.7 Foundation Seepage Cutofs ....................................................................................................................6-164
6.6.2.3.8 Impoundment Water and Slurry Deposition Management ..........................................................6-164
6.6.2.4 Seepage Measurements.................................................................................................................................6-164
6.6.3 Settlement Analysis .............................................................................................................................................6-164
6.6.3.1 Conditions Requiring Deformation Analysis ..........................................................................................6-164
6.6.3.2 Settlement and Deformation Analysis ......................................................................................................6-167
6.6.3.3 Deformation Control Measures ...................................................................................................................6-170
6.6.3.4 Deformation Measurements ........................................................................................................................6-170
6.6.4 Slope Stability Analysis .......................................................................................................................................6-171
6.6.4.1 Conditions Requiring Stability Analysis....................................................................................................6-171
6.6.4.2 Methods of Stability Analysis .......................................................................................................................6-173
6.6.4.2.1 Limit Equilibrium Stability Analysis ........................................................................................................6-173
6.6.4.2.2 Stability/Deformation Analysis Using Finite Element Methods...................................................6-181
6.6.4.3 Acceptable Factors of Safety ........................................................................................................................6-181
6.6.4.4 Stability Control Measures ............................................................................................................................6-184
6.6.4.5 Stability Measurements ..................................................................................................................................6-185
6.6.5 Rock Excavations ..................................................................................................................................................6-185
6.6.5.1 Conditions Requiring Stability Analysis....................................................................................................6-185
6.6.5.1.1 Existing Ground Surface Slope ................................................................................................................6-185
6.6.5.1.2 Overburden Soil Thickness, Type and Quality ....................................................................................6-185
6.6.5.1.3 Bedrock Surface Conditions .....................................................................................................................6-186
6.6.5.1.4 Rock Type, Quality and Confguration...................................................................................................6-187
6.6.5.1.5 Groundwater Conditions ...........................................................................................................................6-189
6.6.5.1.6 Other Factors ..................................................................................................................................................6-189
6.6.5.2 Slope Stability Analysis ...................................................................................................................................6-189
6.6.5.3 Stability Control Measures ............................................................................................................................6-191
6.6.5.4 Stability Measurements ..................................................................................................................................6-193
6.6.6 Conduit Structural Design for Earthen Fill Loads ......................................................................................6-193
6.6.6.1 Conduit Types ....................................................................................................................................................6-194
6.6.6.1.1 Concrete ...........................................................................................................................................................6-194
6.6.6.1.2 Thermoplastic ................................................................................................................................................6-195
6.6.6.1.3 Metal .................................................................................................................................................................6-195
6.6.6.2 Soil-Structure Design ......................................................................................................................................6-196
6.6.6.3 Conduit Installation and Design Details ...................................................................................................6-201
6.6.6.3.1 Installation ......................................................................................................................................................6-201
6.6.6.3.2 Rigid Conduit Support ................................................................................................................................6-201
6.6.6.3.3 Flexible Conduit Support ...........................................................................................................................6-202
6.6.6.3.4 Watertightness ..............................................................................................................................................6-204
6.6.7 Blasting Impacts ....................................................................................................................................................6-206
< PREVIOUS VIEW
xiv MAY 2009
Table of Contents
Chapter 7 Seismic Design: Stability and
Deformation Analyses
7.1 GENERAL ..........................................................................................................................................................................7-1
7.1.1 Design Approach .......................................................................................................................................................7-1
7.1.1.1 Seismic Instability ..................................................................................................................................................7-2
7.1.1.2 Excessive Deformations ......................................................................................................................................7-3
7.1.2 Seismic Design Considerations and Flow Chart .............................................................................................7-3
7.1.3 Sand-Like Versus Clay-Like Material ....................................................................................................................7-9
7.1.4 Susceptibility to Strength Loss .......................................................................................................................... 7-11
7.1.5 Simplifed Steps for Seismic Stability .............................................................................................................. 7-12
7.1.5.1 Step 1 ...................................................................................................................................................................... 7-12
7.1.5.2 Step 2 ...................................................................................................................................................................... 7-12
7.1.5.3 Step 3 ...................................................................................................................................................................... 7-12
7.1.5.4 Step 4 ...................................................................................................................................................................... 7-12
7.1.5.5 Step 5 ...................................................................................................................................................................... 7-13
7.1.5.6 Step 6 ...................................................................................................................................................................... 7-13
7.2 TERMINOLOGY ............................................................................................................................................................ 7-14
7.3 CHARACTERIZATION OF SUBSURFACE CONDITIONS AND MATERIAL PROPERTIES ...................... 7-18
7.4 SEISMIC STABILITY ANALYSES .............................................................................................................................. 7-21
7.4.1 General Discussion ................................................................................................................................................. 7-21
7.4.1.1 Triggering Analyses ........................................................................................................................................... 7-21
7.4.1.2 Evaluation of Post-Earthquake Strength .................................................................................................... 7-22
7.4.1.3 Stability Analysis Using Post-Earthquake Strengths .............................................................................. 7-23
7.4.2 Triggering Analyses ............................................................................................................................................... 7-23
7.4.2.1 Pore-Pressure-Based Method for Triggering of Strength Loss in Clay-Like Material .................. 7-23
7.4.2.2 Pore-Pressure-Based Method for Triggering of Strength Loss in Sand-Like Material ................ 7-23
7.4.2.2.1 Evaluation of Cyclic Stress Ratio (CSR) ..................................................................................................... 7-25
7.4.2.2.2 Evaluation of Cyclic Resistance Ratio (CRR) ........................................................................................... 7-28
7.4.2.2.3 Correction Factors for Earthquake Magnitude, High Overburden, Sloping Ground
and Age of Deposit ........................................................................................................................................ 7-31
7.4.2.3 Strain-Based and Stress-Based Methods to Evaluate Triggering of Strength Loss ......................... 7-32
7.4.2.3.1 Clay-Like Material: Strain-Based Method for Triggering ....................................................................... 7-32
7.4.2.3.2 Sand-Like Material: Strain-Based Method for Triggering ..................................................................... 7-33
7.4.2.3.3 Sand-Like Material: Stress-Based Method for Triggering ..................................................................... 7-35
7.4.3 Evaluation of Post-Earthquake Strength ........................................................................................................ 7-36
7.4.3.1 Correlations of SPT and CPT with S
us
of Sand-Like Material ................................................................ 7-36
7.4.3.2 Laboratory Testing for Measuring Post-Earthquake Strength for Clay-Like Material
and S
us
for Sand-Like Material......................................................................................................................... 7-41
7.4.3.2.1 Laboratory Testing Issues ............................................................................................................................ 7-41
7.4.3.2.2 Laboratory Testing of Soft Clay-like Material to Measure Post-Earthquake Strength ............ 7-41
7.4.3.2.3 Laboratory Testing of Sand-like Material to Measure S
us
..................................................................7-44
< PREVIOUS VIEW
xv MAY 2009
Engineering and Design Manual Coal Refuse Disposal Facilities
7.4.3.3 Field testing to Measure S
us
of Soft Clay-Like Material .......................................................................... 7-51
7.4.4 Analysis Steps .......................................................................................................................................................... 7-51
7.4.4.1 Step 1 - Defne Embankment Geometry .................................................................................................... 7-51
7.4.4.2 Step 2 - Screen for Potential Strength Loss ............................................................................................... 7-51
7.4.4.2.1 Sand-Like Material .......................................................................................................................................... 7-51
7.4.4.2.2 Clay-Like Material ........................................................................................................................................... 7-52
7.4.4.2.3 Screening Results ........................................................................................................................................... 7-54
7.4.4.3 Step 3 - Defne Post-Earthquake Strengths for Limit Equilibrium Stability Analyses ................. 7-54
7.4.4.4 Step 4 - Perform Initial Stability Analysis.................................................................................................... 7-56
7.4.4.5 Optional Step 5 - Perform More Detailed Evaluation of S
us
for Sand-Like Material .................... 7-57
7.4.4.6 Optional Step 6 - Perform Triggering Analysis ......................................................................................... 7-57
7.4.5 Comments on Methods for Evaluating Triggering and Strength .......................................................... 7-58
7.5 SEISMIC DEFORMATION ANALYSES ................................................................................................................... 7-59
7.5.1 General Discussion ................................................................................................................................................. 7-59
7.5.2 Preliminary Screening ........................................................................................................................................... 7-61
7.5.3 Pseudo-Static Procedure for Screening .......................................................................................................... 7-61
7.5.4 Newmark-Type Analysis (No Cyclic Mobility) ............................................................................................... 7-62
7.5.5 Numerical Modeling with No Cyclic Mobility ............................................................................................... 7-65
7.5.6 Numerical Modeling Considering Cyclic Mobility ...................................................................................... 7-65
7.5.7 Acceptable Deformations .................................................................................................................................... 7-66
7.6 EMBANKMENT MODIFICATIONS FOR IMPROVING STABILITY OR RESISTANCE TO
DEFORMATIONS ......................................................................................................................................................... 7-66
7.7 SEISMIC HAZARD ASSESSMENT (SEISMICITY) ............................................................................................... 7-66
7.7.1 Analytical Procedures............................................................................................................................................ 7-71
7.7.1.1 Deterministic Seismic Hazard Analysis (DSHA) ........................................................................................ 7-71
7.7.1.2 Probabilistic Seismic Hazard Analysis (PSHA) ........................................................................................... 7-73
7.7.1.3 DSHA versus PSHA ............................................................................................................................................. 7-74
7.7.2 Seismotectonic Modeling ................................................................................................................................... 7-76
7.7.2.1 Regional Tectonic Framework ........................................................................................................................ 7-77
7.7.2.2 Historical Seismicity ........................................................................................................................................... 7-78
7.7.2.3 Selection of Seismic Sources .......................................................................................................................... 7-78
7.7.2.4 Maximum Magnitude ....................................................................................................................................... 7-79
7.7.2.5 Earthquake Recurrence .................................................................................................................................... 7-81
7.7.3 Design Ground Motion ......................................................................................................................................... 7-81
7.7.3.1 Selection of Design Earthquakes .................................................................................................................. 7-82
7.7.3.2 Ground Motion Attenuation ........................................................................................................................... 7-83
7.7.3.3 Selection of Peak Ground Motion Parameters ......................................................................................... 7-84
7.7.3.4 Selection of Acceleration Time Histories .................................................................................................... 7-84
7.7.3.5 Applicability of Design Ground Motion ..................................................................................................... 7-86
7.7.3.6 Application of Seismic Parameters in Design Process .......................................................................... 7-86
7.7.3.7 Simplifed Design Ground Motion for Areas of Low Seismic Hazard ............................................... 7-86
7.8 SEISMIC DESIGN OVERVIEW .................................................................................................................................. 7-88
< PREVIOUS VIEW
xvi MAY 2009
Table of Contents
APPENDIX 7A DERIVATIONS OF BASIC EQUATIONS FOR STEADY-STATE LABORATORY TESTING
APPENDIX 7B VOID-RATIO MEASUREMENTS DURING UNDISTRUBED TUBE SAMPLING AND
LABORATORY TESTING
APPENDIX 7C PROCEDURE FOR UNDISTURBED, FIXED-PISTON SAMPLING OF COHESIONLESS SOIL
APPENDIX 7D COMBINING NEWMARK-TYPE DISPLACEMENTS COMPUTED USING 1D SITE-
RESPONSE ANALYSIS
Chapter 8 Site Mining and Foundation Issues
8.1 GENERAL CONSIDERATIONS ....................................................................................................................................8-2
8.2 Available Sources of Site Information ................................................................................................................8-3
8.2.1 Mine Maps ....................................................................................................................................................................8-3
8.2.2 Coal Contour Maps and Outcrops .......................................................................................................................8-4
8.2.3 Interpretation and Accuracy of Mine Maps .....................................................................................................8-5
8.3 ON-SITE RECONNAISSANCE AND EXPLORATION ............................................................................................8-6
8.3.1 Surfcial Reconnaissance .........................................................................................................................................8-6
8.3.2 Geophysical Methods ..............................................................................................................................................8-7
8.3.3 Long-Hole Directional Drilling ..............................................................................................................................8-7
8.3.4 Conventional Drilling and Sampling ..................................................................................................................8-7
8.4 EVALUATION OF MINE SUBSIDENCE AND BREAKTHROUGH ......................................................................8-9
8.4.1 Mine Subsidence Considerations ..................................................................................................................... 8-10
8.4.2 Potential Subsidence and Failure Mechanisms ........................................................................................... 8-14
8.4.2.1 Sinkholes ............................................................................................................................................................... 8-15
8.4.2.2 Pillar Failure ........................................................................................................................................................... 8-18
8.4.2.3 Pillar Punching (Floor Failure) ........................................................................................................................ 8-19
8.4.2.4 Outcrop Barrier Failure by Shear or Punching .......................................................................................... 8-20
8.4.2.5 Internal Erosion ................................................................................................................................................... 8-20
8.4.2.6 Bulkhead Failure.................................................................................................................................................. 8-22
8.4.2.7 Trough Subsidence and Subsidence Cracks ............................................................................................. 8-22
8.4.2.8 Failures Related to Auger and Highwall Mining ...................................................................................... 8-23
8.4.2.9 Hillside Movement or Disturbance .............................................................................................................. 8-24
8.4.2.10 Mine Blowout ....................................................................................................................................................... 8-24
8.4.2.11 Seismic and Blasting Events ........................................................................................................................... 8-24
8.4.3 Mine Subsidence Potential and Analysis ....................................................................................................... 8-25
8.4.3.1 Pillar Evaluation and Analysis ......................................................................................................................... 8-25
8.4.3.2 Subsidence Evaluation and Analysis ........................................................................................................... 8-28
8.4.3.3 Subsidence Damage Criteria .......................................................................................................................... 8-29
8.4.4 Mine Breakthrough Potential and Analysis ................................................................................................... 8-30
8.4.4.1 Outcrop Barriers .................................................................................................................................................. 8-30
8.4.4.2 Overburden Barriers .......................................................................................................................................... 8-33
8.5 SUBSIDENCE AND BREAKTHROUGH MITIGATION METHODS ................................................................ 8-34
8.5.1 Mine Subsidence and Breakthrough Mitigation ......................................................................................... 8-35
< PREVIOUS VIEW
xvii MAY 2009
Engineering and Design Manual Coal Refuse Disposal Facilities
8.5.1.1 Use of Safety Zones ............................................................................................................................................ 8-35
8.5.1.2 Mine Backflling ................................................................................................................................................... 8-35
8.5.1.3 Stowing of Mine Openings and Associated Barrier Construction .................................................... 8-36
8.5.1.4 Bulkhead Construction ..................................................................................................................................... 8-38
8.5.1.5 Construction of Compacted Earthen Barriers on the Surface ............................................................ 8-39
8.5.1.6 Conversion to Slurry Cells ................................................................................................................................ 8-39
8.5.1.7 Sealing Sources of Leakage ............................................................................................................................ 8-40
8.5.1.8 Stabilization of Fines ......................................................................................................................................... 8-40
8.5.1.9 Overexcavation and Induced Subsidence ................................................................................................. 8-41
8.5.1.10 Monitoring Provisions ....................................................................................................................................... 8-42
8.5.2 Mine Entry Bulkhead Seal Design ..................................................................................................................... 8-42
8.5.2.1 Site Preparation ................................................................................................................................................... 8-42
8.5.2.2 Bulkhead Types.................................................................................................................................................... 8-43
8.5.2.3 Design Loads ........................................................................................................................................................ 8-47
8.5.2.4 Design of Solid Concrete Block Bulkheads ................................................................................................ 8-47
8.5.2.5 Design of Concrete Bulkheads ....................................................................................................................... 8-48
8.5.2.5.1 Design of Reinforced-Concrete Bulkheads for Flexure ..................................................................... 8-48
8.5.2.5.2 Design of Unreinforced-Concrete Plug Bulkheads for Shear .......................................................... 8-49
8.5.2.5.3 Design of Reinforced-Concrete Bulkheads for Shear ........................................................................ 8-50
8.5.2.6 Monitoring of Bulkheads ................................................................................................................................. 8-51
8.6 FOUNDATION-RELATED CONSTRUCTION AND OPERATIONS MONITORING ................................... 8-52
8.7 MINE BACKFILLING DESIGN .................................................................................................................................. 8-53
8.8 SURFACE MINE SPOIL ISSUES ............................................................................................................................... 8-55
8.8.1 Surface Mine Spoil Characteristics ............................................................................................................... 8-55
8.8.2 Design and Construction Considerations .................................................................................................. 8-56
8.9 SURFACE MINE HIGHWALL ISSUES ..................................................................................................................... 8-59
Chapter 9 Hydrology and Hydraulics
9.1 GENERAL CONSIDERATIONS ....................................................................................................................................9-2
9.2 HYDROLOGY AND HYDRAULICS PRINCIPLES ...................................................................................................9-2
9.2.1 Basic Design Principles ............................................................................................................................................9-5
9.2.2 Defnition and Discussion of Key Runof Elements .......................................................................................9-7
9.2.2.1 Watershed Boundary and Area .........................................................................................................................9-7
9.2.2.2 Precipitation ............................................................................................................................................................9-7
9.2.2.2.1 Rainfall Curves ....................................................................................................................................................9-9
9.2.2.2.2 Rainfall Intensity.................................................................................................................................................9-9
9.2.2.3 Watershed Characteristics ............................................................................................................................... 9-10
9.2.3 Key Storage and Outfow Elements ................................................................................................................. 9-11
9.2.3.1 Impoundment Capacity ................................................................................................................................... 9-11
9.2.3.2 Decants, Principal Spillways and Auxiliary Spillways ............................................................................. 9-11
9.3 GENERAL CONSIDERATIONS FOR COAL REFUSE DISPOSAL FACILITIES ............................................. 9-15
< PREVIOUS VIEW
xviii MAY 2009
Table of Contents
9.3.1 Special Characteristics .......................................................................................................................................... 9-15
9.3.2 Site Conditions ........................................................................................................................................................ 9-15
9.3.2.1 Topography........................................................................................................................................................... 9-16
9.3.2.1.1 Steep Terrain ..................................................................................................................................................... 9-16
9.3.2.1.2 Gently Sloping Terrain ................................................................................................................................... 9-16
9.3.2.1.3 Efects of Slope on Facility Staging .......................................................................................................... 9-17
9.3.2.2 Weather and Climate ......................................................................................................................................... 9-19
9.3.2.3 Geology .................................................................................................................................................................. 9-19
9.3.3 Construction Materials ......................................................................................................................................... 9-20
9.4 DESIGN CONSIDERATIONS FOR DISPOSAL FACILITY EMBANKMENT TYPES .................................... 9-20
9.4.1 Non-Impounding Embankments ..................................................................................................................... 9-20
9.4.1.1 Valley-Fill Embankments .................................................................................................................................. 9-21
9.4.1.2 Side-Hill, Ridge and Heaped Embankments ............................................................................................. 9-22
9.4.2 Slurry Cell Embankments .................................................................................................................................... 9-22
9.4.3 Slurry Impoundments ........................................................................................................................................... 9-24
9.4.3.1 Cross-Valley Impoundment ............................................................................................................................ 9-25
9.4.3.2 Diked Impoundment ......................................................................................................................................... 9-26
9.4.3.3 Incised Impoundment ...................................................................................................................................... 9-26
9.4.4 Other Impounding Structures ........................................................................................................................... 9-26
9.4.4.1 Sedimentation and Treatment Ponds ......................................................................................................... 9-28
9.4.4.2 Fresh Water Impoundments ........................................................................................................................... 9-29
9.5 DESIGN STORM CRITERIA ....................................................................................................................................... 9-29
9.5.1 Design Storms for Impoundments ................................................................................................................... 9-29
9.5.1.1 General Considerations .................................................................................................................................... 9-29
9.5.1.2 Recommended Design Storm Criteria ........................................................................................................ 9-32
9.5.1.3 Size and Hazard-Potential Classifcation .................................................................................................... 9-32
9.5.1.3.1 Impoundment-Size Classifcation ............................................................................................................. 9-33
9.5.1.3.2 Hazard-Potential Classifcation .................................................................................................................. 9-33
9.5.1.4 Determination of Design Storm Precipitation ............................................................................................. 9-34
9.5.1.4.1 Prediction of the PMP and PMTS ............................................................................................................... 9-34
9.5.1.4.2 Prediction of the 100-year and Lesser Design Storms....................................................................... 9-35
9.5.2 Special Considerations for Short-Term Conditions .................................................................................... 9-36
9.5.3 Hydraulic Design Criteria for Drainage Conveyance Installations ........................................................ 9-37
9.6 DETERMINATION OF RUNOFF QUANTITIES .................................................................................................... 9-38
9.6.1 Basic Hydrology Parameters ............................................................................................................................... 9-38
9.6.1.1 Precipitation Intensity-Duration and Distribution .................................................................................. 9-38
9.6.1.2 Unit Hydrographs and Time of Concentration ......................................................................................... 9-40
9.6.1.2.1 Snyder Unit Hydrograph .............................................................................................................................. 9-41
9.6.1.2.2 Clark Unit Hydrograph .................................................................................................................................. 9-42
9.6.1.2.3 SCS Dimensionless Unit Hydrograph ...................................................................................................... 9-42
9.6.1.3 Precipitation-Runof Relationship ................................................................................................................ 9-46
9.6.1.3.1 General Rainfall Conditions ........................................................................................................................ 9-46
< PREVIOUS VIEW
xix MAY 2009
Engineering and Design Manual Coal Refuse Disposal Facilities
9.6.1.3.2 Thunderstorm Rainfall Conditions ........................................................................................................... 9-49
9.6.1.4 Channel and Impoundment Storage Characteristics ............................................................................ 9-52
9.6.2 Runof Determination: Hydrograph Method............................................................................................... 9-54
9.6.3 Peak Runof Determination Key Parameters Method ............................................................................ 9-55
9.6.4 Runof Determination Rational Method ..................................................................................................... 9-58
9.7 DESIGN OF OUTFLOW SYSTEMS ......................................................................................................................... 9-60
9.7.1 Basic Considerations ............................................................................................................................................. 9-61
9.7.2 Hydraulic System Components ......................................................................................................................... 9-63
9.7.2.1 Inlets ........................................................................................................................................................................ 9-63
9.7.2.2 Transport Sections ............................................................................................................................................. 9-67
9.7.2.2.1 Open-Channel Flow ....................................................................................................................................... 9-67
9.7.2.2.2 Pressure Flow ................................................................................................................................................... 9-72
9.7.2.3 Outlets .................................................................................................................................................................... 9-81
9.7.3 Special Design Considerations .......................................................................................................................... 9-83
9.7.3.1 Channel Freeboard ............................................................................................................................................ 9-83
9.7.3.2 Erosion Protection .............................................................................................................................................. 9-83
9.7.3.2.1 Grass-Lined Channels .................................................................................................................................... 9-86
9.7.3.2.2 Riprap-Lined Channels ................................................................................................................................. 9-86
9.7.3.2.3 Grouted-Riprap Channels ............................................................................................................................ 9-90
9.7.3.2.4 Concrete-Lined Channels ............................................................................................................................ 9-92
9.7.3.2.5 Commercially Available Composite Erosion Control Products ...................................................... 9-93
9.7.3.3 Cavitation .............................................................................................................................................................. 9-93
9.7.3.4 Directional Changes in Open-Channel Flow ............................................................................................ 9-94
9.7.3.5 Directional Changes in Conduit Flow .......................................................................................................... 9-94
9.7.3.6 Materials Selection ............................................................................................................................................. 9-95
9.7.3.6.1 Open-Channel Lining Systems .................................................................................................................. 9-95
9.7.3.6.2 Conduit Materials ........................................................................................................................................... 9-95
9.7.4 Types of Hydraulic Systems................................................................................................................................. 9-95
9.7.4.1 Decant Systems ................................................................................................................................................... 9-96
9.7.4.1.1 Inlet ...................................................................................................................................................................... 9-96
9.7.4.1.2 Transport Section ........................................................................................................................................... 9-98
9.7.4.2 Overfow Spillway Systems ............................................................................................................................. 9-99
9.7.4.3 Diversion and Collection Ditches................................................................................................................9-110
9.7.4.4 Culverts ................................................................................................................................................................9-110
9.7.4.5 Natural Streams .................................................................................................................................................9-110
9.8 RESERVOIR ROUTING .............................................................................................................................................9-111
9.8.1 Basic Routing Methodology .............................................................................................................................9-111
9.8.2 Basic Routing Parameters ..................................................................................................................................9-112
9.9 DAM BREACH ANALYSIS AND INUNDATION MAPPING ...........................................................................9-113
9.9.1 Background ............................................................................................................................................................9-113
9.9.2 Failure Scenarios ...................................................................................................................................................9-113
9.9.3 Analytical Methods ..............................................................................................................................................9-115
< PREVIOUS VIEW
xx MAY 2009
Table of Contents
9.9.4 Input Data ...............................................................................................................................................................9-116
9.9.4.1 Breach Parameters ...........................................................................................................................................9-116
9.9.4.2 Initial Conditions ...............................................................................................................................................9-118
9.9.4.3 Flow Channel Geometry and Roughness ................................................................................................9-118
9.9.5 Results of Analysis/Inundation Mapping .....................................................................................................9-118
9.9.6 Sensitivity Analyses..............................................................................................................................................9-119
9.9.7 Hazard Classifcation ...........................................................................................................................................9-119
Chapter 10 Environmental Considerations
10.1 STREAMS AND WETLANDS .................................................................................................................................... 10-1
10.2 AIR QUALITY ................................................................................................................................................................ 10-3
10.2.1 Dust Control ............................................................................................................................................................. 10-3
10.2.2 Combustion Control .............................................................................................................................................. 10-4
10.2.3 Refuse Fire Extinguishment ................................................................................................................................ 10-5
10.2.3.1 Excavation and Removal .................................................................................................................................. 10-6
10.2.3.2 Surface Treatment .............................................................................................................................................. 10-8
10.2.3.3 Water and Slurry Injection ............................................................................................................................... 10-9
10.3 WATER QUALITY .......................................................................................................................................................10-10
10.3.1 Erosion and Sedimentation ..............................................................................................................................10-10
10.3.1.1 Prevention ...........................................................................................................................................................10-10
10.3.1.2 Control ..................................................................................................................................................................10-11
10.3.1.2.1 Sedimentation Ponds ..................................................................................................................................10-11
10.3.1.2.2 Sediment Traps and Check Dams ...........................................................................................................10-12
10.3.1.2.3 Silt Fences ........................................................................................................................................................10-12
10.3.1.2.4 Erosion Control Blankets and Reinforcement Mats ..........................................................................10-12
10.3.1.2.5 Vegetation .......................................................................................................................................................10-12
10.3.2 Acid Generation and Control ...........................................................................................................................10-12
10.3.2.1 Background.........................................................................................................................................................10-13
10.3.2.2 Grading, Compaction and Sealing .............................................................................................................10-13
10.3.2.3 Amendments .....................................................................................................................................................10-13
10.3.2.4 Reclamation and Vegetative Cover ............................................................................................................10-15
10.3.3 Water Quality Control .........................................................................................................................................10-15
10.3.3.1 Diversion of Runof ..........................................................................................................................................10-15
10.3.3.2 Treatment ............................................................................................................................................................10-15
10.3.4 Water Quality Impacts on Construction Materials ...................................................................................10-16
10.3.5 Hydrogeology ........................................................................................................................................................10-17
10.4 LINER SYSTEMS .........................................................................................................................................................10-19
10.4.1 Design Requirements ..........................................................................................................................................10-20
10.4.2 Stability ....................................................................................................................................................................10-23
10.4.3 Performance Considerations ............................................................................................................................10-24
10.5 RECLAMATION ..........................................................................................................................................................10-26
< PREVIOUS VIEW
xxi MAY 2009
Engineering and Design Manual Coal Refuse Disposal Facilities
10.5.1 Design Considerations ........................................................................................................................................10-26
10.5.2 Grading.....................................................................................................................................................................10-27
10.5.3 Impoundment Elimination ...............................................................................................................................10-28
10.5.4 Soil and Topsoil Cover .........................................................................................................................................10-29
10.5.5 Vegetation ...............................................................................................................................................................10-30
Chapter 11 Construction and Disposal Operations
11.1 OPERATIONAL PLANS AND CONSTRUCTION DOCUMENTS .................................................................... 11-1
11.1.1 Design Report .......................................................................................................................................................... 11-3
11.1.2 Construction Drawings ......................................................................................................................................... 11-4
11.1.2.1 Title Sheet and Existing Conditions Plans.................................................................................................. 11-4
11.1.2.2 Initial Site Development Plans ....................................................................................................................... 11-5
11.1.2.3 Refuse Embankment Construction Plan Views ........................................................................................ 11-6
11.1.2.4 Cross Sections and Profles.............................................................................................................................. 11-6
11.1.2.5 Detailed Plan Views and Cross Sections ..................................................................................................... 11-8
11.1.2.6 Miscellaneous Details and Information ...................................................................................................... 11-8
11.1.3 Technical Specifcations ....................................................................................................................................... 11-9
11.1.4 Operation and Maintenance Plan ...................................................................................................................11-11
11.1.5 Calculation Brief ....................................................................................................................................................11-13
11.2 PLANNING AND SCHEDULING ...........................................................................................................................11-14
11.2.1 Planning of Coal Refuse Disposal Operations ............................................................................................11-14
11.2.1.1 Long-Term Planning ........................................................................................................................................11-15
11.2.1.2 Short-Term Planning ........................................................................................................................................11-16
11.2.2 Scheduling Methods and Application ..........................................................................................................11-17
11.2.2.1 Scheduling Data ................................................................................................................................................11-17
11.2.2.2 Applicability of Scheduling Techniques ...................................................................................................11-18
11.3 REFUSE TRANSPORT ...............................................................................................................................................11-19
11.3.1 Trucks and Scrapers .............................................................................................................................................11-20
11.3.2 Conveyor Belt Systems .......................................................................................................................................11-20
11.4 HANDLING EQUIPMENT........................................................................................................................................11-21
11.4.1 Loading and Hauling Equipment ...................................................................................................................11-21
11.4.2 Spreading Equipment .........................................................................................................................................11-22
11.4.3 Compaction Equipment .....................................................................................................................................11-23
11.4.4 Use of Handling Equipment for Related Activities ...................................................................................11-26
11.5 CONSTRUCTION AND PLACEMENT OF EMBANKMENT MATERIALS ...................................................11-26
11.5.1 Embankment Fill ...................................................................................................................................................11-26
11.5.2 Upstream Construction Implementation Procedure ...............................................................................11-33
11.5.3 Excess Coarse Refuse/Inclement Weather Disposal .................................................................................11-35
11.5.4 Disposal of Fine Refuse Slurry ..........................................................................................................................11-36
11.5.5 Disposal of Combined Refuse ..........................................................................................................................11-38
11.5.6 Use of Amendments ............................................................................................................................................11-38
< PREVIOUS VIEW
xxii MAY 2009
Table of Contents
11.6 RELATED ACTIVITIES ...............................................................................................................................................11-38
11.6.1 Haul and Access Road Construction and Maintenance ..........................................................................11-39
11.6.2 Liner Systems .........................................................................................................................................................11-39
11.6.3 Embankment Foundation Preparation .........................................................................................................11-40
11.6.3.1 Clearing and Topsoil Stockpiling ................................................................................................................11-40
11.6.3.2 Soft Soil Foundations ......................................................................................................................................11-40
11.6.3.3 Competent Soil Foundations .......................................................................................................................11-40
11.6.3.4 Rock Foundations .............................................................................................................................................11-40
11.6.3.5 Mine Spoil ............................................................................................................................................................11-41
11.6.4 Water Control .........................................................................................................................................................11-41
11.6.5 General Excavation ..............................................................................................................................................11-43
11.6.6 Embankment Fill Constructed from On-Site Borrow Materials ...........................................................11-43
11.6.6.1 Clayey and Silty Borrow Materials ..............................................................................................................11-43
11.6.6.2 Rock and Granular Borrow Materials .........................................................................................................11-44
11.6.6.3 Use of Mine Spoil ..............................................................................................................................................11-45
11.7 FACILITY APPURTENANT STRUCTURES ..........................................................................................................11-46
11.7.1 Material Requirements .......................................................................................................................................11-46
11.7.2 Filters and Drains ..................................................................................................................................................11-46
11.7.2.1 Granular and Geotextile Filters ....................................................................................................................11-50
11.7.2.2 Granular Underdrains ......................................................................................................................................11-51
11.7.2.3 Drainage Pipe .....................................................................................................................................................11-53
11.7.3 Culverts and Decants ..........................................................................................................................................11-53
11.7.4 Concrete Structures .............................................................................................................................................11-54
11.7.5 Gates, Valves, Pipelines and Other Metal Work ..........................................................................................11-55
11.7.6 Support Structures ...............................................................................................................................................11-55
11.7.7 Mine-Opening and Auger-Hole Seals and Drains.....................................................................................11-56
11.8 QUALITY CONTROL AND FIELD TESTING .......................................................................................................11-56
11.8.1 Compaction Control ............................................................................................................................................11-56
11.8.2 Material Testing .....................................................................................................................................................11-58
Chapter 12 Monitoring, Inspections and
Facility Maintenance
12.1 MONITORING AND INSPECTIONS ....................................................................................................................... 12-2
12.1.1 Monitoring and Inspection Objectives ........................................................................................................... 12-3
12.1.2 Visual Observations ............................................................................................................................................... 12-4
12.1.2.1 Schedule and Checklist for Visual (and Other) Observations ............................................................. 12-7
12.1.2.2 Facility Development ........................................................................................................................................ 12-8
12.1.2.2.1 Site Preparation ............................................................................................................................................... 12-8
12.1.2.2.2 Starter Embankment and Internal Drains .............................................................................................. 12-8
12.1.2.2.3 Decant Pipes ...................................................................................................................................................12-12
12.1.2.2.4 Channel Construction .................................................................................................................................12-12
< PREVIOUS VIEW
xxiii MAY 2009
Engineering and Design Manual Coal Refuse Disposal Facilities
12.1.2.2.5 Instrumentation ............................................................................................................................................12-12
12.1.2.2.6 Reclamation ....................................................................................................................................................12-12
12.1.2.3 Adjoining Area Conditions ................................................................................................................................12-13
12.1.2.3.1 Sloughing or Sliding ....................................................................................................................................12-13
12.1.2.3.2 Signs of Subsidence .....................................................................................................................................12-13
12.1.2.3.3 Evidence of Erosion ......................................................................................................................................12-13
12.1.2.3.4 Springs or Seepage Areas ..........................................................................................................................12-13
12.1.2.3.5 Changes in Vegetation ................................................................................................................................12-14
12.1.2.4 Other Development Work .................................................................................................................................12-14
12.1.2.4.1 Upstream Development .............................................................................................................................12-14
12.1.2.4.2 Downstream Development ......................................................................................................................12-14
12.1.2.4.3 Disposal of Materials Other than Coal Refuse ....................................................................................12-14
12.1.2.5 Refuse Disposal Operations and Maintenance ..........................................................................................12-14
12.1.2.5.1 Refuse Placement and Compaction.......................................................................................................12-14
12.1.2.5.2 Working Surface Maintenance ................................................................................................................12-15
12.1.2.5.3 Embankment Crest Conditions ...............................................................................................................12-15
12.1.2.5.4 Embankment Slope Conditions ..............................................................................................................12-15
12.1.2.5.5 Liner, Foundation and Abutment Conditions ....................................................................................12-17
12.1.2.5.6 Internal Drains ...............................................................................................................................................12-17
12.1.2.5.7 Decant Pipe .....................................................................................................................................................12-17
12.1.2.5.8 Channel Conditions .....................................................................................................................................12-18
12.1.2.5.9 Instrumentation ............................................................................................................................................12-21
12.1.2.5.10 Reclaimed Slopes ..........................................................................................................................................12-21
12.1.3 Quality Control Sampling and Testing ..........................................................................................................12-21
12.2 MONITORING INSTRUMENTATION ...................................................................................................................12-22
12.3 FACILITY AND DAM SAFETY INSPECTIONS ...................................................................................................12-24
12.3.1 Routine Inspections .............................................................................................................................................12-24
12.3.1.1 Impoundments and Dams ............................................................................................................................12-24
12.3.1.2 Non-Impounding Facilities............................................................................................................................12-27
12.3.2 Event-Related Inspections .................................................................................................................................12-27
12.3.3 Annual Dam Safety Inspections ......................................................................................................................12-27
12.4 CERTIFICATION AND CONFORMANCE REPORTING...................................................................................12-28
12.5 MAINTENANCE .........................................................................................................................................................12-29
12.5.1 Routine Maintenance ..........................................................................................................................................12-29
12.5.2 Maintenance after Unusual Events ................................................................................................................12-30
12.5.3 Long-Term Maintenance ...................................................................................................................................12-30
Chapter 13 Instrumentation and
Performance Monitoring
13.1 INSTRUMENTATION PROGRAM PLANNING .................................................................................................... 13-1
13.2 MEASUREMENT TECHNIQUES .............................................................................................................................. 13-2
< PREVIOUS VIEW
xxiv MAY 2009
Table of Contents
13.2.1 Movements and Displacements ....................................................................................................................... 13-3
13.2.1.1 General Observations ....................................................................................................................................... 13-3
13.2.1.2 Movement Measurement Techniques ........................................................................................................ 13-8
13.2.1.2.1 Surveying Methods ........................................................................................................................................ 13-8
13.2.1.2.2 Surface Extensometers ...............................................................................................................................13-16
13.2.1.2.3 Tiltmeters .........................................................................................................................................................13-16
13.2.1.2.4 Probe Extensometers ..................................................................................................................................13-18
13.2.1.2.5 Fixed Embankment Extensometers .......................................................................................................13-19
13.2.1.2.6 Fixed Borehole Extensometers ................................................................................................................13-20
13.2.1.2.7 Inclinometers .................................................................................................................................................13-21
13.2.2 Piezometric Levels and Pore-Water Pressures ............................................................................................13-23
13.2.2.1 Piezometer Types ..............................................................................................................................................13-23
13.2.2.2 Observation Wells .............................................................................................................................................13-24
13.2.2.3 Open-Standpipe Piezometers ......................................................................................................................13-25
13.2.2.4 Closed-System Piezometers .........................................................................................................................13-29
13.2.3 Surface Water Flows and Hydrologic Parameters .....................................................................................13-31
13.2.3.1 Measurement of Impoundment Water Level .........................................................................................13-31
13.2.3.2 Measurement of Seepage Flows and Other Surface Water Flows ..................................................13-32
13.2.3.2.1 Importance of Seepage Flows .................................................................................................................13-32
13.2.3.2.2 Measurement of Seepage Flow Rates ...................................................................................................13-34
13.2.3.3 Measurement of Rainfall and Snowfall .....................................................................................................13-36
13.2.3.4 Observation of Temperature and General Weather Conditions ......................................................13-36
13.2.4 Miscellaneous Instrumentation and Monitoring ......................................................................................13-36
13.2.4.1 Measurement of Soil Pressure ......................................................................................................................13-36
13.2.4.2 Measurement of Blast Vibrations ................................................................................................................13-39
13.2.4.3 Measurement of Temperature .....................................................................................................................13-41
13.3 INSTRUMENTATION MAINTENANCE ................................................................................................................13-41
13.3.1 Importance of Maintenance and Recalibration .........................................................................................13-41
13.3.2 Recalibration and Maintenance during Service Life ................................................................................13-42
13.4 AUTOMATED DATA ACQUISITION .....................................................................................................................13-42
13.5 INSTRUMENTATION COSTS ..................................................................................................................................13-42
APPENDIX 13A MEASUREMENTS, TRANDUCERS AND DATA-ACQUISITION SYSTEMS
Chapter 14 Emergency Action Planning
14.1 PURPOSE ....................................................................................................................................................................... 14-1
14.2 EMERGENCY ACTION PLAN PREPARATION ..................................................................................................... 14-1
14.2.1 Background ............................................................................................................................................................. 14-1
14.2.2 Federal Guidelines for Dam Safety .................................................................................................................. 14-3
14.2.3 Basic Elements of an EAP .................................................................................................................................... 14-3
14.2.4 Dam Inspection and Inundation Area Reconnaissance .......................................................................... 14-4
14.2.4.1 Dam Inspection .................................................................................................................................................. 14-4
< PREVIOUS VIEW
xxv MAY 2009
Engineering and Design Manual Coal Refuse Disposal Facilities
14.2.4.2 Downstream Channel Reconnaissance ..................................................................................................... 14-5
14.2.4.3 Physical Description ......................................................................................................................................... 14-5
14.2.5 Inundation Map Preparation ............................................................................................................................. 14-5
14.2.5.1 Causes of Dam Failures .................................................................................................................................... 14-5
14.2.5.2 Dam Breach Analysis ........................................................................................................................................ 14-6
14.2.5.3 Inundation Map ................................................................................................................................................. 14-6
14.2.6 Participant Responsibilities ................................................................................................................................ 14-6
14.2.7 Implementation Triggers and Emergency Responses ............................................................................. 14-7
14.2.8 Notifcation Requirements ................................................................................................................................. 14-9
14.2.9 EAP Public Copies .................................................................................................................................................. 14-9
14.2.10 Exercises/Training .................................................................................................................................................... 14-9
14.2.11 Plan Maintenance ..................................................................................................................................................14-10
14.2.12 EAP Appendices .....................................................................................................................................................14-11
14.3 PREPARATION FOR POTENTIAL EMERGENCY REMEDIAL ACTIONS ....................................................14-11
14.3.1 Potential Overtopping During a Storm .......................................................................................................14-12
14.3.2 Movement or Sliding of the Embankment .................................................................................................14-12
14.3.3 Internal Erosion or Piping through the Dam, Foundation or Abutments .......................................14-12
14.3.4 Excessive Seepage and a High Level of Saturation in the Embankment ........................................14-13
14.3.5 Spillway Failure or Erosion that Could Cause a Dam Breach................................................................14-13
14.3.6 Excessive Settlement or Subsidence of the Embankment ..................................................................14-13
14.3.7 Formation of a Sinkhole on the Crest or Face of the Dam ....................................................................14-14
14.3.8 Earthquake .............................................................................................................................................................14-14
< PREVIOUS VIEW
xxvi MAY 2009
Table of Contents
< PREVIOUS VIEW
List of Tables
xxvii MAY 2009
TABLE
NO.
TITLE
PAGE
NO.
2.1 Classifcation of Coals by Rank 2-3
2.2 Coal Refuse Disposal Facility Failures and Incidents 2-13
3.1 Guidance for Flood Warning Systems at Surface Mine Pits 3-24
4.1 Coal Refuse Disposal Facility Cost Components 4-5
4.2 Planning Steps for Coal Refuse Facility Design 4-8
4.3 Typical Design Sequence 4-9
6.1 Typical Geotechnical Design Procedure for Coal Refuse Disposal Facilities 6-3
6.2 Basic Characteristics of Coal Refuse Disposal and Design Signifcance 6-5
6.3 Coarse Coal Refuse Characterization Summary of Average/Range of Values 6-9
6.4 Fine Coal Refuse Characterization Summary of Average/Range of Values 6-11
6.5 Combined Coal Refuse Characterization 6-12
6.6 Correlation Between USCS Classifcation and Properties of Compacted Soils 6-14
6.7 Correlation between USCS Classifcation and Relative Desirability of Soils
as Compacted Fill
6-15
6.8 Approximate Correlation Between Engineering Properties and Soil Classifcation
Groups
6-16
6.9 Class F Fly Ash 6-17
6.10 Class C Fly Ash 6-17
6.11 Fly Ash/Coal Refuse Mixture Properties 6-18
6.12 Advantages and Disadvantages of Internal Drain Types 6-23
6.13 Relative Advantages of Vertical and Sloping Embankment Cores 6-25
6.14 Clay Dispersion Potential 6-28
6.15 Guideline Subsurface Exploration Program for a New Impounding Embankment 6-54
6.16 Guideline Subsurface Exploration Program for a New Embankment without an
Impoundment
6-55
6.17 Guideline Subsurface Exploration Program for An Existing Embankment 6-56
xxviii MAY 2009
List of Tables
TABLE
NO.
TITLE
PAGE
NO.
6.18 Common Sampling Methods 6-57
6.19 Advantages and Limitations of the Standard Penetration Test 6-59
6.20 Advantages and Limitations of the Cone Penetrometer Test 6-62
6.21 Advantages and Limitations of the Field Vane Shear Test 6-65
6.22 Field Methods for Measurement of Hydraulic Conductivity 6-68
6.23 Advantages and Limitations of Surface Geophysical Testing 6-71
6.24 Surface Geophysical Methods in Relation to Typical Investigation Objectives 6-72
6.25 Physical Properties and Typical Ground Penetrating Radar Velocities for Common
Earth Materials
6-82
6.26 Applicability of Common Borehole Geophysical Methods 6-87
6.27 Typical Soil and Rock Laboratory Tests for Coal Refuse Disposal Site
Characterization
6-92
6.28 Typical Laboratory Soil Tests for Various Materials 6-94
6.29 Guidelines for Minimum Sample Weights 6-96
6.30 Soil Classifcation Chart (Laboratory Method) 6-98
6.31 In-Place Unit Weight and Specifc Gravity of Coarse Coal Refuse 6-100
6.32 Unit Weight and Specifc Gravity for Fine Coal Refuse 6-101
6.33 Statistical Properties for Atterberg Limits and Moisture Contents of
Fine Coal Refuse at Northern Appalachian Sites
6-103
6.34 Particle-Size Distribution for Fine Coal Refuse 6-105
6.35 Test Methods for Chemical Characterization of Soil, Rock and Refuse Materials 6-106
6.36 Estimated Horizontal Hydraulic Conductivity of Fine Coal Refuse
Based on Piezocone Dissipation Tests
6-111
6.37 Methods for Quality Control Testing of Geotextiles 6-113
6.38 Methods for Quality Control Testing of Geomembranes 6-114
6.39 Methods for Quality Control Testing of Geosynthetic Clay Liners 6-115
6.40 Laboratory Tests for Determining Soil Shear Strength 6-121
6.41 Summary of Fine Coal Refuse Shear Strength Parameters Based on
Triaxial Test Results
6-127
6.42 Standards and Procedures for Laboratory Testing of Intact Rock 6-131
6.43 Slake Durability Classifcation 6-136
6.44 Quality Control Standards for Portland-Cement Concrete 6-136
6.45 Quality Control Standards for CLSM 6-136
6.46 CLSM Mix Design Guide 6-137
< PREVIOUS VIEW
xxix MAY 2009
Engineering and Design Manual Coal Refuse Disposal Facilities
TABLE
NO.
TITLE
PAGE
NO.
6.47 Filter Criteria As A Function of Base Soil and Percent Passing No. 200 Sieve 6-148
6.48 Restrictions on Filter Particle Size to Limit Segregation 6-148
6.49 Guidelines for Evaluating Critical Nature of Severity for Drainage Applications 6-153
6.50 Geotextile Strength Property Requirements for Drainage Geotextiles 6-155
6.51 Recommended Minimum Properties for General Geomembrane Installation
Survivability
6-163
6.52 Stress-Strain Relationships Used for Finite Element Deformation Analyses
of Embankments and Foundations
6-169
6.53 Characteristics of Equilibrium Procedures for Slope Stability Analysis 6-174
6.54 Shear Strengths, Water Pressures, and Unit Weights for Slope-Stability Analyses 6-177
6.55 Guideline Friction Values and Efciencies for Various Geosynthetic and Soil
Combinations
6-178
6.56 Recommended Minimum Factors of Safety for Design of Coal Refuse
Embankments
6-183
6.57 Common Residential Velocity Criteria and Efects 6-207
7.1 Comparison of Basic Criteria for Sand-Like, Clay-Like and Borderline Materials 7-14
7.2 References for Correlations of S
US
with SPT and CPT Data 7-36
7.3 Correlations of S
US
/
V
versus SPT and CPT Data Back-Calculated from Case Histories 7-38
7.4 Upper-Bound Values of S
US
for Corresponding Values of N
1,60
7-45
7.5 Error in Void Ratio and Corresponding Percent Error in S
US
7-49
7.6 MDE and Seismic Hazard Assessment Recommendations 7-72
7.7 Safe Shutdown Earthquake (SSE) Peak Horizontal Ground Accelerations
for U.S. Nuclear Power Plants
7-85
8.1 Sources of Information for Evaluation of Breakthrough Potential 8-4
8.2 Sources for Mine Maps 8-5
8.3 Exploration Considerations for Assessing Impoundment Breakthrough Potential To
Underground Mines
8-8
8.4 Outcrop Barrier Exploration Considerations 8-9
8.5 Summary of Guidelines for Mining Under Or Near Bodies of Water 8-11
8.6 Guidelines for Cement Grout Programs 8-63
9.1 Hydrologic and Hydraulic Design Procedures for Coal Refuse Disposal Facilities 9-3
9.2 Hydrologic and Hydraulic Design Considerations for Coal Refuse Disposal Facilities 9-6
9.3 Facility Characteristics Infuencing Hydraulic System Design 9-15
9.4 Recommended Minimum Design Storm Criteria for Coal Refuse Disposal
Impoundments
9-30
< PREVIOUS VIEW
xxx MAY 2009
List of Tables
TABLE
NO.
TITLE
PAGE
NO.
9.5 NWS Precipitation Frequency Publications 9-37
9.6 Typical Design Criteria for Minor Hydraulic Structures 9-38
9.7 Methods for Determining Runof Rate and Volume 9-39
9.8 Recommended Correction Factors for T
c
for Watersheds West of the 105
th
Meridian 9-43
9.9 Roughness Coefcients (Mannings n) for Sheet Flow 9-44
9.10 Runof Curve Numbers for Watershed Complexes and AMC-II 9-50
9.11 Runof Curve Numbers (CN) for Antecedent Moisture Conditions 9-52
9.12 Runof Coefcients for Use in the Rational Method 9-60
9.13 Mannings Coefcients of Channel Roughness 9-72
9.14 Head Loss Coefcients for Conduits 9-77
9.15 References for Secondary Hydraulic Design Issues 9-81
9.16 Permissible Velocity and Tractive Force for Channel Linings 9-85
9.17 Vegetal Retardance Classes 9-87
9.18 Riprap Size Designation 9-89
9.19 Recommended Grading of Grouted Rock Riprap Lining 9-91
9.20 Inundation Analysis Software 9-115
9.21 Hazard Potential Classifcations 9-119
10.1 Fire Extinguishment Techniques 10-7
10.2 Summary of Suggested Criteria for Interpreting Acid-Base Accounting 10-14
10.3 Chemical Compounds Used in AMD Treatment 10-16
10.4 Design Considerations for AMD Passive Treatment Systems 10-17
10.5 Overview of the Hydrogeologic Process 10-18
10.6 Site Specifc Factors to Consider in Liner System Design 10-19
10.7 Available Materials or Procedures for Liner System Components 10-21
10.8 Stability Considerations in Liner Design 10-24
10.9 Liner System Performance Factors and Considerations 10-25
10.10 Potential Final Land Uses 10-26
10.11 Design Considerations for Reclamation Soils 10-27
10.12 Summary of Reclamation Guidelines 10-30
10.13 Characteristics of Common Species Potentially Suitable for Reclamation 10-31
11.1 Refuse Disposal Operations 11-2
11.2 Transport Systems and Associated Handling Equipment at the Point of Disposal 11-22
< PREVIOUS VIEW
xxxi MAY 2009
Engineering and Design Manual Coal Refuse Disposal Facilities
TABLE
NO.
TITLE
PAGE
NO.
11.3 Compaction Capabilities of Transport and Spreading Equipment 11-24
11.4 Compaction Guide for Material Types 11-28
11.5 Suggested Compaction Equipment and Methods 11-30
11.6 Efects of Corrosion and Countermeasures 11-47
12.1 Typical Observation and Inspection Checklist 12-9
12.2 Field Quality Control Testing Summary 12-25
13.1 Instrumentation Planning and Design Steps 13-3
13.2 Performance Questions for Slurry Impoundments 13-4
13.3 Summary of General Methods for Detecting Embankment Movements
and Displacements
13-6
13.4 Instrument Categories for Measuring Movements 13-9
13.5 Surveying Methods 13-12
13.6 Instruments for Measuring Ground-Water Pressure 13-24
13.7 Cement-Bentonite Grout Mix 13-26
13.8 Guidance for Accurate Weir Flow Measurement 13-39
14.1 Historical U.S. Dam Failures 14-2
14.2 Guidance for EAP Events and Situations 14-8
< PREVIOUS VIEW
xxxii MAY 2009
List of Tables
< PREVIOUS VIEW
List of Figures
xxxiii MAY 2009
FIGURE
NO.
TITLE
PAGE
NO.
2.1 Coarse Coal Refuse Embankment and Sample Gradation 2-8
2.2 Fine Coal Refuse Slurry Characteristics 2-9
2.3 Combined Coal Refuse Materials 2-10
3.1 Cross-Valley Impounding Embankment 3-5
3.2 Incised Impounding Facility 3-6
3.3 Side-Hill Impounding Embankment 3-7
3.4 Diked-Pond (Impounding) Embankment 3-8
3.5 Upstream Staging Method 3-8
3.6 Downstream Staging Method 3-9
3.7 Centerline Staging Method 3-9
3.8 Valley-Fill, Non-Impounding Embankment 3-12
3.9 Cross-Valley, Non-Impounding Embankment 3-12
3.10 Side-Hill, Non-Impounding Embankment 3-13
3.11 Ridge (Non-Impounding) Embankment 3-14
3.12 Heaped (Non-Impounding) Embankment 3-15
3.13 Slurry Cell in Valley Fill, Non-Impounding Embankment 3-16
3.14 Slurry Cell, Side-Hill, Non-Impounding Embankment 3-17
3.15 Slurry Cells in Cross-Valley Fill, Impounding Embankment 3-19
5.1 Typical Slurry Impoundment Components 5-4
5.2 Typical Slurry Cell Facility (Impounding Embankment) 5-5
5.3 Typical Refuse Embankment Components 5-6
5.4 Illustration of Impoundment Design Storm Control 5-10
5.5 Underground Mining at Slurry Impoundment 5-16
6.1 Drain Confgurations for Homogeneous Embankments with Impoundments 6-22
xxxiv MAY 2009
List of Figures
FIGURE
NO.
TITLE
PAGE
NO.
6.2 Zoned Embankments with Impoundments 6-24
6.3 Resistance of Core Materials to Piping and Cracking 6-27
6.4 Seepage Cutofs for Pervious Foundations 6-29
6.5 Impoundment Internal Drain Concept 6-30
6.6 Non-Impounding Embankments 6-32
6.7 Example Spring Collection Zones and Rock Underdrain 6-33
6.8 Upstream Construction Following Initial Development 6-35
6.9 Downstream Embankment Construction Followed by Expansion Using the
Downstream and Upstream Methods
6-38
6.10 Internal Drainage Systems for Existing Impounding Embankments 6-42
6.11 Flow Chart for Site Exploration, Material Property Testing, and Facility Design 6-45
6.12 Piezocone Penetrometer 6-62
6.13 Field Vane Shear Test Setup 6-64
6.14 Vane Correction Factor 6-66
6.15 Typical Seismic Refraction Data with Interpretation 6-73
6.16 Correlation of Rippability with P-Wave Velocity for Caterpillar D-9 6-74
6.17 Seismic Refection Principle and Schematic of Refection Data Record 6-75
6.18 Seismic Refection Survey Profle Over Abandoned Coal Mine Workings 6-76
6.19 Commonly Used Electrode Confgurations for Ground Electrical Measurements 6-77
6.20 Resistivity Profle Across Fine Coal Refuse Impoundment 6-77
6.21 Electrical Survey of Flooded Mine Workings 6-78
6.22 Illustration of Use of SP Method to Identify Area of Seepage 6-79
6.23 Delineation of Flooded Mine Workings with TDEM Method 6-81
6.24 Resistivity of Coal Refuse from EM-31 Measurements 6-81
6.25 GPR Record of Shallow Coal Mine Workings 6-83
6.26 Theoretical Response of Gravity Geophysical Method Over 20-Foot-Diameter, Air-Filled
Tunnel
6-84
6.27 Magnetic Intensity Recorded at Various Heights Above Well Casing 6-86
6.28 Field Setup for Crosshole Seismic Survey 6-88
6.29 Field Setup for Downhole Seismic Survey 6-89
6.30 Image of Mine Entry From Borehole Laser Scanner 6-90
6.31 Laser Image of Mine Workings 6-91
6.32 Plasticity Chart 6-99
< PREVIOUS VIEW
xxxv MAY 2009
Engineering and Design Manual Coal Refuse Disposal Facilities
FIGURE
NO.
TITLE
PAGE
NO.
6.33 Efect of Carbon Content on Specifc Gravity of Coal Refuse Material 6-102
6.34 Particle Size Distribution 6-103
6.35 Typical Moisture-Density Relationships for Dynamic Compaction 6-107
6.36 Particle-Size Distribution of Coarse Coal Refuse After Compaction 6-108
6.37 Results of Consolidation Test on Fine Coal Refuse 6-116
6.38 Efective Stress Mohrs Circle for CU Triaxial Test 6-118
6.39 Efective Stress Paths and Failure Envelope for CU Triaxial Test 6-119
6.40 Direct Shear Testing Arrangement 6-122
6.41 Results of Drained Direct Shear Test on Clay 6-123
6.42 Unconfned Compression Test Results 6-124
6.43 Variation of Strength of Fine Coal Refuse with Fines Content 6-127
6.44 Typical Results for Resonant Column Test on Fine Coal Refuse 6-130
6.45 Point-Load Test Apparatus 6-132
6.46 Rotating Drum Assembly and Setup of Slake Durability Equipment 6-134
6.47 Approximation for Estimated Reduction in Hydraulic Conductivity Under
Turbulent Flow
6-142
6.48 Flow Net Analysis 6-144
6.49 Flow Net Analysis for Anisotropic Conditions 6-145
6.50 Filter Diaphragm Design for Conduit 6-158
6.51 Example Design Confguration for Filter Diaphragm and Decant Pipe 6-159
6.52 Typical Relief Well Installation 6-161
6.53 Example Horizontal Drain Installation 6-162
6.54 Cracking Patterns Observed for Dams in Narrow Valleys 6-167
6.55 Histogram of Factor of Safety Distribution 6-180
6.56 Horizontal and Vertical Extent of Cuts for Base Width of 10 Feet 6-186
6.57 Schematic Geological Profle of A Bedrock Slope 6-187
6.58 Unfavorable Orientation of Bedding Planes 6-188
6.59 Favorable Orientation of Bedding Planes 6-190
6.60 Stepped Slope Example for Cuts in Weathered Rock 6-191
6.61 Examples of Rock-Cut Control 6-192
6.62 Deformations Associated With Conduits In Embankments 6-198
6.63 Reinforced Concrete Cylinder Pipe (RCCP) Details 6-204
6.64 Prestressed Concrete Cylinder Pipe (PCCP) Details (Lined Cylinder) 6-205
< PREVIOUS VIEW
xxxvi MAY 2009
List of Figures
FIGURE
NO.
TITLE
PAGE
NO.
6.65 Comparison of Predicted and Observed Conduit Settlements 6-206
6.66 Acceptable Limits of Vibration for Houses 6-208
7.1a Seismic Stability Screening 7-6
7.1b Triggering and Seismic-Stability Analysis 7-7
7.1c Deformation Screening and Analysis 7-8
7.2 Typical Undrained Stress-Strain Curve 7-11
7.3 Comparison of Peak Base and Crest Transverse Accelerations Measured at Earth Dams 7-27
7.4 Variation of Maximum Acceleration Ratio with Depth of Sliding Mass 7-28
7.5 SPT Clean-Sand-Based Curve for Magnitude 7.5 Earthquakes with Data from
Liquefaction Case Histories
7-30
7.6 Calculation of CRR from CPT Data Along with Empirical Liquefaction Data from
Compiled Case Histories
7-31
7.7 S
us
versus N
1,60(CS)
7-37
7.8 S
us
versus N
1,60
7-38
7.9 S
us
/
v
versus q
t1
7-39
7.10 Steady-State Line 7-46
7.11 Method for Correcting Laboratory S
US
to In-Situ S
US
7-48
7.12 Interpretation of Laboratory S
US
Testing 7-49
7.13 Consolidation Curve and Steady-State Line 7-50
7.14 Use of CPT Data to Screen Clay-Like Material 7-53
7.15 Cyclic Stress-Strain Curve Demonstrating Racheting Mechanism 7-59
7.16 Double Integration to Obtain Displacement 7-64
7.17 Location of U.S. Coal Resources 7-67
7.18 Seismic Hazard Map of the U.S. with Overlay of Main Bituminous Coal Fields 7-68
7.19 Illustration of Moderate- to High-Hazard Region 7-69
7.20 Quaternary Faults and Signifcant Seismic Sources with Overlay of Main Bituminous
Coal Fields
7-70
7.21 Return Period for M > 5.5 Event at a Distance Less than 50 km 7-80
7.22 Location of Nuclear Power Plants with Respect to U.S. Coal Fields 7-83
7.23 Seismic Hazard Assessment Flow Diagram 7-87
7.24 Example 1: Time History for Low-Seismic-Hazard Area 7-89
7.25 Example 2: Time History for Low-Seismic-Hazard Area 7-90
7.26 Example 3: Time History for Low-Seismic-Hazard Area 7-91
< PREVIOUS VIEW
xxxvii MAY 2009
Engineering and Design Manual Coal Refuse Disposal Facilities
FIGURE
NO.
TITLE
PAGE
NO.
8.1 Strata Disturbance and Subsidence Caused By Mining 8-12
8.2 Infuence of Extraction Width on Subsidence 8-12
8.3 Displacement and Strain for Various Extraction Widths 8-13
8.4 Development of Subsidence Through and Strains with Face Advance 8-14
8.5 Ground Movements Caused By Subsidence 8-15
8.6 Safety Zone Guidelines for Mining in the Vicinity of Dams and Impoundments 8-17
8.7 Hillside Disturbance at Coal Outcrop 8-21
8.8 Comparison of Pillar Strength Formulas with Respect to Width-to-Height Ratio 8-26
8.9 Outcrop Barrier Breakthrough Analysis for Shear Failure 8-32
8.10 Illustration of Overburden Barrier 8-34
8.11 Stowed Mine Openings and Barrier Construction 8-37
8.12 Barrier Against Outcrop with Low-Permeability Fill and Drain 8-38
8.13 Basic Bulkhead Design Concepts 8-44
8.14 Drift Opening Bulkhead Design Concepts 8-45
8.15 Bulkhead Confguration Showing Concrete-Block Retaining Walls and Concrete Center 8-46
8.16 Classifcation and Interpretation of Spoil Durability 8-58
8.17 Excavation Slopes for Preparation of Dam Abutments in Rock 8-60
8.18 Slope Modifcation to Reduce Diferential Settlement and Potential for Cracking 8-61
8.19 Slope Modifcation and Seam Treatment for Sedimentary Rock Strata 8-62
9.1 Runof and Impoundment Infow Sources 9-8
9.2 Watershed Boundary Delineated on USGS Topographic Map 9-9
9.3 Typical Impoundment Infow and Outfow Hydrographs 9-10
9.4 Typical Impoundment Area and Storage Volume Curves 9-12
9.5 Typical Decant Systems 9-13
9.6 Typical Decant and Spillway Systems 9-14
9.7 Channel Construction in Moderately and Steeply Sloping Terrain 9-17
9.8 Channel Construction in Gently Sloping Terrain 9-18
9.9 Drainage Control for Valley-Fill, Non-Impounding Embankment 9-21
9.10 Drainage Control for Side-Hill and Heaped Embankments 9-23
9.11 Drainage Control for Slurry-Cell Facility 9-25
9.12 Drainage Control for Cross-Valley Impoundment 9-27
9.13 Drainage Control for Diked Impoundment 9-28
< PREVIOUS VIEW
xxxviii MAY 2009
List of Figures
FIGURE
NO.
TITLE
PAGE
NO.
9.14 U.S. Regions Covered by Generalized PMP Studies 9-35
9.15 Dimensionless Design Storm Distribution 9-40
9.16 Average Velocities for Estimating Travel Time for Shallow, Concentrated Flow 9-45
9.17 Direct Runof for Rainfall Less than or Equal to 8 Inches 9-47
9.18 Direct Runof for Rainfall Greater than or Equal to 8 Inches 9-48
9.19 Estimated Total Runof Volume for CN = 70 9-56
9.20 Estimated Total Runof Volume for CN = 80 9-57
9.21 Estimated Peak Runof Rate for 6-Hour Design Storm and CN = 70 9-58
9.22 Estimated Peak Runof Rate for 6-Hour Design Storm and CN = 80 9-59
9.23 Example of Rational Method of Infow Calculation 9-62
9.24 Typical Head-Discharge Curves for Hydraulic Systems 9-64
9.25 Spillway Approach Channels 9-65
9.26 Drop-Inlet Decant Systems 9-66
9.27 Open-Channel Cross Sections 9-68
9.28 Channel Flow Conditions 9-70
9.29 Critical Depth for Trapezoidal and Circular Sections 9-71
9.30 Closed-Conduit Hydraulic Systems 9-73
9.31 Typical Flow Conditions Culvert Conduits on Mild Slopes 9-74
9.32 Typical Flow Conditions Culvert Conduits on Steep Slopes 9-75
9.33 Culvert Inlet Confgurations 9-78
9.34 Flow and Discharge Characteristics of Decant-Inlet, Closed-Conduit System 9-80
9.35 Examples of Channel and Conduit Outlets 9-82
9.36 Mannings n Versus VR for Various Retardance Classes 9-88
9.37 Grouted Riprap Lining Thickness as a Function of Flow Velocity 9-91
9.38 Examples of Decants 9-100
9.39 Example Decant System at Diked Impoundment 9-101
9.40 Example Decant System at Cross-Valley Impoundment 9-102
9.41 Example Spillway System A 9-104
9.42 Example Spillway System B 9-105
9.43 Example Spillway System C 9-106
9.44 Spillway Inlet Control 9-107
9.45 Spillway Outlet Control 9-108
< PREVIOUS VIEW
xxxix MAY 2009
Engineering and Design Manual Coal Refuse Disposal Facilities
FIGURE
NO.
TITLE
PAGE
NO.
9.46 Sectional View Through Impounding Facility 9-112
9.47 Breach Parameters 9-117
10.1 Liner Systems Used at Coal Refuse Disposal Facilities 10-23
11.1 Compaction by Hauling Units Using Successive Ofset Passes 11-25
11.2 Procedures for Controlled Placement of Refuse 11-27
11.3 Benching of Embankment Fill Into Natural Slope 11-33
11.4 Slurry Disposal 11-37
11.5 Stockpiling and Handling of Granular Materials 11-52
12.1 Failure Modes, Warning Signs and Surveillance Measures 12-5
12.2 Example Data Collection Form 12-6
13.1 Location of Monuments Using Triangulation 13-14
13.2 Example of Benchmark in Rock 13-15
13.3 Typical Monument Installations 13-17
13.4 Probe Extensometer 13-19
13.5 Settlement Platform 13-20
13.6 Fixed Borehole Extensometer 13-21
13.7 Inclinometer Operation 13-22
13.8 Approximate Response Time for Various Types of Piezometers 13-25
13.9 Open-Standpipe Piezometer/Observation Well 13-27
13.10 Normally-Closed Pneumatic Piezometer 13-30
13.11 Vibrating-Wire Piezometer 13-31
13.12 Bonded Electrical-Resistance Piezometer 13-33
13.13 Measurement of Flow from Pipes 13-37
13.14 Typical V-Notch Weir 13-38
13.15 Flow Depth-Discharge Relationship for 90-Degree, V-Notch Weir 13-40
13.16 Automated Data Acquisition System at Coal Refuse Disposal Facility 13-43
14.1 Inundation Map for Coal Refuse Impoundment 14-7
< PREVIOUS VIEW
xl MAY 2009
List of Figures
< PREVIOUS VIEW
Acronyms
MAY 2009
ACRONYM REPRESENTS
AASHTO American Association of State Highway and Transportation Ofcials
ABS Acrylonitrile-Butadiene-Styrene
ACI American Concrete Institute
AC alternating current
ALD anoxic limestone drain
AMC antecedent moisture condition
AMD acid mine drainage
AMRL AASHTO Materials Reference Laboratory
AOS apparent opening size
ArcGIS GIS software developed by Environmental Systems Research Institute (ESRI)
ARMPS Analysis of Retreat Mining Pillar Stability (software)
ARMPS-HWM ARMPS program for highwall mine pillars (software)
ASCE American Society of Civil Engineers
ASTM American Society for Testing and Materials
AWWA American Water Works Association
BOD biochemical oxygen demand
BREACH breach parameter computation software
CADD computer-aided design and drafting
CANDE Culvert Analysis and Design (software)
CD consolidated drained
CGS Canadian Geotechnical Society
CIDC consolidated isotropic drained compression
CISPM Comprehensive and Integrated Subsidence Prediction Model (software)
CIUC consolidated isotropic undrained compression
CLSM controlled low-strength material
CMP corrugated metal pipe
CMRR Coal Mine Roof Rating (software)
xli
< PREVIOUS VIEW
xlii
Acronyms
MAY 2009
ACRONYM REPRESENTS
CN curve number
CO carbon monoxide
COSMOS Consortium of Organizations for Strong-Motion Observation Systems
COV coefcient of variability
CPE chlorinated polyethylene
CPM Critical Path Method (software)
CPP corrugated plastic pipe
CPT cone penetrometer test
CPTu piezocone penetrometer test
CRR cyclic resistance ratio
CFR Code of Federal Regulations
CSI Construction Specifcations Institute
CSPE chlorosulfonated polyethylene
CSR cyclic stress ratio
CU consolidated undrained
CU
consolidated undrained with pore pressure measurements
CWA Clean Water Act
DAMBRK Dam Break (software)
DC direct current
DCDT direct current diferential transformer
DEM digital elevation model
DEP Department of Environmental Protection (state environmental agencies)
DI degradation index
DMLR Virginia Department of Mines, Minerals, and Energy, Division of Mined Land
Reclamation
DOD Department of Defense
DOQ digital orthophoto quadrangle
DOQQ digital orthophoto quarter quadrangle
DRG digital raster grid
DSHA deterministic seismic hazard analysis
DWOPER Dynamic Wave Routing Model (software)
EAP Emergency Action Plan
EDM electronic distance measuring
< PREVIOUS VIEW
xliii
Engineering and Design Manual Coal Refuse Disposal Facilities
MAY 2009
ACRONYM REPRESENTS
EEGS Environmental and Engineering Geophysical Society
EERT Eastern Energy Resources Team
EM electromagnetic method
EMA Emergency Management Agency
EMC Emergency Management Coordinator
EMS Emergency Medical Service
EPA U.S. Environmental Protection Agency
EPRI Electric Power Research Institute
EROS Earth Resources Observation and Science
ESRI Environmental Systems Research Institute
FBC fuidized bed combustion
FE fnite element
FEMA Federal Emergency Management Agency
FERC Federal Energy Regulatory Commission
FGD fue gas desulfurization
FHWA Federal Highway Administration
FLAC Fast Lagrangian Analysis of Continua (software)
FLDWAV Flood Wave (software)
FLUSH seismic soil-structure interaction software
FR friction ratio
FS factor of safety
FVST feld vane shear test
GAI-LAP Geosynthetic Accreditation Institute - Laboratory Accreditation Program
GCL geosynthetic clay liner
GEI Geotechnical Engineers, Inc.
GIS geographic information system
GPR ground penetrating radar
GPS global positioning system
GRM generalized reciprocal method
HDPE high-density polyethylene
HEC USACE Hydrologic Engineering Center
HEC-1 open channel fow analysis software
HEC-GeoHMS GIS -based version of HEC-HMS software
< PREVIOUS VIEW
xliv
Acronyms
MAY 2009
ACRONYM REPRESENTS
HEC-HMS Hydraulic Engineering Center Hydrologic Modeling System (software)
HEC-RAS Hydraulic Engineering Center River Analysis System (software)
HMR Hydrometeorological Report
HSG hydrologic soil group
ICODS Interagency Committee on Dam Safety
IDF infow design food
LAMODEL software for computing stresses and displacements in mines
LI liquidity index
LIDAR light detection and ranging
LL liquid limit
LLNL Lawrence Livermore National Laboratory
LVDT linear variable diferential transformer
MAE Mid-America Earthquake
MARV minimum average roll value
MCE maximum credible earthquake
MDE maximum design earthquake
MIBC methylisobutyl carbinol
MMI Modifed Mercali Intensity
MPBX multiple-point borehole extensometer
MSF magnitude scaling factor
MSHA Mine Safety and Health Administration
NAS National Academy of Sciences
NASA National Aeronautics and Space Administration
NCB National Coal Board (Britain)
NEH National Engineering Handbook
NEHRP National Earthquake Hazard Reduction Program
NIOSH National Institute for Occupational Safety and Health
NMO normal moveout
NMSZ New Madrid Seismic Zone
NOAA National Oceanic and Atmospheric Administration
NR not reported
NRC National Research Council
NRCS Natural Resources Conservation Service
NSF National Science Foundation
< PREVIOUS VIEW
xlv
Engineering and Design Manual Coal Refuse Disposal Facilities
MAY 2009
ACRONYM REPRESENTS
NSSGA National Stone, Sand and Gravel Association
NWS National Weather Service
OBE operating basis earthquake
OLC open limestone channel
OSHA Occupational Safety and Health Administration
OSM Ofce of Surface Mining
PCCP prestressed concrete cylinder pipe
PCPT piezocone penetrometer test
PE polyethylene
PERT scheduling software
PGA peak ground acceleration
PI plastic index
PL plastic limit
PMF Probable Maximum Flood
PMP Probable Maximum Precipitation
PMTS Probable Maximum Thunderstorm
POA percent open area
PSHA probabilistic seismic hazard analysis
PV prefabricated vertical
PVC polyvinyl chloride
Q equivalent to UU
QA quality assurance
QC quality control
QUAD4 software for fnite element seismic analysis of earth structures
R equivalent to CU
RCCP reinforced concrete cylinder pipe
RCP reinforced concrete pipe
RMR rock mass rating
RQD rock quality designation
RVSP reverse vertical seismic profle
S equivalent to CD
SAGEEP Symposium on the Application of Geophysics to Engineering and Environmental
Problems
< PREVIOUS VIEW
xlvi
Acronyms
MAY 2009
ACRONYM REPRESENTS
SAPS successive alkalinity producing system
SCPTu seismic piezocone penetrometer test
SCS Soil Conservation Service (now NRCS)
SDI slake durability index
SDPS Surface Deformation Prediction System (software)
SEE safety evaluation earthquake
SF safety factor (for channel linings)
SHAKE software for seismic analysis of subsurface layers
SHANSEP stress history and normalized soil engineering parameters
SL shrinkage limit
SMCRA Surface Mining and Reclamation Act
SMPDBK Simplifed Dam Break (software)
SP spontaneous or self potential
SPBX single-point borehole extensometer
SPECTEXT specifcation software
SPT standard penetration test
SSE safe shutdown earthquake
SSHAC Senior Seismic Hazard Analysis Committee
TDEM time-domain electromagnetic method
TDR time-domain refectometry
TR technical release
USACE U.S. Army Corps of Engineers
USBM U.S. Bureau of Mines
USBR U.S. Bureau of Reclamation
USCS Unifed Soil Classifcation System
USDA U.S. Department of Agriculture
USDOE U.S. Department of Energy
USEPA U.S. Environmental Protection Agency
USGS U.S. Geological Survey
USNRC U.S. Nuclear Regulatory Commission
UU unconsolidated undrained
VLDPE very low density polyethylene
VLF very low frequency
< PREVIOUS VIEW
xlvii
Engineering and Design Manual Coal Refuse Disposal Facilities
MAY 2009
ACRONYM REPRESENTS
VST vane shear test
VSP vertical seismic profle
WMS Watershed Modeling System (software)
WP wetted perimeter
WVDOT West Virginia Department of Transportation
< PREVIOUS VIEW
xlviii
Acronyms
MAY 2009 < PREVIOUS VIEW
Chapter 1
INTRODUCTION
11 MAY 2009
This Engineering and Design Manual for Coal Refuse Disposal Facilities (Manual), originally pub-
lished by the Mining Enforcement and Safety Administration (MESA, 1975), has been updated for
the Mine Safety and Health Administration (MSHA) of the United States Department of Labor to
present current guidelines and procedures for design, construction and operation of coal refuse dis-
posal facilities. Guidance related to dam safety issues for other impounding structures at coal mine
sites, including fresh water impoundments and sedimentation and treatment ponds, is also provided
in individual sections or through discussion of the associated design features. Emphasis has been
placed on facility planning and materials handling and placement techniques. The input provided
by MSHA, other regulatory agencies, the coal mining industry, consulting engineers and the general
public has been benefcial to the preparation of this document. The guidance presented in this Man-
ual is advisory and not intended to discourage new and innovative methods that may be applicable
in specifc situations.
The focus of this Manual is on design, construction and operational practices for achieving stable em-
bankment and impoundment conditions based upon performance criteria, established regulations,
and accepted engineering standards. Typical performance criteria are associated with long-term
stability, food and seismic conditions, and abandonment. Where the application to disposal prac-
tices requires interpretation that may not have been envisioned by the developers of design criteria,
or where such criteria may not be available (as with some elements of seismic stability analyses),
guidance is provided along with commentary describing the basis for development of the guidance.
Criteria for environmental control of impacts from coal refuse disposal facilities vary from state to
state. The Manual discusses features currently incorporated into disposal facilities for environmental
control and provides guidance for their design and construction, as related to dam and embankment
safety concerns of MSHA. Guidance for evaluating potential environmental impacts or for establish-
ing environmental design criteria is lef to other references and regulatory agencies. For example, the
Manual provides information on the design and construction of liner systems for disposal facilities,
but guidance relative to establishing the need for or the required hydraulic conductivity of a liner is
available in other references.
Before the Bufalo Creek coal refuse facility failure in February 1972, coal refuse disposal was ofen
subject to only cursory engineering, planning and design. Through the atention of the industry, en-
suing regulatory programs, and guidance provided in the original 1975 Manual, improvements were
made in the design and construction of disposal facilities. Since 1975, mining and coal processing
< PREVIOUS VIEW
12
Chapter 1
MAY 2009
have undergone changes that afect the design of disposal facilities; advances in engineering practice
have led to modifcation of design criteria; and other events such as the breakthrough of impound-
ment basins into abandoned underground mine workings have disclosed potential considerations
that must be addressed. Thus, the original Manual has been updated to refect these changes. While
the structure and sequence of the previous Manual have been maintained, additional chapters have
been added to address design components such as seismic stability, site mining and foundation is-
sues, instrumentation, and emergency action planning.
This Manual is intended as a reference document for use by coal industry personnel and engineering
staf, design consultants, and dam safety regulators involved with the planning, design, construction
and regulation of refuse disposal facilities and other mine-related impoundments. It may be particu-
larly valuable to industry personnel who perform design coordination, as a guidance document for
exploration, testing, engineering analysis, design and construction document preparation, and con-
struction monitoring and inspection. For design engineers, the Manual provides background infor-
mation on coal refuse, along with methods and procedures for detailed design of coal refuse disposal
facilities. Construction personnel will fnd descriptions of critical engineering issues that should be
addressed during initial development and throughout the operational life of disposal facilities.
Chapter 2 presents an introduction to coal refuse including its origin, characteristics of the various
materials making up coal refuse, and disposal practices. Current challenges in the disposal of coal
refuse are also discussed.
Chapter 3 describes various types of coal refuse disposal facilities (both impounding and non-im-
pounding) and other site impoundments and the operations and hazard potential associated with
each. The classifcation and terminology used for embankments and impoundments are presented,
consistent with conventions adopted in the 1975 Manual and MSHA publications. Additionally, un-
derground injection of coal refuse and recovery (remining) of coal from refuse disposal facilities are
also discussed.
Chapter 4 presents updated planning and technical considerations associated with disposal site selec-
tion and facility design and emphasizes the importance of close coordination between the designer
and the coal operators production staf. Past experiences with improving and expanding existing fa-
cilities and the planning of new facilities demonstrate the importance of integrating planning, design,
construction and operations. The continually changing structure of refuse disposal facilities and the
need to eventually reach an acceptable abandonment condition are important factors in the planning
process.
Chapter 5 presents a new discussion of the design components of coal refuse disposal facilities, with
emphasis on the objective of meeting the generation rates of coal preparation plants. The chapter pro-
vides the design coordinator and the engineer with an overview of the objectives of disposal facility
design. The interrelationship between accommodating coal refuse generation rates, storm water and
environmental control requirements, stable embankment development, and post-mining land use
objectives are discussed.
Chapter 6 presents updated and new procedures for geotechnical exploration, engineering analysis,
and design of coal refuse disposal facilities. Initially, the discussion addresses the unique charac-
teristics of refuse disposal that afect facility design, as compared to the design of more commonly
encountered geotechnical structures. Design concepts for embankment zoning, internal drainage
control, and foundations are discussed with reference to related site exploration, testing, and geo-
technical engineering analyses. The site exploration section provides new and updated guidance
on cone penetration testing and geophysical methods applied to siting and design of impounding
and non-impounding refuse embankments. The laboratory testing section presents descriptions and
< PREVIOUS VIEW
13
Introduction
MAY 2009
references to standards for commonly used tests for refuse disposal facility design, along with sum-
maries of coal refuse material properties. Geotechnical analysis topics discussed include seepage,
setlement, stability, rock excavation, and conduit design. Chapter 6 is intended to provide an over-
view of geotechnical engineering requirements for disposal facilities for design coordinators who
are familiar with coal mining operations, but may not routinely practice civil engineering as related
to embankments and dams. It should also be a useful source to the practicing engineer for guidance
on approaches, methods, and design elements of geotechnical engineering applied to coal refuse
disposal facilities.
Chapter 7 presents new guidance for analysis of seismic stability and deformations applicable to
dams and coal refuse embankments. Many slurry impoundments are constructed using the upstream
construction method. Such sites are susceptible to instability during an earthquake because a por-
tion of the dam is founded on relatively loose and saturated fne material. Major advances in the
analysis of seismic stability of embankments and impoundments have occurred since the original
1975 Manual was published. Chapter 7 presents recommended design methods and fow charts to
assist in seismic stability analyses. Also discussed is the importance of variation in material gradation
(clay-like versus sand-like) that afects the selected approach and methods employed. Seismic hazard
assessment is discussed, and guidance is provided for determining ground motion parameters con-
sidering the potential source zones and conditions found in coal regions. Deformational analysis is
discussed along with criteria for tolerable deformations. Chapter 7 provides a condensed discussion
of a complex subject that should allow design coordinators to gain an appreciation for the engineer-
ing required for addressing seismic stability. It will also help practicing engineers with this portion of
the design by providing simplifying approaches in low-seismic-hazard regions. In regions of greater
seismic hazard and for signifcant- and high-hazard-potential structures, more complex approaches
for use by geotechnical engineers with experience in earthquake engineering are presented.
Chapter 8 has been added to the Manual to address site mining conditions, embankment foundation
conditions, and impoundment stability at coal refuse facilities and the potential for subsidence and
breakthrough into underground mine workings. Mitigation measures are also discussed. Addition-
ally, unique foundation issues that may be encountered at mine sites are addressed, including evalu-
ation of the efect of mine spoil materials and surface mine benches and highwalls on embankment
and foundation design.
Chapter 9 summarizes hydrologic and hydraulic engineering methods applicable to the design of
coal refuse disposal facilities. The Manual has been updated to refect current design storm criteria.
Examples of spillways, decants and channels with methods for performing design analyses and rout-
ing storms through disposal facilities are presented.
Chapter 10 addresses environmental considerations for coal refuse disposal facility design, including
streams and wetlands, air quality, water quality, and reclamation. Updated guidance is provided as to
methods for mitigating water quality concerns, including amendments and liner design. References
are provided for evaluation of water quality impacts and treatment methods. Reclamation issues are
also discussed. Design coordinators and engineers will fnd these sections useful when dealing with
environmental issues and mitigation methods, but they will also need to be familiar with applicable
state regulations and guidance.
Chapter 11 addresses disposal facility construction and operation, and Chapter 12 discusses con-
struction monitoring, inspection and maintenance. These chapters refect updates in construction
and disposal practices since the 1975 Manual was prepared and discuss some of the issues associ-
ated with large production mines such as material handling, placement, and compaction require-
ments. Guidance has been added relative to the content of construction documents and quality
control programs.
< PREVIOUS VIEW
14
Chapter 1
MAY 2009
Chapter 13 provides a discussion of instrumentation for monitoring the performance of coal refuse
disposal embankments and impoundments.
Chapter 14 provides a discussion of emergency action planning. It is recommended practice in dam
safety that emergency action plans (EAPs) be developed and maintained for all dams that will have a
signifcant downstream impact in the event of their failure.
< PREVIOUS VIEW
Chapter 2
BACKGROUND AND
CHARACTERIZATION
OF COAL REFUSE
21 MAY 2009
Safe, economical and environmentally acceptable disposal of coal refuse has been made possible
through the application of technology from the disciplines of geology, soil and rock mechanics,
hydrology, hydraulics, geochemistry, soil science, agronomy and environmental sciences. These
engineering and scientifc disciplines have traditionally supported the design and construction
of earthfll and rockfll embankments and dams, as well as a wide variety of waste disposal struc-
tures used in other industries. Among the various types of coal refuse disposal facilities, im-
poundments are one of the most critical types of structures, and they entail many of the same
features and safety concerns as typical water-impounding dams. However, there are distinct dif-
ferences between coal refuse disposal impoundments and conventional dams that must be con-
sidered in the transfer of dam engineering technology to coal refuse disposal, as discussed in the
following:
The primary purpose of a coal refuse disposal facility is to dispose of unusable ma-
terials from mining, not to construct an embankment to satisfy a secondary function.
However, at some sites coal refuse impoundments do serve secondary purposes such
as providing water storage capacity for coal processing and food atenuation.
Coal refuse is composed of rock fragments such as friable shale materials, and it
typically includes varying amounts of coal that can have an efect on material be-
havior. For example, the low specifc gravity of coal results in the refuse having
lower specifc gravities and densities than the materials encountered in typical earth
embankments.
Metals and sulfur present in some coal refuse materials may result in environmen-
tally undesirable leachate water with serious corrosive characteristics that require
special measures such as liner and collection systems or even amendments to neu-
tralize acid production of the refuse.
The potential for variations in coal refuse production as a result of mining or pro-
cessing, plus the long-term phased method of refuse placement, should be accounted
for in the disposal facility design at inception. The potential for changes in genera-
tion rates and engineering properties of the coal refuse during the life of the facility
should also be recognized and provisions to accommodate these changes should be
included in the design.
< PREVIOUS VIEW
22
Chapter 2
MAY 2009
The confguration of a disposal facility changes continually throughout its opera-
tional period, resulting in complications relating to control, but providing increased
fexibility for planning, design and construction.
Coal refuse disposal occurs year round, through prolonged inclement weather peri-
ods, such that the design and maintenance of the facility has to accommodate con-
struction under adverse weather conditions.
The construction of a disposal facility over an operational period of many years or
even decades introduces the potential for discontinuity in construction oversight,
quality control, monitoring, and recognition of performance factors that can afect
operation and safety. The design plans should include specifcations for oversight,
monitoring, and quality control sufcient so that facility construction and operation
meets the design intent.
A primary planning and construction goal for a disposal facility is achievement of
a safe and environmentally acceptable condition during use and upon closure and
abandonment. Elimination of the impounding capability of a coal refuse disposal fa-
cility is an important design issue associated with abandonment unless maintenance
of the facility is part of the planned post-mining operation of the site.
2.1 GEOLOGICAL NATURE OF COAL REFUSE
Coal refuse is composed of rock fragments unavoidably removed from the earth during the coal min-
ing process and small amounts of coal not separated during processing. The primary source of these
rocks and minerals is normally the formations immediately above and below the coal seam and the
sediments within the seam. Surface mine overburden and rock removed to provide shafs, haulage-
ways and other working space for the mining operation may constitute an important percentage of
the total refuse material.
The proportion of rock, coal and associated minerals in any disposal facility will depend upon: (1)
the geologic formation of the coal stratum, (2) the geochemical properties of the coal and adjacent
materials, (3) the geometry of the coal seam, (4) the method used for mining, and (5) the process used
to separate the coal from the refuse. A preparation plant may process coal from multiple seams and
mining operations, resulting in some variability of the refuse product. Also, a disposal facility may
receive refuse from multiple preparation plants.
2.1.1 Origin of Coal
Coal is a sedimentary rock resulting from past accumulation of plant materials in swampy conditions
and subsequent metamorphism by pressure and heat to form a hard, rock-like organic material. The
diversity of the original plant materials and the degree of metamorphism (coalifcation) that has af-
fected these materials cause coal to be a non-uniform, combustible substance varying in both physical
and chemical composition.
The predominant elements in coal are carbon, hydrogen and oxygen. Coals are rich in carbon and are
classifed based primarily as to the content of volatile mater. Table 2.1 presents the American Society
for Testing and Materials (ASTM) system for classifcation of coals by rank. For the anthracite and
low- and medium-volatile bituminous coals, the carbon content exceeds about 70 percent. When the
volatile mater exceeds about 30 percent, it becomes difcult to classify coals on the basis of volatile
mater alone, and caloric value is used to distinguish high-volatile-bituminous, subbituminous, and
lignite coals.
Impurities present in coal include nitrogen, sulfur, iron and various other inorganic materials (ash).
These impurities can be divided into the following classes: (1) impurities that are chemically or struc-
< PREVIOUS VIEW
23
Background and Characterization of Coal Refuse
MAY 2009
turally a part of the coal and (2) impurities that can be mechanically removed from the coal. The
amount and type of impurities that cannot be economically separated from a particular coal vary
greatly and afect value. The impurities that are separable afect the refuse characteristics from a geo-
chemical and physical standpoint. These characteristics contribute to the engineering properties of
coal refuse and are thus critical to the engineering and design of disposal facilities.
TABLE 2.1 CLASSIFICATION OF COALS BY RANK
Group
Fixed Carbon
Limits
(1)
(%)
Volatile Matter
Limits
(1)
(%)
Calorifc Value
Limits
(2)
(BTU/lb)
Anthracite
Meta-anthracite 98 < 2
Anthracite 92 to 98 2 to 8
Semianthracite 86 to 92 8 to 14
Bituminous
Low-volatility bituminous coal 78 to 86 14 to 22
Medium-volatility bituminous coal 69 to 78 22 to 31
High-volatility-A bituminous coal < 69 31 14,000
High-volatility-B bituminous coal 13,000 to 14,000
High-volatility-C bituminous coal 10,500 to 13,000
Subbituminous
Subbituminous-A coal 10,500 to 11,500
Subbituminous-B coal 9,500 to 10,500
Subbituminous-C coal 8,300 to 9,500
Lignite
Lignite A 6,300 to 8,300
Lignite B < 6,300
Note: 1. Dry mineral-matter-free basis.
2. Moist mineral-matter-free basis.
(ADAPTED FROM ASTM, 2008a)
Sulfur, an impurity of major importance, occurs in three principal forms: (1) organic sulfur, (2) sul-
fate sulfur, and (3) pyritic sulfur. Generally, pyritic sulfur (FeS
2
) is the predominant form. Pyrites are
of signifcance in the design of coal refuse disposal facilities because, when exposed to air and wa-
ter, they oxidize to create acidic conditions that adversely afect the weathering resistance of certain
rocks, cause negative environmental impacts, and cause corrosion of some construction materials.
2.1.2 Coal-Related Rocks
The rock formations immediately above and below a coal seam represent a large portion of the mate-
rial in coal refuse. The character of this rock varies from seam to seam and with the geographic loca-
tion of an individual coal seam, depending upon the geologic conditions preceding and following the
deposition of the coal-forming materials.
Inorganic sedimentary rocks, such as claystone, siltstone, shale, sandstone and occasionally lime-
stone, compose the bulk of the strata where coal is found. These sediments are frequently found in
< PREVIOUS VIEW
24
Chapter 2
MAY 2009
the coal as inorganic debris deposited during the formation of the organic layer. The most abundant
minerals contained in these inorganic rocks are quartz, feldspar, mica (usually muscovite with minor
chlorite) and calcite. However, numerous minor components may also be present, and these can vary
greatly from site to site. The following sedimentary rocks are listed in an order of decreasing resis-
tance to weathering:
Limestone A relatively hard, dense rock consisting largely of calcite (CaCO
3
). This rock
is very durable in the weathering cycle, except in the presence of acid (in-
cluding acid conditions that result from pyritic sulfur), which will dissolve it.
Sandstone A rock composed of relatively coarse particles cemented together by calcite,
silica (SiO
2
) argillaceous material or, less commonly, iron. Its strength and
resistance to weathering depend upon the cementing agent.
Siltstone A rock quite similar to sandstone, except for the generally smaller grain size
and substantial amount of clay present. The clay portion is subject to rapid
weathering, and high clay content can cause rapid deterioration.
Claystone A rock composed predominantly of clay-sized particles. It may be massive
(mudstone) or fssile (shale) in appearance, but in either case, it has a low
resistance to weathering
Iron and sulfur impurities are also common in many coal-bearing sedimentary rocks. As with the
coal itself, the most signifcant of these impurities are the pyrites.
2.2 COAL MINING AND COAL PREPARATION (CLEANING)
Coal is mined in more than 400 coal felds and small deposits in 38 states, split between the Eastern
(Appalachian and Interior) and Western regions. In the Eastern region, the quality of raw coal has
declined as higher quality reserves have been depleted, requiring more cleaning to meet product re-
quirements. The thicker Western region coal seams typically contain fewer in-seam and out-of-seam
rock, and much of that coal is shipped raw without extensive processing (NRC, 2002). Accordingly,
coal cleaning and the generation of coal refuse is a signifcant part of mining projects in the Eastern
region, which is where most coal refuse impoundments are operated.
In general, the removal of extraneous materials with the coal during the mining process is controlled
by three primary factors: (1) the general geology of the mining area and to a greater extent the im-
mediate type and quality of the foor, partings and roof materials, (2) the thickness of the seam to be
mined, and (3) the type of mining equipment selected to remove the coal. For example, in longwall
mining where a shearer or cuting head is drawn across the face of a coal seam, a constant thickness
of material is removed and will contain roof or foor rock in areas of decreased coal seam thickness.
This extraneous material, ofen referred to as out of seam dilution, is a major source of coal refuse,
particularly in the coarser (> 19-mm) size fractions.
Coal preparation (cleaning) removes refuse from the coal. The preparation technique afects the refuse
disposal primarily in three ways: (1) the proportion of coal in the refuse (efciency of the cleaning
process), (2) the gradation of the refuse particles, and (3) the moisture content of the refuse as it leaves
the preparation plant. Additionally, the use of focculants and other chemical additives can also afect
the composition of refuse materials. Most modern coal preparation plants today recover coal down to
100-mesh (150-micron) size, and some have incorporated cleaning all of the raw coal. Usually a prepa-
ration plant will have either two or three circuits (possibly four if the preparation plant cleans all the
way to zero impurities) that handle the distinct size fractions discussed in the following paragraphs.
Ofen raw coal is processed through a scalping screen or rotary breaker ahead of the preparation
plant. This is done to control the maximum size of the raw coal arriving at the preparation plant.
< PREVIOUS VIEW
25
Background and Characterization of Coal Refuse
MAY 2009
Historically, screening has been done at the 4- to 6-inch size range (100- to 150-mm). Recently, how-
ever, the trend has been to reduce the separation size down to as small as 1 inch (37 mm). This has
been done to accommodate the use of the simpler two-circuit coal preparation plant. Depending on
the mining method and coal characteristics, the rejected material that must be handled at the refuse
disposal site can represent 5 to 15 percent of the raw coal.
Typically, coal preparation plants size raw coal into discrete fractions that are processed separately
using various types of process technologies. Currently, heavy media technology is typically used for
processing raw coal down to 1 mm. Heavy media vessels (down to 9 mm) or heavy media cyclones
(down to 1 mm) or a combination of these two technologies are used to process the plus 1-mm raw
coal. The heavy media process involves adding magnetite to water to create slurry with a specifc
gravity that improves the separation efciency of the process. The magnetite slurry creates a sus-
pension with specifc gravity high enough that the coal foats and the refuse materials with higher
specifc gravity sink. These types of processes are the most efcient and produce a refuse stream with
minimal coal. Afer the separation of the coal from the refuse, the materials are drained and rinsed of
the magnetite media on a vibrating screen. Rinse water is added to the vibrating screen to improve
the efciency of the magnetite removal. Typically, the refuse material discharging from the end of
the vibrating screen afer being rinsed will have a surface moisture in the range of 8 to 15 percent
depending upon the minimum size of the refuse being processed.
The 1-mm by 100-mesh (150-micron) size fraction is typically processed using water-only technolo-
gies. Technologies used to process this size range include water-only cyclones, spirals, and hindered-
bed setlers. Afer processing, the refuse materials are typically dewatered using a high frequency
dewatering screen designed to retain the plus 100-mesh material. The installation of this type of
screen ahead of the thickener can signifcantly reduce the amount of plus 200-mesh particles in the
refuse slurry.
If coal cleaning below 100-mesh size is performed, then typically the diferences between the surface
chemistry of the coal and refuse are exploited to perform a separation (e.g., froth fotation). Froth
fotation requires the addition of a collector (typically fuel oil) and a frother (typically glycol or me-
thylisobutyl carbinol (MIBC)) to the slurry. The fne coal particles are hydrophobic and collect on
the surface of the air bubbles created by the fotation machine and rise up through the slurry and
are removed with the froth. Alternatives to fotation include high gravity concentrators and fne coal
spirals used to process the 100-mesh by 325-mesh (44-micron) size fraction.
Due to the high volume of water used in the processing of coal, thickeners are typically used to re-
claim the wash water for recirculation. Generally, the use of focculants is sufcient to remove the
majority of coal fnes, but when there are high levels of clay in the mined coal, it may be necessary
to also add coagulants. The sequential addition of focculants and coagulants aids in controlling the
turbidity of recycled process water and setling of solids that are pumped to the impoundment.
Technologies for dewatering of the thickener underfow include belt flter presses, vacuum flters,
plate and frame flters, solid bowl centrifuges, horizontal belt flters, and paste thickeners. When fnes
are dewatered, the most commonly used technology is the belt flter press. The dewatered product
from these technologies typically has moisture contents in the range of 35 to 45 percent. The amount
of minus 325-mesh solids in the feed to these unit operations dictates the fnal moisture of the prod-
uct. The higher the amount of minus 325-mesh particles in the feed (which is controlled by the geol-
ogy and mining method), the more difcult it is for the system to operate and to produce a product
with manageable moisture.
If fne refuse dewatering is included in the process, refuse streams must be handled either individu-
ally or as a combined product. Amendments may be required for pH control or to aid in the handling
< PREVIOUS VIEW
26
Chapter 2
MAY 2009
of the refuse products. Depending on the amount of fne refuse and its moisture content, the disposal
of the combined product may be difcult.
Prior to 1970, many preparation plants did not atempt to separate coal and impurities in the fne-
washed material, resulting in a fne refuse with high coal content. This practice resulted in a relatively
high ratio of fne refuse volume to coarse refuse volume, and these disposal sites are of interest for
remining and recovery of the coal. In general, modern coal preparation plants are more efcient (i.e.,
lower coal losses to refuse). In addition, the amount of out-of-seam dilution has increased, result-
ing in higher coarse to fne ratios. This is particularly true in areas where the coal reserve is being
depleted. As coal preparation techniques have advanced, the amount of fne coal deposited with the
fne refuse slurry has decreased, thus increasing the amount of coarse refuse disposed in relation to
the amount of fne refuse.
2.3 REFUSE TRANSPORT AND DISPOSAL PLACEMENT
2.3.1 Coarse Refuse
The method of transporting coarse coal refuse from the preparation plant to the disposal facility is de-
termined by: (1) material gradation, (2) production rate, (3) the distance and topography of the route
to the disposal facility and (4) the size, type and confguration of the disposal facility. Although refuse
transport methods vary widely, the most common method for transporting coarse refuse is by trucks
or scrapers for relatively short distances over moderate terrain, by conveyors for long distances over
moderate to steep terrain, or by a combination of these methods. The ramifcations of these transport
methods on the disposal facilities are further described in Chapter 11.
The placement of coarse refuse at a disposal facility depends upon the method of transport to the
disposal facility and the facility confguration. Prior to 1970, coarse refuse transported by truck was
generally end-dumped from the truck and allowed to form a progressing embankment at the angle
of repose of the material. Another method from that era was to transfer the coarse refuse from a
conveyor or continuous bucket tram to an aerial tram that dumped the refuse from a high elevation
along a fxed axis, allowing the refuse to form a progressing embankment at the angle of repose of the
material except as modifed by dozers or scrapers.
Current practice typically includes the use of conveyor systems and/or trucks. Trucks dump the re-
fuse in piles on the disposal surface with subsequent spreading by dozers. If the refuse is transported
by conveyor, it is normally dumped into a hopper bin and transferred to trucks or scrapers for place-
ment, as described above. There has been some use of mobile conveyors to deposit coarse refuse di-
rectly on the embankment surface for spreading by dozers. The common confgurations of coal refuse
embankments are described in Chapter 3.
2.3.2 Fine Refuse and Combined Refuse
A common method of transporting fine refuse material has been as slurry pumped to a disposal
pond for settling of the solids and clarification of the remaining water. Use of thickeners with
the application of flocculants, and in some cases coagulants, improves clarification and settling
of solids and facilitates deposition within a slurry impoundment. These additives typically over-
come the use of reagents used in the separation and concentration processes of coal preparation.
Another method for transporting fne refuse involves partially dewatering the slurry at the prepara-
tion plant. The resulting material is then disposed separately or mixed and placed with the coarse
refuse material as combined refuse. There have been many problems associated with transporting
and placing the combined material due to the relatively high moisture content of the dewatered fne
refuse. Consideration has recently been given to transporting the fne refuse as a paste (thickened
< PREVIOUS VIEW
27
Background and Characterization of Coal Refuse
MAY 2009
tailings) that can be pumped to a disposal location. The objective in using thickened tailings is to
minimize the use of water, generally because of limited availability.
2.4 COAL REFUSE CHARACTERIZATION
Coal refuse is composed of rock fragments, sand, silt and clay particles and contains small amounts
of coal not separated during processing.
2.4.1 General Geochemical Characteristics
Coal refuse geochemical characteristics result from the depositional environment of the coal and
adjacent strata, the mining and coal preparation process, and the weathering process when the re-
fuse is exposed to air. Coal refuse shares many properties with the coal seam where it originated,
with sulfur usually representing the most signifcant environmental concern. Mining processes may
take mine roof and foor materials that contain signifcant pyrite materials, further afecting sulfur
content. Processing of the coal includes crushing, separation, and in some plants recombining coarse
and fne refuse. The efciency of the processing plant at removing sulfur and the degree to which the
sulfde fragments are fractured and reduced in size infuences the reactivity of the fnal refuse prod-
uct. Reagents and additives such as surfactants, oils, and strong bases are used in various separation
processes.
Most of the environmental problems associated with coal refuse result from the oxidation of pyrite and
subsequent production of acidity. The exposure of coal refuse to weather (air and moisture) enhances
the production of acid, leading to a condition of low pH. The pH of a refuse deposit may depend not
only on the pyrite content, but also on the length of exposure time and the acid-neutralizing capacity
of the refuse. The majority of the coal refuse in the Appalachian region of the U.S. contains an excess of
oxidizable sulfur compared to neutralizing carbonates and is therefore net acid producing over time.
During the oxidation process, metals are released into soluble forms, resulting in impacts to water qual-
ity from drainage through or of the surface of refuse deposits and accumulation of phytotoxic concen-
trations of acid-soluble metal ions and sulfate salts on refuse surfaces that inhibit the establishment of
vegetation during reclamation of the disposal site (Daniels and Stewart, 1993). Western coal regions in
the U.S. tend to be less acid producing, but may still pose environmental concerns.
Variability may be found in coal refuse within the same disposal area because: (1) the coal prepara-
tion plant may process coal from multiple seams, (2) the process may change over time, and (3) the
rate of processing (and thus accumulation and successive covering of earlier refuse deposits) may
change due to production and mining factors. The management of refuse at the disposal site can
afect the weathering process, with the result that geochemical characteristics can change over rela-
tively short periods of time.
The geochemical characteristics of coal refuse are important planning considerations for coal refuse
disposal facilities and can be addressed in a variety of ways. Amendments (additives with neutral-
ization characteristics such as lime products) may be added to provide additional acid-neutralizing
capacity that can afect the physical properties of the deposit. Liners (natural soil or synthetic) may
be employed for control of seepage from the refuse disposal area and thus will afect foundation re-
quirements and facility confguration.
2.4.2 General Geotechnical Characteristics
Each type of coal refuse has characteristic geotechnical properties. Coarse refuse from the preparation
process represents a substantial portion of and sometimes the entire waste stream. Where mechanical
crushers are included in the coal preparation process, large-diameter rock (on the order of 6 inches in
size and referred to as breaker rock) may also be generated, although it typically represents a small
fraction of the refuse materials.
< PREVIOUS VIEW
28
Chapter 2
MAY 2009
Fine refuse is a separate waste stream resulting from the wet processing of coal. It may be: (1) dis-
posed as a slurry (fne coal refuse slurry) separate from the coarse refuse, (2) dewatered and disposed
with the coarse refuse as combined refuse, or (3) dewatered and disposed separately from the coarse
refuse (dewatered fne coal refuse or flter cake). An additional form of refuse, although not common,
is the slurrying of both coarse and fne refuse (coarse and fne refuse slurry).
Most slurry disposal operations use slurry pumps to transport fne coal refuse at between 5 and 20
percent solids. Thickened tailings are increasingly being considered for fne coal refuse disposal. This
practice involves processing the fne coal refuse into a non-setling and non-segregating suspension
of solids taking the form of a paste with about 50 percent solids.
2.4.2.1 Coarse Coal Refuse
Coarse coal refuse is typically a well-graded material with particle sizes ranging up to 3 inches and a
fnes content (as measured by the amount passing the No. 200 sieve or 0.075 mm) ranging from less
than 10 percent to more than 20 percent. It is generally classifed as a silty, clayey sand with gravel to
a clayey, silty gravel with sand. Coarse coal refuse typically has a specifc gravity ranging from 1.8 to
2.3, which is lower than many natural soils and aggregates due to the carbon content. Run-of-plant
coarse coal refuse generally has a water content of between 8 and 15 percent and loses moisture dur-
ing transport to the disposal site. As a result, it can readily be compacted to form dense fll in embank-
ments. Figure 2.1 provides the range of grain-size distributions usually encountered for coarse coal
refuse at disposal sites and a photograph of a coarse refuse embankment.
Coarse coal refuse typically exhibits weathering and some physical degradation during placement
and compaction. Thus, the resulting percentage of fnes in compacted, weathered coarse coal refuse
can be greater than in fresh or recently placed material. Sieve analyses performed on fresh samples
before and afer compaction have exhibited an average increase in fnes content of approximately 4
percent in one northern Appalachian mine. Coarse coal refuse generally has strength, setlement and
hydraulic conductivity properties refective of a well-graded soil and rock fll. Further characteriza-
tion of coarse refuse is discussed in Chapter 6.
2.4.2.2 Fine Coal Refuse Slurry
Fine coal refuse slurry, when discharged into an impoundment, meanders across previous deposits
and the sand-size material commences setling in shallow water. As the slurry discharge velocity
FIGURE 2.1 COARSE COAL REFUSE EMBANKMENT AND SAMPLE GRADATION
FIGURE 2.1 COARSE COAL REFUSE EMBANKMENT AND SAMPLE GRADATION
100 10 1 0.1 0.01 0.001
(HEGAZY ET AL., 2004)
10
20
30
40
50
60
70
80
90
100
0
SIEVE SIZE (mm)
P
E
R
C
E
N
T
P
A
S
S
I
N
G
SAMPLE GRAIN-SIZE
RANGE FOR COARSE
COAL REFUSE
AVERAGE
FIGURE 2.1 COARSE COAL REFUSE EMBANKMENT AND SAMPLE GRADATION
< PREVIOUS VIEW
29
Background and Characterization of Coal Refuse
MAY 2009
slows and deeper water conditions are encountered, setling of the fner materials occurs, resulting in
greater variability in grain size than encountered with other forms of coal refuse placement. Samples
collected from or close to the slurry discharge point are predominantly sand- and silt-size material,
whereas samples collected away from the discharge point are predominantly silt- and clay-size ma-
terial. Because the discharge location is typically shifed periodically, there is variability in the grain
size of fne refuse materials with elevation as well as location. Figure 2.2 illustrates typical grain-size
distributions encountered with discharge of fne coal refuse slurry at impoundments.
Fine coal refuse consists of particles that are mostly less than 1 millimeter in size and frequently has
a fnes content (as measured by the amount passing the No. 200 sieve) of typically between 30 and 80
percent. The specifc gravity of fne coal refuse tends to be lower than that of coarse refuse, typically
in the range of 1.4 to 2.0, depending on the percentage of coal fnes. Water content will vary signif-
cantly dependent upon the grain-size distribution and location within the impoundment. Setled fne
coal refuse slurry has strength, setlement and hydraulic conductivity properties refective of hydrau-
lically-placed deposits of sand, silt and clay, as further discussed in Chapter 6.
2.4.2.3 Dewatered Fine Coal Refuse
Dewatering of fne coal refuse results in a flter cake material with less variability in grain size than
observed in hydraulically-placed fne refuse slurry. While the grain-size distribution is relatively con-
sistent for a given coal seam and processing plant, there is typically variation between samples from
diferent coals and processing systems. Filter cake may have a retained water content in excess of 30
percent, and considering the signifcant amount of silt and clay fraction in the material, frequently
cannot be placed and compacted into an embankment without further drying and/or the addition
of other materials such as coarse refuse or amendments for moisture control and stabilization. Con-
sequently, dewatered fne coal refuse is generally disposed with coarse refuse either intermixed as
combined refuse or segregated to form containment cells.
2.4.2.4 Combined Coal Refuse
Combined coal refuse is a mixture of coarse and fne coal refuse, and thus has a greater fnes content
and water content than coarse coal refuse. While still well graded, combined refuse may have fnes
FIGURE 2.2 FINE COAL REFUSE SLURRY CHARACTERISTICS
FIGURE 2.2 FINE COAL REFUSE SLURRY CHARACTERISTICS
SIEVE SIZE (mm)
P
E
R
C
E
N
T
P
A
S
S
I
N
G
(HEGAZY ET AL., 2004)
100 10 1 0.1 0.01 0.001
10
20
30
40
50
60
70
80
90
100
0
SAMPLE GRAIN-SIZE
RANGE FOR FINE
COAL SLURRY
AVERAGE
FIGURE 2.2 FINE COAL REFUSE SLURRY CHARACTERISTICS
< PREVIOUS VIEW
210
Chapter 2
MAY 2009
content in the range of 20 to 40 percent, and water content between 12 and 30 percent. When the fnes
and/or water content are relatively high, combined refuse may be difcult to compact into a dense
embankment and may also be sensitive to inclement weather impacts. Figure 2.3 presents a photo-
graph of combined coal refuse and an illustration of a typical grain-size distribution for combined
refuse with a moisture content of 30 percent.
2.4.2.5 Fine Coal Refuse Paste
To reduce the quantity of water associated with fne refuse disposal, thickened tailings technol-
ogy is currently under study. It is anticipated that the resulting paste will have properties similar
to dewatered fne coal refuse from flter presses, but with greater water content. The atraction to
this technology lies in certain perceived advantages over flter press operating requirements and
the ability to transport the resulting paste by pumping, using less water than required for slurry
impoundments. Placement and containment issues are anticipated to be similar to those encoun-
tered with disposal of dewatered flter cake. The use of thickened tailings has been implemented
at two existing impoundments in the United States (Henry, 2007; Gupta et al., 2008), achieving a
paste solids content of 45 to 55 percent that reportedly builds or stacks without separation as the
paste is deposited.
2.5 DISPOSAL PRACTICES
2.5.1 Disposal Practices Prior to the Bufalo Creek Failure
In February 1972, a coal refuse disposal facility located on a tributary to Bufalo Creek in southern
West Virginia failed afer several days of rainfall, sending food water and refuse material through
the narrow downstream valley. A number of coal mining communities were devastated, and 125
people were killed. This monumental failure, known as the Bufalo Creek failure, brought the po-
tential danger of similar disposal facilities to the atention of the nation. A summary description
of this failure along with other incidents reported at coal refuse disposal facilities is provided in
Table 2.2.
Prior to the Bufalo Creek failure, minimal technical efort was devoted to the planning and design of
coal refuse disposal facilities in the U.S. Geotechnical and hydraulic characteristics of refuse materials
and disposal facilities were seldom considered. Instead, disposal facilities were ofen developed at
FIGURE 2.3 COMBINED COAL REFUSE MATERIALS
FIGURE 2.3 COMBINED COAL REFUSE MATERIALS
100 10 1 0.1 0.01 0.001
10
20
30
40
50
60
70
80
90
100
0
SIEVE SIZE (mm)
%
P
A
S
S
I
N
G
SAMPLE GRAIN-SIZE
RANGE FOR COMBINED
COAL REFUSE
NOTE: DATA WERE OBTAINED FROM UNPUBLISHED STUDIES.
FIGURE 2.3 COMBINED COAL REFUSE MATERIALS
< PREVIOUS VIEW
211
Background and Characterization of Coal Refuse
MAY 2009
sites chosen for expediency and were ofen operated with equipment selected by availability rather
than suitability for site materials. Most research devoted to coal refuse disposal in the U.S. before
1972 was related to: (1) the prevention of refuse embankment burning, (2) the extinguishment of
embankment burning, and (3) the development of vegetation on refuse materials through refuse
conditioning and seeding.
Prior to 1972, little or no planning or design was devoted to the control of water entering an
impoundment from a preparation plant or from watershed runoff upstream of the disposal
facility. Experience had suggested that excess preparation plant water could be clarified natu-
rally by settlement at the impoundment and filtration through the refuse embankment. In most
coal refuse impoundments, the water level above the settled slurry reached an equilibrium
condition at a depth of less than 10 to 15 feet, at which time the outflow of water from seepage
equaled the inflow to the impoundment. This effect, plus the large volume of coarse refuse
often placed in the embankment, generally resulted in the maintenance of a large freeboard
between the normal water surface and the embankment crest that provided storage for runoff
from moderate storms.
The potential risk of coal refuse disposal facilities is related to the storage of water and waste that may
be fowable and that, if released, may result in downstream inundation with substantial public safety,
economic and environmental consequences. A major reason for emphasizing the potential hazard of
disposal facilities with impoundments, as compared to disposal facilities without impoundments,
is the large geographical area that may be afected by the rapid release of large volumes of water
or waste following embankment failure. Unless a non-impounding embankment is located directly
above a populated area such as the disastrous failure in Aberfan, Wales in 1966 (Bignell et al., 1977),
or adjacent to a stream where materials movement could block fow, the slippage of such an embank-
ment will cause only localized problems related more to environmental and aesthetic considerations
than to public safety. However, embankment failure allowing the rapid release of large volumes of
water may cause fooding that is costly not only in terms of miner and public safety, and property
and environmental damage, but also in terms of extensive cleanup of the dispersed wet refuse for
many miles downstream. In mountainous mining areas complementary industrial, commercial and
residential developments are usually located in the downstream valley botom.
2.5.2 Disposal Practices Subsequent to the Bufalo Creek Failure
Following the Bufalo Creek failure in February 1972, regulations were promulgated for the design
of embankment structures, and programs and resources were developed, including the 1975 Manual
to assist in design, inspection and construction of coal refuse disposal facilities. Hundreds of exist-
ing and planned refuse disposal facilities were upgraded to new and safer standards, and for slurry
impoundments, signifcant emphasis was placed on dam safety.
The various programs and resources have produced the following major, positive results:
It has been recognized that coal refuse disposal facilities can represent a threat to
public safety and the environment, and positive steps toward eliminating these
threats in future disposal operations have been taken.
The principles of geotechnical engineering, hydrology, hydraulics, and dam-safety
engineering have been applied to the design and construction of coal industry im-
poundments to enhance the safety of these structures.
There is industry-wide recognition of the importance of coal refuse disposal to
preparation plant operation, and disposal practices are being incorporated into the
feasibility, planning, and management of mine operations.
< PREVIOUS VIEW
212
Chapter 2
MAY 2009
Data that contribute to the application of geology, soil and rock mechanics, hydrol-
ogy, hydraulics, dam engineering and geochemistry technologies to the evaluation
and design of disposal facilities have been collected and analyzed.
Research has yielded important advances in the prediction of the performance of
disposal facilities, and areas have been identifed where new research is needed for
the design and operation of disposal facilities.
The potential economic beneft of integrating the disposal facility design into the
total plan for coal production operations has been demonstrated.
While efective engineering and design measures have been implemented to minimize the potential
for failure of coal refuse disposal facilities, incidents have occurred that may be instructive. A sum-
mary of these incidents is presented in Table 2.2.
Challenges remain for the coal industry, including the following:
Advances in mining technology, including the more widespread application of long-
wall mining systems, increases the potential for subsidence and associated impacts
to the overburden strata that must be addressed at disposal facility sites, either as
part of siting and new facility design, or to permit mining beneath or adjacent to ex-
isting disposal facilities. Additionally, the threat of subsidence from old mine work-
ings at some sites may require evaluation and remedial action.
Production capacity at mines has continued to increase as a result of larger longwall
mining systems and operations with continuous miners, and in some cases more
reject material (non-coal rock) is being mined, with a corresponding increase in the
generation of coal refuse. This has resulted in larger embankments and impound-
ments that must accommodate refuse placement at a greater rate.
Longwall systems and larger mines with high production rates may exhibit variabil-
ity in the refuse generation rate. Disposal facilities should be planned and designed
considering this potential variability.
Integration of surface mine operations with coal refuse disposal poses opportunities
for more efcient construction, provided design issues associated with mine spoil/re-
fuse embankments are adequately addressed.
Amendments and admixtures have become common, whether as part of operations
(transport of combustion ash from power plants to mine sites) or to address environ-
mental requirements (neutralization of acid generation from refuse). These material
additions may afect the performance of the refuse embankments and should be
considered during the design process.
Re-mining of existing coal refuse facilities to recover coal has developed into a niche
industry. Development plans must be prepared for the safe excavation and removal
of coal refuse from disposal facilities while maintaining the original geotechnical and
hydraulic design criteria and considering the potential for shutdowns experienced
with such operations.
Slurry impoundments should be designed considering the potential impact of
seismic events. While seismic failures of such facilities have not been reported in the
U.S., the potential exists at some refuse disposal facilities.
Multiple regulatory criteria from federal and state agencies with difering perspec-
tives must be accommodated in site selection and design.
< PREVIOUS VIEW
213
Background and Characterization of Coal Refuse
MAY 2009
TABLE 2.2 COAL REFUSE DISPOSAL FACILITY FAILURES AND INCIDENTS
February 26, 1972: Buffalo Mining Company,
Buffalo Creek, West Virginia
On February 26, 1972, the most destructive food in West Virginias history occurred when a coal waste impounding
structure collapsed on the Buffalo Creek tributary of Middle Fork. Shortly before 8:00 a.m., the impounding structure
collapsed, releasing approximately 132 million gallons of water. The water passed through two more piles of coal waste
blocking the Middle Fork. At that time, there were no federal standards requiring either impoundments or hazardous
refuse piles to be constructed and maintained in an approved manner.
Around 1957, as part of its surface mining operations, the Buffalo Mining Company (a subsidiary of the Pittston Coal
Company) had begun depositing mine waste consisting of rock and coal in Middle Fork. Buffalo Mining constructed
its frst impounding structure, near the mouth of Middle Fork in 1960. Six years later, it added a second impounding
structure, 600 feet upstream. By 1968, the company was depositing more waste another 600 feet upstream. By 1972,
the height of this third impounding structure ranged from 45 to 60 feet.
Between February 24 and 26, 1972, the National Weather Service measured 3.7 inches of rain in the area of
Logan County and Buffalo Creek. The impounding structure probably failed because foundation defciencies led
to sliding and slumping of the front face of the refuse bank. The waterlogged refuse bank accelerated the failure.
The slumping lowered the top of the refuse bank and allowed the impounded water to breach and then rapidly
erode the crest of the bank. Upon failure of the refuse bank, the foodwater moved into pockets of burning coal
waste.
As result of the food, 125 people were killed, 1,100 were injured, and more than 4,000 were left homeless. In addition,
the food completely demolished 1,000 cars and trucks, 502 houses, 44 mobile homes, and damaged 943 houses and
mobile homes to varying extents. Property damage was estimated at $50 million.
Source: Davies et al., 1972
August 14, 1977: Island Creek Coal Company,
Boone County, West Virginia
An embankment under construction failed at Island Creek Coal Companys impoundment in Boone County, West
Virginia, on August 14, 1977. Heavy rainfall overfowed a temporary diversion ditch, causing the water level in the
impoundment to rise. Because the embankment was still under construction, storage capacity had not yet reached the
required minimum, and the sudden infux of additional water overtopped the embankment. Meanwhile, the water eroded
the embankment, reducing its height 23 feet during a two-day period. During this time, 6.8 acre-feet were released,
clogging a drainage pipe downstream.
Source: Owens, 1977
December 18, 1981: Eastover Mining Company,
Harlan County, Kentucky
On December 18, 1981, Eastover Mining Companys Hollow No. 3 combined refuse disposal site failed, releasing about
25 million gallons of saturated coal refuse. The operation, which had been permitted for disposal of coarse coal refuse
and dewatered slurry flter cake that contained approximately 30 percent moisture behind an embankment 192 feet
high, had reached 90 percent of its planned capacity. Several factors contributed to the increased pore water pressure
in the dewatered fne refuse zone, including: (1) the flter cake layers had not been allowed suffcient time to dry before
additional material was added; (2) layers of flter cake were not completely covered with coarse coal refuse; (3) a stream
fowed into the impoundment material, increasing saturation; and (4) material used in construction of the embankment
did not allow water to seep out. The failure released a mudfow approximately 5 feet deep that traveled 4,400 feet down-
stream (500 feet in vertical distance) into the community of Ages, Kentucky. One resident was killed, three houses were
destroyed and 30 homes were damaged.
Source: Cannon, 1981
April 8, 1987: Peabody Coal Company,
Raleigh County, West Virginia
On April 8, 1987, a breach developed in the principal spillway pipe in the Lower Big Branch impoundment at Peabodys
Montcoal No. 7 complex in Raleigh County, West Virginia. The 36-inch-diameter pipe ran through the impoundment and
under part of the embankment at a depth of 55 feet. The rupture released nearly 23 million gallons of water, slurry, and
fne coal refuse.
< PREVIOUS VIEW
214
Chapter 2
MAY 2009
The exact cause of the accident was not identifed but was probably the result of a combination of factors: (1) Heavy
snowfall (16 inches of snow with a rainfall equivalent to 1.9 inches), followed by rapid temperature increases and
snowmelt, sent excessive amounts of water through the pipe. (2) Two landslides occurred in the slope above the rupture.
Although the relative timing of the landslides and the breach is not known, the slides could have caused the pipe to
collapse or separate. (3) Erosion of particles near the pipe connections could have reduced the bearing strength of the
pipe. (4) The strength of an elbow in the piping may have been exceeded by massive and rapid fuid fow. In addition,
a sinkhole that developed from the rupture threatened the stability of the embankment. The sinkhole came within 100
feet of several upstream-constructed additions to the cross-valley embankment before stability was maintained through
mitigation of the breach.
The impoundment, upstream from several communities, was rated at the time as high hazard. A 50-mile stretch of Coal
River from Montcoal to its mouth at St. Albans was visibly affected, and fve water plants were shut down. Although 1,700
customers water supply was disrupted in the Racine Public Service District, no human injuries or fatalities occurred as
a result of this incident.
Source: Owens, 1987
January 28, 1994: Consolidation Coal Company,
Morgantown, West Virginia
On January 28, 1994, a 5-foot earthen berm failed at a slurry refuse impoundment at the Arkwright Mine in Granville,
West Virginia. Heavy rain and melting snow resulted in 30 inches of water collecting behind the berm; it was determined
that the 4-inch discharge pipe and rock underdrain at the site were insuffcient to prevent water accumulation. The
incident released 375,000 gallons of water into the town of Granville. Although no one was injured, three residences
directly downstream were damaged.
Source: Betoney, 1994
May 22, 1994: Martin County Coal Corporation,
Davella, Kentucky
On May 22, 1994, a breakthrough occurred at Martin County Coal Corporations Big Hollow slurry impoundment in
Davella, Kentucky. Nearly 32 million gallons of black water inundated an abandoned and sealed-off portion of the
mine. The breakthrough resulted either from collapse or water penetration of the Coalburg coal seam bordering the
impoundment. Slurry had been impounded 32 feet higher than the coal seams elevation. The mines 16-inch, concrete-
block seals held the black water inundating the mine, but water broke through portal seals and a coal seam outcrop
barrier. Although the slurry level dropped by 6 feet, the embankment structure was not damaged, and no injuries or
fatalities occurred.
Source: Stewart and Robinson, 1994
August 9, 1996: Lone Mountain Processing Incorporated,
St. Charles, Virginia
On August 9, 1996, there was a breakthrough at Lone Mountain Processings Miller Cove slurry impoundment. The
evening before the failure, approximately 2.75 inches of rain had fallen, and most of it within an hour and a half.
Approximately 1 million gallons of black water were released into Gin Creek through an abandoned mine. (Underground
mines had operated in areas adjacent to the impoundment from the 1920s to the 1980s.)
Excavation of the breach showed that the leak occurred in an area where available mine maps indicated a barrier of at
least 25 feet of solid coal between the outcrop and the underground mine workings. Further exploration revealed that the
barrier was in fact less than 2 feet thick. It is believed that hydrostatic pressure from the slurry opened cracks in the coal
seam and began a piping-type failure. The thin coal barrier was progressively eroded, allowing slurry to fow uncontrolled
into the abandoned mine.
Source: Michalek et al., 1996
October 24, 1996: Lone Mountain Processing Incorporated,
St. Charles, Virginia
On October 24, 1996, a second breakthrough occurred at Lone Mountain Processings Miller Cove impoundment, but in
another area of the abandoned mine. This release was more serious than the event in August 1996 because the water
contained more solids. Approximately 6 million gallons of water and slurry exited the abandoned mine into Gin Creek
and fowed 11 miles, where it entered the Powell Rivers North Fork. Reportedly, the river was discolored for more than
40 miles.
TABLE 2.2 COAL REFUSE DISPOSAL FACILITY FAILURES AND INCIDENTS
(Continued)
< PREVIOUS VIEW
215
Background and Characterization of Coal Refuse
MAY 2009
The failure resulted from two large sinkholes that had developed on the northwestern end of the impoundment. When
the site was excavated to locate the breach, it was determined that the slurry had entered through a fracture in the mine
roof that coincided with these sinkholes.
Source: Michalek et al., 1996
November 26, 1996: Consolidation Coal Company,
Oakwood, Virginia
On November 26 1996, the Buchanan No. 1 impoundment in Buchanan County, Virginia, failed. In the 1960s, the
Kennedy coal seam at the site had been excavated by both surface area mining and underground auger mining. After the
impoundment was constructed (1984), another company mining underground in the adjacent drainage area apparently
intersected the historic auger mine workings, providing a conduit for the slurry.
Coal refuse and slurry from the impoundment broke into an abandoned underground mine and discharged about 1,000
gallons per minute at its peak through two mine portals into the adjacent North Branch Hollow of the Levisa Fork of the
Big Sandy River. There was no detrimental impact on the embankment, and no one was killed or injured.
Source: Michalek el al., 1996
October 11, 2000: Martin County Coal Corporation,
Inez, Kentucky
On October 11, 2000, a coal waste impoundment of the Martin County Coals preparation plant near Inez, Kentucky,
released slurry containing an estimated 250 million gallons of water and 31 million gallons of coal waste into local
streams. Reportedly, the failure was caused by the collapse of the slurry pond into underground coal mine workings next
to the impoundment. The slurry broke through an underground mine seal and discharged from mine entrances 2 miles
apart into two different watersheds (Wolf Creek and Coldwater Fork).
Although no human life was lost, the release killed aquatic life along the Tug Fork of the Big Sandy River and its tributaries.
Public water supplies were disrupted when communities along the rivers in both Kentucky and West Virginia shut down
water plants to prevent contamination with black water.
Source: Various issues of the Herald Leader, the Courier-Journal, and the Charleston
Gazette from 2000 and 2001; National Research Council (2002)
TABLE 2.2 COAL REFUSE DISPOSAL FACILITY FAILURES AND INCIDENTS
(Continued)
(NATIONAL RESEARCH COUNCIL, 2002)
< PREVIOUS VIEW
Chapter 3
COAL REFUSE DISPOSAL
FACILITIES AND OTHER
IMPOUNDING STRUCTURES
3-1 MAY 2009
This chapter presents an overview of the types of coal refuse disposal facilities and other impound-
ing structures that are employed at mine sites and provides a discussion of terminology applied to
these embankments and impoundments. Refuse disposal facilities are generally classifed in the fol-
lowing terms: (1) refuse piles (non-impounding facilities), (2) impounding facilities, (3) slurry cell
facilities, and (4) underground injection facilities. Non-impounding refuse piles typically consist of
an embankment fll where drainage is directed away from the site without retention. These facilities
are employed for disposal of coarse coal refuse and dewatered fne coal refuse.
Impounding facilities have the capacity to retain water, sediment and/or fne coal refuse slurry. Under
MSHA regulations, a facility or structure requires an approved impoundment plan when it exceeds
a threshold height or has the potential for impounding a threshold volume of water and/or fne coal
refuse slurry and/or represents a potential hazard to miners. In practice, MSHA also considers any
potential hazard to the public downstream. While not always true, slurry disposal facilities typi-
cally exceed impoundment threshold parameters, resulting in their being subject to MSHA impound-
ment plan regulations. Slurry disposal may also occur in slurry cell facilities that do not exceed the
impoundment threshold height or volume. In addition to disposal in embankments, impoundments
and cells, fne coal refuse slurry may also be disposed in underground mine workings at under-
ground injection sites.
3.1 HAZARD POTENTIAL FOR DISPOSAL FACILITIES AND OTHER IMPOUNDING STRUCTURES
Coal refuse disposal facilities and other mining impoundments are evaluated relative to their hazard-
potential classifcation, which dictates the design criteria that must be incorporated in their planning,
development, and construction. While hazard-potential classifcation primarily relates to impound-
ments, it can also be applied to non-impounding facilities where dewatered fne coal refuse may
exhibit fowable characteristics afer disposal. Consistent with the hazard-potential-classifcation
system and criteria for dams in use by federal agencies (FEMA, 2004a), the three hazard-potential
classifcations for disposal facilities are as follows:
Low hazard potential Facilities where failure would result in no probable loss of
human life and low economic and/or environmental losses. Such facilities are usu-
ally located in rural or agricultural areas where losses are limited principally to the
owners property or where failure would cause only slight damage to farm build-
ings, forest and agricultural land, and minor roads.
< PREVIOUS VIEW
3-2
Chapter 3
MAY 2009
Signifcant hazard potential Facilities where failure would likely not result in loss
of human life, but can cause economic loss, environmental damage, or disruption of
lifeline facilities. Such facilities are generally located in predominantly rural areas,
but could be in populated areas with signifcant infrastructure and where failure
could damage isolated homes, main highways, and minor railroads or disrupt the
use of service of public utilities.
High hazard potential Facilities where failure will probably cause loss of human
life. Such facilities are generally located in populated areas or where dwellings are
found in the food plain and failure can reasonably be expected to cause loss of life;
serious damage to homes, industrial and commercial buildings; and damage to
important utilities, highways or railroads.
The purpose of hazard-potential classifcation is not to determine the likelihood of a failure occur-
ring, but rather to assess the potential impacts should a failure occur and to establish appropriate
criteria for use in the design and operation of the facility. Thus, more conservative design and
operations criteria apply as the potential for loss of life or property damage from failure increases.
For example, more subsurface exploration and material property testing is normally performed
for a facility with high hazard potential than for one with low hazard potential. An impound-
ment dam with high hazard potential would be designed to accommodate the probable maximum
food (PMF), while a dam with low hazard potential would be designed for a smaller storm event.
While this chapter introduces hazard-potential classifcation, specifc design criteria that are asso-
ciated with the hazard-potential classifcation are discussed in later chapters. The application of
the hazard-potential classifcation to storm water management is discussed in Chapter 5; stability
considerations are discussed in Chapters 6 and 7; and hydrology and hydraulic engineering issues
are addressed in Chapter 9.
Determination of possible damage due to failure of an impounding refuse disposal facility must be
based upon an evaluation of conditions for an appropriate downstream distance. This distance is
normally determined by performing a breach analysis that defnes an inundation area resulting from
a breach of the impounding embankment. This analysis is particularly important in mountainous
mining areas where complementary industrial, commercial, and residential developments are usu-
ally located in valley botom. Conditions downstream from the disposal facility may also be impor-
tant even if the facility embankment is incapable of impounding water. An example is failure of an
embankment that blocks a stream, temporarily forming a dam that impounds water that suddenly
fails, releasing a food wave of water and coal refuse.
Many mine impoundments currently under MSHA jurisdiction have underground mine works either
beneath or near the dam or reservoir. This creates a situation where a failure of the impoundment
may release fowable fne coal refuse and water into the underground mine workings. A release of
fowable fne coal refuse and water can also occur due to failure of natural ground or man-made bar-
riers, resulting in breakthrough into the mine workings. In the event of breakthrough, not only will
mine personnel potentially be endangered, but the water or slurry may subsequently discharge from
the mine and potentially afect an area diferent from that afected by a dam failure. MSHA addresses
this possibility as follows:
The ofcial hazard-potential classifcation for an impoundment facility is based on
the three classifcations discussed previously and is assigned regardless of whether
the potential hazard is from a failure of the dam or from a breakthrough into the
mine workings and subsequent discharge.
For the purpose of selecting appropriate design criteria for a coal refuse dam, the
hazard classifcation is based on the appropriate rating for the type of failure con-
< PREVIOUS VIEW
3-3
Coal Refuse Disposal Facilities and Other Impounding Structures
MAY 2009
sidered. For example, an impoundment could have a high-hazard-potential rating
based solely on the potential for breakthrough into the underground mine workings,
with low consequences associated with a failure of the dam itself. In such a case, the
dam embankment can be designed based on the low-hazard-potential classifcation
(e.g., the 100-year design storm), while the breakthrough evaluation and prevention
measures, with respect to the extent of exploration, testing, monitoring, etc., would
need to be appropriate for a high-hazard-potential facility.
3.2 COAL REFUSE IMPOUNDMENT FACILITIES
Coal refuse dams are typically designed using coarse refuse and soil or rock fll materials for con-
struction with placement of fne coal refuse slurry in the associated impoundment. The development
frequently initiates with a starter dam constructed of earthen materials, coarse refuse or a combina-
tion of both for the period of initial slurry disposal. Coarse refuse generated by the coal preparation
plant is then used to expand and raise the starter dam, thus providing additional disposal capacity
for the fne coal refuse slurry. Typical features and construction practices associated with a slurry
impoundment facility include:
Temporary surface drainage diversion away from areas of embankment dam
construction.
Removal and stockpiling of topsoil and soils for future reclamation.
Removal of unsuitable foundation materials that would adversely afect the con-
struction of the embankment dam.
Collection and conveyance of groundwater springs.
Construction of impoundment and embankment liners, as required by state
regulations.
Construction of the dam and emergency spillway to design grades to provide ample
freeboard for design storm runof storage and discharge capacity.
Construction of a foundation cutof, underdrains and internal drainage struc-
tures within the embankment dam to control seepage and phreatic surface
development.
Construction of decant pipe structures for removal of accumulated water (surface
runof and/or clarifed water) from the impoundment.
Construction and maintenance of haul roads on the embankment or adjacent areas to
transport refuse to the disposal area.
Grading of the refuse embankment surface to maintain drainage toward collection
and/or diversion ditches.
Grading and sealing of the refuse embankment for drainage control and minimizing
water retention on the embankment.
Collection of surface drainage on the embankment and delivery to sediment control
structures.
Reclamation of completed surfaces such as the embankment downstream face and,
at completion of slurry disposal operations, covering of the setled fne coal refuse
by evacuating accumulated water, covering the fne coal refuse with coarse refuse or
earthen materials, and placement of soil and topsoil, and vegetation.
A signifcant issue in the development of a slurry impoundment is the design and construction of
the starter dam. The size, materials, and hydraulic appurtenances associated with the starter dam
impact future operations of the disposal facility, require substantial engineering oversight during
construction, and can represent a signifcant part of the development cost for the facility. The sched-
< PREVIOUS VIEW
3-4
Chapter 3
MAY 2009
ule for construction of a starter dam for a new mine is normally critical, as the coal preparation
plant will not usually be able to operate at full capacity until there is a functioning slurry disposal
area that meets initial short-term design storm criteria. The starter dam must also be designed so
that it can be transitioned to meet long-term criteria. For existing preparation plants, the designer
should coordinate the closure of an existing impoundment with the activation of a new impound-
ment to avoid lapses in available refuse disposal capacity and to facilitate reclamation of completed
areas.
Coarse refuse embankment dams are usually large in cross section in order to accommodate the
production of refuse from the preparation plant. They may be constructed either as a homogeneous
or zoned embankment. If an embankment dam is constructed as a zoned embankment, it may be
built with earthen materials of fner gradation in upstream areas of the cross section and with coarse
refuse in downstream zones. While the term homogeneous is ofen applied to coarse refuse embank-
ment dams, the coarse refuse may exhibit variability in grain size and specifc gravity, but still have
adequate shear strength to meet design assumptions and be sufciently competent to support con-
struction equipment.
3.2.1 Size and Hazard Considerations
3.2.1.1 Impoundment Defnition
MSHA currently requires that plans be prepared and approved for the design, construction, opera-
tion and maintenance of structures that impound water, sediment or slurry if such structures:
Impound water, sediment, or slurry to an elevation of 5 feet or more above the
upstream toe of the structure and have a storage volume of 20 acre-feet or more; or
Impound water, sediment, or slurry to an elevation of 20 feet or more above the
upstream toe of the structure; or
Are determined by the MSHA District Manager to present a hazard to coal miners.
For purposes of determining inclusion under MSHA regulation, the height of an embankment dam
is measured from the upstream toe of the structure to the lowest point on the crest of the dam. Other
regulatory agencies may defne height diferently. Figure 3.1 illustrates this measurement for a cross-
valley impounding embankment. If the lowest point on the crest of the structure is the invert of a
properly designed open-channel spillway capable of conveying the maximum water fow from the
design storm with sufcient freeboard and erosion protection, then that point is the proper location
for the upper measurement. Where decant pipes and pipe/box spillways are used, the elevation must
still be measured to the lowest point on the crest of the structure or embankment, not to the invert of
the decant riser or spillway pipe.
The storage volume of a dam is calculated using either the invert elevation of the spillway or the
lowest point on the crest of the structure. However, the storage volume based on a measurement to
the invert of the spillway may only be used if two conditions are satisfed. First, the spillway must be
an open-channel confguration. Second, the open-channel spillway must be designed to convey the
fow from the design storm with adequate freeboard and must have appropriate erosion protection.
If either of these requirements is not met, the capacity of a dam must be determined based on a mea-
surement made to the lowest point on the crest of the structure without reference to the spillway. The
design of the open-channel spillway can be verifed using a food routing analysis of the design storm
(selected based on the purpose of the structure and projected hazard potential classifcation) through
the impoundments open-channel spillway and other outlet works (if applicable).
Small ponds are sometimes used for disposal of fne refuse slurry and are referred to as slurry cells.
These cells may be located within an embankment and sized and sequenced to preclude classifcation
< PREVIOUS VIEW
3-5
Coal Refuse Disposal Facilities and Other Impounding Structures
MAY 2009
of the embankment as an impoundment. However, there are also situations where the slurry cells are
of sufcient size and are operated such that the facility may be classifed as an impoundment. Section
3.4 addresses slurry cell facilities. If a pond or impoundment meets any of the preceding criteria,
the facility is subject to regulation by MSHA as an impoundment. It is important to note that state
regulatory agencies may defne structures that are regulated as dams using diferent dimensions or
defnitions.
3.2.1.2 Impoundments in Series
In the case of multiple impounding structures in series that individually do not meet the size cri-
teria cited in Section 3.2.1.1 and where a failure of one of these structures can result in the failure
of another, the cumulative storage capacity should be used when applying the impoundment size
criterion.
In the case of multiple slurry cells (addressed separately in Section 3.4) that individually do not meet
the size criteria cited in Section 3.2.1.1 and where failure of one slurry cell can lead to the failure of
others, or where a slope failure can result in the release of water, sediment or slurry from multiple
cells, the cumulative capacity of the potentially afected cells should be used when applying the
impoundment size criterion.
3.2.1.3 Incised Impoundments
An incised impoundment is one created by excavating below the natural ground surface, as illustrated
in Figure 3.2. An impoundment may be totally below natural ground, or an embankment may be con-
structed so that only a portion of the impoundment is below natural ground. For purposes of determin-
ing the height or storage volume of an impoundment with respect to the criteria presented in Section
3.2.1.1, the portion contained below natural ground is not included in the height or storage volume
calculation. However, even for an incised impoundment, if it is determined that the impoundment
could become a hazard, appropriate design and construction requirements should be implemented. In
situations where mining is planned or occurs near an incised impoundment, mine operators should be
sure that there is a sufcient thickness of undisturbed ground lef in place to preclude failure.
FIGURE 3.1 CROSS-VALLEY IMPOUNDING EMBANKMENT
SECTION A - A
AA
PLAN
HEIGHT
NOTE: HEIGHT DIMENSION SHOWN
SOLELY FOR DETERMINING
COVERAGE BY MSHA REG-
ULATIONS.
FIGURE 3.1 CROSS-VALLEY IMPOUNDING EMBANKMENT
FIGURE 3.1 CROSS-VALLEY IMPOUNDING EMBANKMENT
< PREVIOUS VIEW
3-6
Chapter 3
MAY 2009
Cases may arise where a surface mining pit is used as an impoundment. If any portion of the ground
creating the impoundment is mine spoil or fll, then that portion should not be considered to be
incised (MSHA, 2007).
3.2.2 Disposal Facility Confguration
Impounding coal refuse disposal facilities are classifed based on their confguration and develop-
ment/construction staging. The disposal facility confguration is dependent upon the terrain and
size requirements, and the development staging refects the general sequence or direction of dis-
posal activity. The facility confguration categories are: (1) cross-valley impounding embankment
(Figure 3.1), (2) incised impoundment (Figure 3.2), (3) side-hill impounding embankment (Figure
3.3), and (4) diked impounding embankment (Figure 3.4).
Planning and design of an impounding embankment generally involves distinct development/con-
struction stages that are associated with intermediate points in the facility construction and the gen-
eral timing and directional development of the disposal operations. These stages typically refect a
few months to a few years of disposal operation and usually refect the direction in which develop-
ment occurs. The direction of construction normally falls into two categories: upstream and down-
stream. Upstream construction, as shown in Figure 3.5, involves initial construction and placement
of coarse refuse in downstream areas to form the impoundment with sequential placement during
subsequent stages in upstream locations, typically at higher elevations. Downstream construction,
as shown in Figure 3.6, involves initial construction and placement of coarse refuse in upstream
areas with placement during subsequent stages in downstream locations. It is common to have both
upstream and downstream construction stages as part of a disposal facility design.
An intermediate development condition is centerline construction (which is essentially the same
as alternating upstream and downstream construction), where refuse stages are constructed both
upstream and downstream of the previous stage, with the crest of the two stages generally in
alignment, but separated by the elevation increment of the stage (Figure 3.7). This terminology for
SECTION A - A
FIGURE 3.2 INCISED-POND (IMPOUNDING) FACILITY
HEIGHT
A A
PLAN
NOTE: HEIGHT DIMENSION SHOWN
SOLELY FOR DETERMINING
COVERAGE BY MSHA REG-
ULATIONS.
FIGURE 3.2 INCISED IMPOUNDING FACILITY
FIGURE 3.2 INCISED IMPOUNDING FACILITY
< PREVIOUS VIEW
3-7
Coal Refuse Disposal Facilities and Other Impounding Structures
MAY 2009
impounding embankments is relevant primarily for identifying whether portions of the coarse refuse
embankment will be constructed on setled fne coal refuse slurry and the timing for reclamation of
completed surfaces of the embankment.
Upstream construction, and to a lesser degree centerline construction, with placement of coarse
refuse embankments on setled fne coal refuse, introduces stability concerns due to the potentially
low strength of the fne coal refuse during initial covering and the potential for seismically-induced
strength degradation. The feld exploration, testing and analysis, and design methodology and crite-
ria used for upstream construction plans are typically more extensive than for downstream or center-
line construction. On the other hand, upstream construction, when the fnal toe and lower elevations
of the embankment face are established early in the disposal facility development, allows concurrent
reclamation of the completed face and thus improved erosion and sediment control.
Downstream construction mitigates issues related to the stability of the embankment due to the sof
foundation conditions associated with fne refuse, but the reclamation of the embankment face is
delayed until much later in the facility life, thus requiring more comprehensive erosion and sediment
control measures.
At some sites, both upstream and downstream stages are constructed in order to: (1) ft site ter-
rain conditions, (2) meet disposal capacity requirements, and (3) provide multiple coarse refuse dis-
FIGURE 3.3 SIDE-HILL IMPOUNDING EMBANKMENT
FIGURE 3.3 SIDE-HILL IMPOUNDING EMBANKMENT
SECTION A - A
A A
PLAN
FIGURE 3.3 SIDE-HILL IMPOUNDING EMBANKMENT
< PREVIOUS VIEW
3-8
Chapter 3
MAY 2009
posal areas. This type of construction is sometimes referred to as the modifed upstream construction
method. Thacker (1997) recommends initiating upstream construction early in the project life, with
subsequent placement of coarse refuse in a downstream butress zone to allow pore pressure to dis-
sipate in the previously loaded fne refuse.
The decision whether to construct a refuse facility utilizing upstream, downstream and centerline
construction sequencing is also dependent upon the geometry of the valley and the refuse disposal
capacity requirements. For instance, downstream construction generally requires use of greater
quantities of coarse refuse to gain embankment elevation than upstream construction. Centerline
construction normally requires a greater amount of coarse refuse during the initial stages of the facil-
ity with lesser amounts required as construction progresses. Additional discussion of the advantages
and disadvantages of various types of embankment construction is provided in Chapter 5.
FIGURE 3.4 DIKED-POND (IMPOUNDING) EMBANKMENT
STAGE I
SLURRY DISCHARGED AT EMBANKMENT
FIGURE 3.5 UPSTREAM STAGING METHOD
STAGE II
STAGE III
FIGURE 3.5 UPSTREAM STAGING METHOD
FIGURE 3.4 DIKED-POND (IMPOUNDING) EMBANKMENT
SECTION A - A
A A
PLAN
FIGURE 3.4 DIKED-POND (IMPOUNDING) EMBANKMENT
FIGURE 3.5 UPSTREAM STAGING METHOD
< PREVIOUS VIEW
3-9
Coal Refuse Disposal Facilities and Other Impounding Structures
MAY 2009
3.2.3 Hazard Potential Rating
The determination of the hazard-potential classifcation for impoundments discussed in Section 3.1
is based on the probable loss of life in the downstream area and the presence of structures and infra-
structure that could be afected if the dam were to fail. It may be apparent from the presence of
nearby, downstream mining facilities or of-site occupied structures that a dam failure could result
in loss of life and that the classifcation should be high hazard. Since this is the most severe hazard
classifcation, no further analyses may be necessary for hazard potential determination, although the
evaluation of potential downstream impacts would still be required for preparation of an emergency
action plan. If the hazard potential classifcation is not apparent, dam breach analyses should be per-
formed, with evaluation of the resulting food inundation levels and fow velocities at downstream
structures to determine the potential for loss of life and severity of impact to structures and infra-
structure. Dam breach analyses are discussed in Section 9.9.
The assignment of high hazard potential principally hinges on the determination of probable loss of
human life. In situations where there is no probable loss of human life due to the dam failure, other
site-specifc factors are applied to determine the hazard potential. Estimates of economic loss and
damage to infrastructure can be generated to provide guidance in diferentiating between low and
signifcant hazard, but assessment of environmental damage is more controversial.
Where there are mine workings beneath or near a refuse embankment or impoundment, a release
of water and slurry into the mine could result due to failure of the natural ground or a man-made
barrier between the impoundment and mine workings. Such a potential failure is referred to as an
FIGURE 3.6 DOWNSTREAM STAGING METHOD
FIGURE 3.7 CENTERLINE STAGING METHOD
SLURRY DISCHARGED AT EMBANKMENT
STAGE I
STAGE II
STAGE III
STAGE IV
FIGURE 3.7 CENTERLINE STAGING METHOD
FIGURE 3.6 DOWNSTREAM STAGING METHOD
SLURRY DISCHARGED AT EMBANKMENT
STAGE I
STAGE II
STAGE III
STAGE IV
FIGURE 3.6 DOWNSTREAM STAGING METHOD
FIGURE 3.7 CENTERLINE STAGING METHOD
< PREVIOUS VIEW
3-10
Chapter 3
MAY 2009
impoundment breakthrough and not only poses a threat to mine workers, but the water and slurry
could discharge from the mine and potentially afect an area diferent from the one that would be
afected by a dam failure. Chapter 8 addresses the design requirements for this situation, and MSHA
(2007) provides the following guidance on hazard potential classifcation:
The ofcial hazard-potential classifcation for an impoundment is based on the three
classifcations indicated in Section 3.1 and is assigned regardless of whether the
potential hazard is from a failure of the dam or a failure into mine workings.
For the purpose of selecting the appropriate design criteria for the impounding
embankment, the hazard-potential classifcation is based on the appropriate rating in
the event of a failure of the embankment itself. For example, an impoundment could
have a high-hazard-potential rating based solely on the potential for a failure into
underground mine workings, but have low consequences due to a failure of the dam
itself. In such case, the dam can be designed based on low hazard potential (e.g., the
100-year design storm), while the breakthrough evaluation and prevention measures
(with respect to the extent of exploration, testing, monitoring, etc.) must be appropri-
ate for a high-hazard-potential site.
3.3 COAL REFUSE NON-IMPOUNDING FACILITIES (REFUSE PILES)
Coal refuse disposal facilities that do not retain water or slurry are considered to be non-impounding
structures and are also referred to as refuse piles. As indicated in the MSHA Coal Mine Impoundment
Inspection and Plan Review Handbook (MSHA, 2007), a refuse pile may have small isolated sediment
control facilities and cells for the disposal of flter cake, sediments, etc. provided that the size of these
cells would not result in the classifcation of the structure as an impoundment, their location would
not afect structural stability, and the confguration does not impede drainage (e.g., block a drainage
course). Where this material is not compacted in two-foot lifs, the disposal should be approved by
the MSHA district manager. In addition to flter cake, fne coal refuse slurry can also be disposed in
appropriately designed cells, as discussed in Section 3.4.
3.3.1 Coarse Refuse Embankments
If not part of an impoundment plan, a coarse refuse embankment is typically designed for separate
disposal of coarse coal refuse and fne coal refuse. Typical features and construction practices associ-
ated with a coarse refuse disposal facility include the following:
Surface drainage diversion away from the limits of the disposal embankment.
Removal and stockpiling of topsoil and soils for future reclamation.
Removal of unsuitable foundation materials that would adversely afect the con-
struction or stability of the embankment.
Collection of discharges from springs under the foot print of the embankment that
may adversely afect stability.
Placement and compaction of run-of-plant coarse refuse in lifs to the designed lines
and grades of the embankment.
Construction and maintenance of haul roads on the embankment for transporting
the refuse to the disposal area.
Grading and sealing of the refuse embankment to maintain drainage control and to
prevent water retention behind the embankment.
Collection of surface drainage on the embankment and delivery to sediment control
structures.
Reclamation of completed surfaces, consisting of placement of soil and topsoil and
re-vegetating.
< PREVIOUS VIEW
3-11
Coal Refuse Disposal Facilities and Other Impounding Structures
MAY 2009
Coarse refuse embankments are usually developed as homogeneous embankments, without multi-
ple zones. While the term homogeneous is applied to such embankments, the coarse refuse materials
may exhibit some variability in grain size and specifc gravity. Typically, however, coarse refuse has
the strength to support construction equipment and to meet the embankment design assumptions.
3.3.2 Combined Refuse Embankments
Combined refuse embankments are designed for disposal of coarse refuse and dewatered fne coal
refuse as a mixed material. While combined refuse embankments may be homogeneous, they can
also be constructed as zoned embankments where more workable refuse materials such as coarse
refuse are used (either periodically, or in sufcient continuous quantity) to construct a well-com-
pacted downstream shell of sufcient width (sometimes referred to as a structural zone) to contain
the combined refuse and to construct haul roads within the disposal area. Typical features of a com-
bined refuse disposal facility are similar to those for a coarse refuse embankment. Because of the
presence of dewatered fne coal refuse and overall weter material conditions, large disposal areas
are generally needed so that the dumped materials can drain before spreading and compaction, and
more extensive internal drainage systems may also be required. In some cases amendments may be
needed to stabilize the wet materials.
3.3.3 Segregated Refuse Embankments
Dewatered fne coal refuse can be disposed in segregated areas of a coarse refuse embankment.
This approach results in a zoned embankment in which coarse refuse forms a downstream zone
or shell that supports the overall stability of the embankment, and isolated areas are provided
within the upstream or interior of the embankment for depositing the dewatered fne coal refuse.
Typically, these segregated refuse embankments are designed with provisions for haulage routes
within the embankment and for containment of the dewatered fne refuse without creating depres-
sions that could be classifed as impoundments. Such containment structures require measures to
control surface drainage that collects in the dewatered fne refuse disposal area. Typical features
of segregated refuse embankments are similar to those for coarse refuse embankments, but with
more extensive internal drainage systems and control of drainage within the dewatered fne refuse
containment area.
3.3.4 Confguration and Development Staging
Non-impounding facilities are constructed with a range of confgurations and development staging.
The disposal facility confguration depends upon the terrain, and the development staging refects
the general sequence or direction of disposal activity. Non-impounding facility confguration catego-
ries are: valley-fll (Figure 3.8), cross-valley (Figure 3.9), side-hill (Figure 3.10), ridge-dump (Figure
3.11), and heaped (Figure 3.12).
Planning and design of a non-impounding embankment generally involves development stages
related to intermediate points in the facility construction and the general development direction of
the disposal operations. The development staging is characterized by the direction in which devel-
opment occurs. Upstream construction involves the construction and placement of refuse in down-
stream areas initially, with sequential placement during subsequent stages in upstream locations,
typically at higher elevations. Downstream construction involves the construction and placement
of refuse in upstream areas initially, with sequential placement during subsequent stages in down-
stream locations. This terminology for non-impounding embankments is more relevant for valley-fll
confgurations, where upstream and downstream directions refect the shape of the valley.
Cross-valley non-impounding embankments (Figure 3.9) can potentially impound water, and
they may be classifed as impoundments by MSHA if they can impound water for such a period
< PREVIOUS VIEW
3-12
Chapter 3
MAY 2009
of time that they can create a hazard. MSHA (2007) presents factors considered for such classifca-
tion. Abandonment of cross-valley non-impounding embankments can require substantial regrading
in order to satisfy long-term drainage concerns. Therefore, this type of embankment is generally
avoided.
3.4 COAL REFUSE SLURRY CELL FACILITIES
Small cells or ponds in a coarse refuse facility that receive fne refuse are commonly referred to as
slurry cells. Disposal of fne coal refuse using slurry cells has been implemented at sites where con-
PERMANENT DRAINAGE PIPE
OR CULVERT
A A
PLAN
FIGURE 3.9 CROSS-VALLEY, NON-IMPOUNDING EMBANKMENT
SECTION A - A
CULVERT
FIGURE 3.9 CROSS-VALLEY, NON-IMPOUNDING EMBANKMENT
FIGURE 3.8 VALLEY-FILL, NON-IMPOUNDING EMBANKMENT
FIGURE 3.8 VALLEY-FILL, NON-IMPOUNDING EMBANKMENT
SECTION A - A
AA
PLAN
STAGE III
STAGE II
STAGE I
UPSTREAM
STAGING
METHOD
FIGURE 3.9 CROSS-VALLEY, NON-IMPOUNDING EMBANKMENT
FIGURE 3.8 VALLEY-FILL, NON-IMPOUNDING EMBANKMENT
< PREVIOUS VIEW
3-13
Coal Refuse Disposal Facilities and Other Impounding Structures
MAY 2009
straints may deter or preclude the construction of a large slurry impoundment, and/or the quantity
of fne coal refuse slurry is relatively small and can be reasonably accommodated with a small cell
structure. Small cells are created, usually within the upstream zone of a coarse refuse embankment,
and are flled with slurry. The upstream zone and cells are contained by a well-compacted down-
stream shell (sometimes referred to as a structural zone). Once an individual cell is flled, it is allowed
to drain, and the fne coal refuse is covered with coarse coal refuse while another cell is operated.
Because of the typical small cell size, operation of multiple cells intermitently will facilitate setling
and draining of fne refuse and optimize the capacity of each cell.
Slurry cells may also be used to dewater fne refuse prior to fnal disposal. Instead of encapsulating
the slurry cell in a coarse refuse embankment, the fne refuse is excavated from the cell afer substan-
tial dewatering has taken place and then is typically mixed with coarse refuse prior to being incorpo-
rated into an embankment for fnal disposal.
3.4.1 Size and Hazard Classifcation
The slurry cell concept is typically focused on: (1) sequencing the construction of cells to match the slurry
generation rate, (2) accommodating drainage and covering of each cell, and (3) maintaining a total capac-
ity of all open cells at a level that does not meet impoundment classifcation or, if classifed as an impound-
ment, has a low hazard potential. In the case of multiple slurry cells where they individually do not meet
the size criteria requirements of 30 CFR 77.216(a), if the failure of one cell can result in the failure of
another or if a slope failure can result in the release of water, sediment, or slurry from multiple cells, then
the cumulative storage capacity of the afected cells is used for application of the impoundment size crite-
ria. This includes cases where the initial slurry cells are covered with coarse refuse and additional cells
are to be placed above the initial set of cells (MSHA, 2007).
The operation of multiple cells, particularly if subsequent cells are built on previously covered
cells similar to the upstream construction method for an impoundment, can lead to classifcation
as an impoundment. If multiple cells are arranged such that they afect structural stability and
there is a possibility of sequential failure and a large release with signifcant downstream impacts,
FIGURE 3.10 SIDE-HILL, NON-IMPOUNDING EMBANKMENT
A
A
PLAN
SECTION A - A
FIGURE 3.10 SIDE-HILL, NON-IMPOUNDING EMBANKMENT
FIGURE 3.10 SIDE-HILL, NON-IMPOUNDING EMBANKMENT
< PREVIOUS VIEW
3-14
Chapter 3
MAY 2009
the impounding facility could receive a signifcant- or high-hazard-potential classifcation, as dis-
cussed in Section 3.1. In such a case, the design storm requirement will increase, necessitating
greater diversion and freeboard, thus limiting the feasibility of the concept for some confguration
categories (e.g., valley flls).
The advantages of the slurry cell concept are the following:
As a low-hazard-potential facility with limited cell size, managing the design storm
runof can generally be accommodated with minimal storage, enabling more capac-
ity for disposal of slurry and clarifcation of process water. Control of surface drain-
age from the surrounding embankment area is critical, along with efective diversion
of runof from watershed areas around the facility.
Slurry cells have limited capacity, comprising generally a thin deposit of slurry
with coarse refuse containment and covering layers that allow the fne refuse to
dewater and consolidate, making the total mass less fowable. This limited capacity
is viewed as an advantage at sites that are undermined and have the potential for
breakthrough.
With the fnes compartmentalized in cells, a problem at one location is less likely
to afect the entire facility, which also helps to mitigate concerns for breakthrough
potential.
A A
PLAN
SECTION A - A
FIGURE 3.11 RIDGE (NON-IMPOUNDING) EMBANKMENT
FIGURE 3.11 RIDGE (NON-IMPOUNDING) EMBANKMENT
FIGURE 3.11 RIDGE (NON-IMPOUNDING) EMBANKMENT
< PREVIOUS VIEW
3-15
Coal Refuse Disposal Facilities and Other Impounding Structures
MAY 2009
As a low-hazard-potential facility, fewer and less extensive hydraulic appurtenances
are required, with lower associated construction costs.
The features of a slurry cell facility are similar to those discussed for a non-impounding facility. It
has become common practice to design a well-compacted shell (sometimes referred to as a structural
zone) at the face of the disposal facility embankment as the downstream containment structure for
cells. The design of the well-compacted shell is based upon evaluation of factors such as width, slope,
benches, internal drainage, and access or haul roads. Similar to non-impounding refuse embank-
ments, diversion of runof from upstream and hillside areas is accomplished by locating ditches and
channels, which are separate structures from the slurry cells, either above or on the embankment sur-
face. To keep runof from entering the disposal area, the construction and maintenance of the diver-
sion ditches requires sequential planning and periodic construction activities apart from disposal
operations. Additionally, control and discharge of process water from the slurry operation is required
at each cell. Typically, this will require small-diameter decant systems or stabilized channels. These
structures usually have limited service requirements because of the low capacity of the cells.
In addition to the critical sequencing of cells to match the fne refuse generation rate and maintain
low-hazard-potential classifcation, a slurry cell facility has many of the same design considerations
as impoundments: (1) geotechnical investigation to determine embankment and foundation charac-
teristics, (2) testing for relevant material properties including shear strength and hydraulic conduc-
tivity, (3) static and seismic slope stability analyses, (4) underdrains to control the phreatic level, (5) a
dam breach analysis, unless the hazard potential is otherwise apparent, to determine the appropriate
hazard potential rating and design storm, (6) instrumentation to confrm design assumptions and
performance, (7) preparation of construction specifcations, and (8) construction monitoring.
Some disadvantages of slurry cell facilities include requirements for: (1) frequent construction of
diversion ditches, new cells, and cell spillways (decant structures) as the site elevation increases, (2) a
FIGURE 3.12 HEAPED (NON-IMPOUNDING) EMBANKMENT
SECTION A - A
A A
PLAN
FIGURE 3.12 HEAPED (NON-IMPOUNDING) EMBANKMENT
FIGURE 3.12 HEAPED (NON-IMPOUNDING) EMBANKMENT
< PREVIOUS VIEW
3-16
Chapter 3
MAY 2009
relatively large ratio of coarse refuse to fne refuse, and (3) more detailed planning and supervision of
the site so that the construction, flling, and covering of cells is accomplished in the proper sequence.
Also, a slurry cell facility can potentially be reclassifed as having high hazard potential, increasing
the diversion and cell spillway requirements and impacting the long-term feasibility of the concept.
While one of the atractions of the slurry cell concept is disposal of fne refuse slurry in a structure
with low-hazard-potential classifcation and thus less stringent design storm requirements, it may
be advantageous to employ the concept at an impoundment site to mitigate potential impacts from
underground mining and breakthrough potential. In such a case, the embankment design would be
in accordance with the appropriate impoundment criteria, and additional operating plans would be
required for disposal of slurry within cells constructed in the impoundment area.
3.4.2 Disposal Facility Confguration and Development
Non-impounding facility confgurations established by MSHA may be developed using the slurry
cell concept, although the quantity of available coarse refuse typically requires a valley-fll or side-hill
confguration. The valley-fll confguration is generally developed in the upstream direction afer a
sufcient embankment height and top surface is reached to enable individual cells to be constructed.
Consequently, beginning a slurry cell system may require operation of another disposal facility (e.g.,
underground injection), an existing disposal facility (e.g., old impoundment), or available fll mate-
rial (e.g., mine spoil from other site development work) in order to achieve a sufcient working
surface and embankment confguration to initiate slurry cell operation. As the disposal embankment
is raised in height and additional cells are constructed over covered cells, hazard classifcation may
become an issue. Therefore, the use of a valley fll for a slurry cell facility will likely have limitations.
Figure 3.13 shows a slurry cell facility developed in a valley-fll confguration.
Use of slurry cells with side-hill and heaped confgurations is less common, unless backup disposal
capacity is required in conjunction with the underground injection of fne coal refuse. These con-
fgurations require a larger quantity of coarse refuse for development of the structural shell, but this
FIGURE 3.13 SLURRY CELL IN VALLEY FILL, NON-IMPOUNDING
EMBANKMENT
SECTION A - A
PLAN
A A
ACTIVE SLURRY CELL
FILLED AND COVERED
SLURRY CELL
FIGURE 3.13 SLURRY CELL IN VALLEY FILL, NON-IMPOUNDING EMBANKMENT
FIGURE 3.13 SLURRY CELL IN VALLEY FILL, NON-IMPOUNDING EMBANKMENT
< PREVIOUS VIEW
3-17
Coal Refuse Disposal Facilities and Other Impounding Structures
MAY 2009
3.5 UNDERGROUND INJECTION SITES
Fine coal refuse slurry can be disposed within underground mines by injection. Background informa-
tion for the technical feasibility of underground disposal can be found in a publication by the National
Academy of Sciences (NAS, 1975). There are state and federal regulatory programs for these operations,
and the Virginia Department of Mines, Minerals and Energy, Division of Mined Land Reclamation has
developed guidance for planning, design, operation and monitoring (VA DMLR, 2006). Planning and
design for underground slurry disposal is discussed in the following subsections.
3.5.1 Siting
Background information useful for planning and design includes property ownership, geologic strata,
hydrogeologic conditions, and data describing the mine workings to be used for disposal, including map-
ping, extraction information, mine dewatering and discharge conditions, proximity to active mine areas
FIGURE 3.14 SLURRY CELL SIDE-HILL FILL, NON-IMPOUNDING EMBANKMENT
SECTION A - A
FILLED AND COVERED
SLURRY CELL
ACTIVE SLURRY CELL
A
A
PLAN
FIGURE 3.14 SLURRY-CELL, SIDE-HILL, NON-IMPOUNDING EMBANKMENT
material may not be available considering the need for coarse refuse to construct individual small
cells. Figure 3.14 illustrates a slurry cell facility developed with a side-hill confguration.
Slurry cells may also be incorporated into an impoundment confguration, if issues of underground
mining and breakthrough potential cause concern over the quantity of fowable impounded material.
Figure 3.15 illustrates a slurry cell facility developed at an impounding embankment.
FIGURE 3.14 SLURRY-CELL, SIDE-HILL, NON-IMPOUNDING EMBANKMENT
< PREVIOUS VIEW
3-18
Chapter 3
MAY 2009
and associated barrier information (coal barrier and bulkheads). The capacity of the underground work-
ings should be determined, and the potential duration for slurry disposal should be estimated based upon
setling and the ultimate solids content of deposits. Consideration should be given to the following:
Lineaments, faults, and fractures as possible conduits of slurry.
Recharge and discharge of groundwater into the mine workings as an indication of
the hydraulic connection of the mine works with the fracture fow system.
Surface openings that could or will become a discharge point, and coal barrier
requirements and stability, with an assessment of the amount of weathering of out-
crop, jointing, and previous or planned surface or auger mining.
Presence of groundwater users, radius of infuence of withdrawal wells, and connec-
tivity with mine voids.
Mine works overlying and underlying the mine work receiving injected slurry,
including inventory and assessment of vertical dewatering holes, subsidence fractur-
ing, and associated surface openings.
Proximity of and impacts on active mines.
Slurry injection and movement through mine workings, including water balance
analysis of injection fows and surface and groundwater conditions.
Leaching of contaminants from emplaced slurry, including potential impacts from
process chemicals, and sorption characteristics of processing chemicals on coal and
sand/clay particles.
Hydrologic conditions that could develop afer cessation of injection, including
development of equilibrium groundwater conditions.
Potential impacts to near-surface groundwater resources
3.5.2 Injection System Design
The following issues related to injection system design should be addressed:
Drilling methods for penetrating the abandoned mine to reduce the potential of cre-
ating an ignition source.
Design of bulkheads and seals that may be required to control the deposition of
slurry or direct drainage toward acceptable discharge locations.
Injection site and slurry line/alignment, including construction and service access,
drainage control, and secondary containment.
Injection well design including maximum pressure (below level for hydraulic fracturing of
overburden or mine barriers, and system components such as casing), complete casing of
overburden (with double casing provisions when penetrating upper mine voids), casing
grouting methods through overburden, and wellhead completion with air gap to prevent
pressurization of the well, if applicable, or pressure control and monitoring. Where pres-
sure injection is required, a control system to limit pressures to the maximum design value
must be employed, along with backfow prevention in the event of slurry line rupture.
Secondary containment for piping and injection site, including monitoring and con-
trols that shut down the slurry pumps in the event of piping failure.
Operating procedures and fow rates, including hours of slurry disposal/process
water pumping, and measures to prevent the introduction of oxygen into the mine.
< PREVIOUS VIEW
3-19
Coal Refuse Disposal Facilities and Other Impounding Structures
MAY 2009
3.5.3 Risk Assessment and Response Plan
The following are issues that should be considered in the evaluation of risks associated with release
and potential response actions:
Avenues of release (both surface and underground) and human exposure pathways.
A concern is ingestion through drinking water wells with capture zones that may
draw water from underground mine works or be connected to mine works through
fracture systems.
FIGURE 3.15 SLURRY CELLS IN CROSS-VALLEY FILL IMPOUNDING EMBANKMENT
PLAN
FILLED AND COVERED
SLURRY CELLS
A A
ACTIVE SLURRY CELLS
ABANDONED MINE WORKINGS
IN COAL SEAM
SECTION A - A
FIGURE 3.15 SLURRY CELLS IN CROSS-VALLEY-FILL, IMPOUNDING EMBANKMENT
FIGURE 3.15 SLURRY CELLS IN CROSS-VALLEY-FILL, IMPOUNDING EMBANKMENT
< PREVIOUS VIEW
3-20
Chapter 3
MAY 2009
Blowout potential associated with the injection pressures, slurry and water accumula-
tion, and barrier stability. The barrier stability evaluation should refect subsequent
surface, auger, or highwall mining that may have thinned or penetrated the barriers.
Future restrictions on mining in barriers should be considered if such potential exists.
Environmental receptor identifcation relative to groundwater and surface water,
considering contamination potential due to acid forming materials and petroleum
solvents or other chemicals used in the coal preparation plant and solids deposition
within streams.
Response to release, including: (1) steps to determine the extent of release, (2) identi-
fcation of emergency containment and cleanup resources (in-house and contracted)
and associated steps to initiate action, (3) disposal of wastes generated during
cleanup, and (4) implementation of operational contingency plan for ongoing slurry
disposal.
3.5.4 Contingency Plan
Contingency disposal options should be established for fne coal refuse slurry if injection is sus-
pended for emergency or performance reasons. The plan should include steps to be taken to shut
down the current injection operation; facilities and designs for interim slurry disposal such as
alternate injection locations (in the event of performance based shutdown) or emergency ponds
and drying cells, dewatering equipment, etc.; and longer-term options for disposal such as slurry
disposal cells.
3.5.5 Monitoring Plan
Monitoring for slurry injection, emplacement, mine water levels and mine discharge rates/quality for
the injection target and adjacent mines, as applicable, should be developed, along with groundwa-
ter and surface water monitoring. The parameters, methods, and frequency for monitoring should
be established based on site conditions, sound engineering judgment, and applicable regulatory
programs.
Slurry injection monitoring for organic and inorganic parameters should be per-
formed to assess potential groundwater and surface water impacts, as required by
regulatory programs.
Slurry emplacement monitoring should include monitoring wells and discharge
points within the mine works for detecting the presence of slurry solids for compari-
son with predicted slurry and water accumulation and movement, and planning
subsequent injection sites.
Mine pool/discharge water monitoring should be performed and the data should be
compared to the maximum predicted mine pool level and discharge rates, and water
balance analysis.
Groundwater monitoring should include wells located in hydrogeologic units most
likely to be infuenced by the injection operation, such as the fracture fow system,
the subsidence fracture zone overlying workings, and the coal seam adjacent to mine
workings impounding the slurry. Lineaments should also be considered when select-
ing locations to be monitored. Potable wells within the expected zone of infuence of
the injection operation should be monitored, but these wells should not be consid-
ered as groundwater monitoring unless well construction details are known.
With the use of an underground injection site for slurry, the coal preparation plant will also require a
surface disposal facility for coarse coal refuse. As indicated above, contingency disposal options should
< PREVIOUS VIEW
3-21
Coal Refuse Disposal Facilities and Other Impounding Structures
MAY 2009
be developed if underground injection proves to be infeasible. Such options could include use of slurry
cells or conversion to combined refuse if the preparation plant is equipped for fne refuse dewatering.
3.6 RECOVERY (REMINING) OF COAL REFUSE DISPOSAL FACILITIES
Recovery (sometimes referred to as remining) of coal refuse from existing active or previously
abandoned embankments and impoundments is performed at some sites and involves the use of
advanced processing methods to obtain additional coal for fuel or power generation. In such situa-
tions, a ground control plan addressing safety issues associated with remining operations should be
submited to MSHA. The remining process involves the excavation and removal of coal refuse and,
where processing is performed on site, disposal of waste from the coal processing plant. Thus, a
modifcation of the existing site refuse disposal plan will generally be necessary.
Recovery or remining may involve exploration, excavation and handling, fnal grading, and recla-
mation on sof or loose materials including operation in wet, saturated or submerged conditions.
Exploration will typically include initial as well as periodic operational borings and test pits for
evaluation of geotechnical and groundwater conditions as well as marketability of the excavated coal
refuse. Low-ground-pressure equipment, upstream pushout of a coarse refuse pad, or barge opera-
tion may be needed depending on impoundment conditions, and procedures and safety precautions
appropriate to the method used should be developed.
Under relatively dry conditions, dozers, excavators, end-loaders, scrapers, or a clamshell can be used
for excavation of refuse and fnes from an impoundment. Procedures and safety precautions that are
suitable for the conditions encountered should be developed, and when sof or loose fne refuse is
present, stability analyses to determine if the excavation slopes will be stable should be conducted.
For impoundments built by the upstream construction method, a bufer zone or distance should be
maintained between the area of excavation and the upstream slope of the impounding embankment
or other fll that may be present.
Surface runof should be diverted around the operation area to the extent possible, and the active
work area should be graded so that water drains toward ditches and sumps. Lowering the phreatic
level in the fne coal refuse to the extent possible by surface runof control and removal of ground-
water will improve ground conditions for equipment operation, allowing maximum recovery of
coal. Consideration should be given to grading, deep sumps, and other dewatering measures for
lowering the phreatic level, and the phreatic levels should be continually monitored during opera-
tion. At sites where control and removal of surface and groundwater is not practical at all times
of the year, safe and economic removal of fnes may only be possible during dry periods. In such
instances, evaluation of the degree and depth of saturation of the fnes should be made prior to
resumption of operations.
When wet and submerged conditions are present, dredges and hydraulic sluicing (water cannons) are
sometimes employed to remove fnes from an impoundment. Accidents have occurred when steep-
ened slopes of fnes have collapsed, sending a wave of material across the impoundment. Besides the
operating concerns discussed above, additional safety concerns must be addressed, and it is impor-
tant to establish guidance for equipment operation. Land anchors used to position dredges should
be located on solid ground and kept well back from areas that might be susceptible to slope failure
as the dredging progresses. The safe distance that any operator or equipment not on solid natural
ground (including anchor equipment) approaches the working face should be evaluated and should
generally be limited to no closer than twice the vertical height of the working face. When dikes built
across the impoundment to divide it into sections are present, personnel and equipment should not
be allowed on dikes adjacent to active dredging operations until dredging has ceased and the stabil-
ity of the area has been assessed. Personnel and operating equipment must have the proper safety
equipment for working at a dredging operation.
< PREVIOUS VIEW
3-22
Chapter 3
MAY 2009
Upon completion of the recovery operation, fnal grading and reclamation should provide long-term
stability of slopes and drainage channels. This work is typically performed with conventional earth-
moving equipment, and many of the operating procedures and safety precautions for this phase of
recovery are similar to typical reclamation requirements.
For embankments or impoundments that are to be remined, engineering plans addressing the fol-
lowing should be prepared for review and approval:
Excavation of coal refuse, including the method of removal, temporary slopes,
sequence, and resulting confguration of the facility, with associated provisions for
controlling drainage and maintaining stability during operations and following
completion.
Slope stability analyses (for both permanent and temporary slopes) demonstrating
acceptable factors of safety where miners are subject to potential slope failure haz-
ards. The slope stability analyses should include consideration of rapid drawdown
conditions associated with the removal of fne refuse. If excavated slopes that are
intended to be temporary and are designed for short-term conditions remain in place
much longer than planned (or become permanent because of the idling of recovery
operations), they should be re-designed with long-term factors of safety.
Changes in seepage resulting from pooling of water as fne coal refuse is removed.
Operating procedures and precautions specifc to the excavation and recovery meth-
ods used (e.g., conventional construction equipment, barge and dredge equipment,
water cannon operation) to provide for safe access and mining.
Monitoring of compliance with the excavation plan and inspection of slopes and
drainage control structures.
Disposal plans for waste from the reprocessing of the coal refuse.
Reclamation and abandonment of the disposal facility.
If fne coal refuse is to be removed from a slurry impoundment, the potential impacts on the safety
of the impoundment should be addressed in the plans. This should include potential impacts on the
outlet structures and facility operation during the design storm and limitations on the extent and
slope of the excavation so that other parts of the facility are not compromised. For instance, if there
is upstream construction or barriers related to mine workings are built over slurry, removal of fnes
could remove support for the embankment crest or barrier structure. In such cases, an analysis for
determining the necessary bufer and excavation slope such that fnes recovery can take place with-
out compromising the stability of the embankment or barrier structure is required.
3.7 OTHER IMPOUNDING STRUCTURES
In addition to coal refuse disposal facilities, mine sites may have other impounding structures such
as fresh-water reservoirs to provide make-up water for the processing of coal and sedimentation and
treatment ponds to handle runof and drainage from refuse embankments, surface mined areas and
other disturbed surfaces. In some limited surface mining situations primarily in the western U.S.,
dams may be used for food control during the temporary period when the mine pit advances near a
water course. These structures are generally traditional dams and reservoirs that, while having many
of the same features as a slurry impoundment, may also have additional features associated with
infuent and efuent controls. Additionally, these impoundments are generally constructed during a
limited time period that may or may not be related to mining or coal processing rates.
Fresh water impoundments are typically located close to the coal preparation plant and are sized and
located within a watershed to provide an adequate quantity of process water. These structures are
< PREVIOUS VIEW
3-23
Coal Refuse Disposal Facilities and Other Impounding Structures
MAY 2009
typically dams, generally built of soil and rock fll, with primary and emergency spillways and piping
tied into the preparation plant. They are diferentiated from slurry impoundments by their design to
maximize water storage and the presence of gated spillways and distribution pipelines. Additionally,
they may have a signifcant depth of water and thus be subject to greater hydraulic head than would
be expected for a slurry impoundment.
Sedimentation ponds are required as part of erosion and sediment control measures for runof from
disturbed mining areas, including coal refuse disposal facilities. These ponds are smaller structures
than slurry impoundments and fresh water reservoirs, and they are sized to meet state regulatory
criteria based on the contributing disturbed area and other hydrologic factors. They are frequently
designed so as not to be classifed as an impoundment regulated by MSHA. They typically have pri-
mary and emergency spillways.
Treatment ponds are similar to sedimentation ponds and are used to treat drainage from disturbed
mining areas, including drainage from coal refuse disposal facilities and water pumped from under-
ground mine workings to meet suspended solids and water quality efuent requirements. In many
cases, chemicals are added to these ponds to neutralize acidic conditions and to precipitate metals
such as iron and manganese. Because they are used for treatment of drainage to improve water qual-
ity, the associated drainage area is limited and the surface runof entering the ponds is minimal. These
ponds are generally small structures, frequently below the MSHA size classifcation for impound-
ments. They typically have primary and emergency spillways, and gated controls are ofen part of
the primary spillway.
Flood-control dams are sometimes constructed in a water course to prevent or mitigate fooding
of a surface mine pit. The water course may have signifcant fow, particularly during thunder-
storms or periods of snowmelt. These dams may be small or large temporary structures, generally
located in the western U.S., and are required only during the period when the mine pit could be
afected by fooding from the water course. In situations where there is no threat to the public of
of mine property, it may be possible to design the dam using low- or signifcant-hazard-potential
hydrologic criteria, provided that a warning system and plan is developed and maintained for
notifying and evacuating personnel involved with the mining operation when the water behind
the dam reaches a specifed level. Table 3.1 presents guidance that should be considered or evalu-
ated as part of the MSHA impoundment plan if a warning system is being used to support the
selection of low- or signifcant-hazard-potential hydrologic criteria for a food-control dam at a
surface mine pit.
The geotechnical and hydrology/hydraulic engineering requirements for fresh water impoundments,
food control dams, sedimentation ponds, and treatment ponds are substantially the same as for
slurry impoundments, and design criteria are typically identical. Aspects of engineering analyses
and design that may be diferent for fresh water impoundments than for slurry impoundments are
addressed in other chapters of this Manual.
3.8 SMALL PONDS AND SIMILAR STRUCTURES
Some structures that are capable of impounding water or temporarily storing slurry on mine sites
may not exceed the threshold size that would require an approved MSHA plan, but they should be
designed in a manner consistent with the engineering guidance provided herein. Many sedimenta-
tion ponds and emergency slurry holding ponds at preparation plants are deliberately sized below
impoundment threshold criteria and generally do not have signifcant hazard potential. These struc-
tures are typically regulated by states as ponds or small dams and are designed in accordance with
applicable state criteria.
< PREVIOUS VIEW
3-24
Chapter 3
MAY 2009
TABLE 3.1 GUIDANCE FOR FLOOD WARNING SYSTEMS AT SURFACE MINE PITS
(1)
1. A warning system should typically consist of power supply equipment (primary and emergency back up),
water level monitors, and automated communications equipment. The entire system should be designed by
a qualifed engineer.
2. The evacuation warning level should be established based upon the potential time required for evacuation
of the surface mine pit and the potential rate of infow. Ideally, the warning should be triggered at a level
that allows for evacuation before overtopping of the dam by the PMF and resulting inundation within the pit
to a critical level. The food control structure should typically be maintained either in a dry condition of with
a limited water level and storage volume that would not present a hazard to downstream personnel. When
determining the water level for the warning system, any water stored behing the dam must be taken into
account.
3. Multiple warning levels should be considered. For example, in addition to the evacuation warning level, an
alert should also be issued at a lower level that would result in mine personnel coming to the dam to verify
that the system is working correctly and to monitor the situation.
4. While the mining operation is immediately downstream of the dam, the warning system should be tested on
a frequent basis and, if practicable, before expected large meteorologic events.
5. The warning plan should clearly defne the procedures to be followed when the warning system is activated,
and mine personnel should be trained on the appropriate response to a warning.
6. The warning plan should address the status of the dam after the mine pit has advanced beyond the infu-
ence of the water course. The plan should state whether the dam will be removed or will remain in place. If
the dam is to remain in place, it should be evaluated, as necessary, to verify that the hazard potential clas-
sifcation is appropriate for the downstream area potentially affected by a dam failure. The warning system
approach should only be in effect for the potential hazard posed to the mining operation for the temporary
period when the pit could be affected.
Note: 1. Use of a warning system must not be a substitute for appropriate dam design and construction. MSHA has
indicated that a warning system may be acceptable on a case-by-case basis for support of the use of low- or
signifcant-hazard-potential design criteria at food-control structures to prevent or mitigate fooding of a surface
mine pit. The guidance presented herein refects conditions that should be considered or evaluated as part of
an MSHA impoundment plan.
(FREDLAND, 2008)
< PREVIOUS VIEW
Chapter 4
PROJECT PLANNING
41 MAY 2009
Planning for a coal refuse disposal facility should be integrated with overall mine development and
operation plans. Procedures for the planning and design of safe, environmentally acceptable and
economical refuse disposal facilities are stressed throughout this and subsequent chapters of this
Manual. This chapter presents important planning and design sequences and procedures that can be
employed to optimize refuse disposal operations. Topics covered include:
Unique aspects of refuse disposal
Operations and site-related considerations
Sustainable mining practices
Economic considerations
Environmental and regulatory considerations
Planning sequence
Design sequence
Understanding these topics is essential to the design and construction of a refuse disposal facility that
imposes minimal restraints on the production of coal, assures the safety of miners and the surround-
ing public, and protects the environment. A refuse facility that meets these goals ultimately optimizes
control of the overall cost of mining operation.
4.1 UNIQUE ASPECTS OF REFUSE DISPOSAL
The disposal of coal refuse materials from preparation to transport and placement imposes problems
that are similar in many ways to more routine types of civil or mining construction projects. How-
ever, some aspects of refuse disposal may provide opportunities for optimization as follows:
Dissimilar refuse materials are continually generated. These include: (1) coarse
refuse in solid form, (2) fne refuse in slurry or dewatered sludge form, and (3) for
some sites, combined coarse and fne refuse. The characteristics and quantities of
coal refuse may change over time due to changes in geology, production, mining
methods, and coal preparation. Anticipating these changes in advance provides an
opportunity to develop a cost-efective disposal plan.
< PREVIOUS VIEW
42
Chapter 4
MAY 2009
Refuse materials can be disposed in a manner that allows for post-mining land use
that can enhance the future value of the property.
Coarse refuse, at times supplemented with amounts of borrowed soil and rock
materials, may be used to develop the structural capability to retain less carefully
placed or slurried refuse. In some instances, amendments such as industrial by-
products containing lime to address acid generation within the refuse are neces-
sary and these amendments can make up a signifcant percentage of the material
placed.
Co-disposal of coal combustion waste (fy ash and botom ash) with coal refuse may
be an efective contracting strategy or necessity for the mine, and this waste can be
used as an amendment to beneft the refuse disposal facility. Combustion waste can
be a signifcant percentage of the material placed and, like amendments, can alter
embankment material properties.
A disposal facility may be adapted to serve other operating requirements, such as the
recovery of clarifed water for return to the preparation plant.
Short- or long-term water impoundments may be created behind properly construct-
ed refuse embankments.
Some structural or hydraulic features must be adaptable to modifcation as the facil-
ity grows in size and shape over its service life.
Transport methods and equipment may change as the facility grows horizontally
and vertically.
Technological advancements, market demands and regulation changes may occur
within the service life of the facility.
Final portions of coarse refuse from normal operations, rather than borrow material,
may be used to make the completed facility structurally safe and environmentally
acceptable for abandonment.
Portions of refuse, properly mixed with appropriate amendments, may be used to
facilitate planting on a completed portion of the facility when reclamation soils are
scarce.
A major factor in many of the above items is the abundance of refuse material available for construc-
tion, at litle or no added cost. Proper planning must be employed to take advantage of the favorable
characteristics of refuse and to minimize the need for using more costly borrow soil and rock and
specially purchased construction materials.
4.2 OPERATIONS AND SITE-RELATED FACTORS
Each mine has characteristics that uniquely afect overall coal production and processing op-
erations. Proper planning on a site-by-site basis will assure that processing and refuse disposal
procedures that best suit the mining operation are established. It is possible, however, to develop
a list of considerations for the planning of practically all refuse disposal facilities, corresponding
to elements of federal and state permiting requirements. The length of this list and its subject
breadth illustrates the importance of meaningful coordination between management, operating
personnel and the design/permiting team. These factors can be grouped into two categories, as
follows:
Factors Primarily Related to Disposal
Capacity, number and location of potential disposal areas at the site.
General topographical, soils, geologic, seismological, hydrologic and cultural char-
< PREVIOUS VIEW
43
Project Planning
MAY 2009
acteristics of the site. Cultural characteristics may include buildings, infrastructure
(such as roadways, pipelines, power lines, etc.), and cemeteries.
Environmental and ecological conditions at the site, including surface water, ground-
water, wetlands, wildlife and plant species.
Mining (past, current, or potential future), oil and gas development, and other natu-
ral resources at the site. Analyzing the potential impacts from active or abandoned
underground mines that may underlie a refuse site is of utmost importance. Previ-
ous site development and refuse disposal activities should be identifed.
Capital requirements for initiating disposal at each disposal area.
Operating and maintenance costs at each disposal area.
The relationship between the capacity of each disposal area and its distance from the
preparation plant.
Proximity of population centers and residences.
Current property boundaries and land availability.
Land-use practices within and in adjacent areas and post-mining land-use consider-
ations.
Contingencies for unexpected conditions (i.e., changes in mine life resulting in
changes to the fnal embankment confguration and abandonment provisions).
Factors Primarily Related to Overall Operations
Public relations.
Source of coal and type of mining.
Extent of preparation of coarse and fne refuse.
Production rates of coarse and fne refuse and their reliability.
Changes in geologic conditions as the mine advances.
Anticipated life of the mine.
Geochemical characteristics of coarse and fne refuse.
Water requirements for the preparation process.
Modifcations needed to meet future market demands and regulation requirements.
Available and preferred materials handling equipment.
Flexibility in location of preparation plant.
Required infrastructure (roads, portals, slopes, shafs, conveyor beltlines, railroads,
etc.) during the life of the mine.
Additional materials generated at the mine by other operations or by nearby con-
struction projects (site developments, power plants, etc.) that might provide eco-
nomic value and be benefcial when combined with the disposed refuse.
Water treatment facilities required for operation or environmental control.
Corporate policy on capital expenditures versus operating costs.
Data relative to each of the above factors can be obtained through preliminary investigations and
discussions with appropriate mine personnel. This information can then be used to formulate specifc
studies leading to refnement of the project plan, performance of site-selection studies, and initial
design. As discussed in Sections 4.6 and 4.7, the designer must coordinate closely with the operating
staf during the early project stages and must incorporate environmental, permiting and regulatory
requirements into the facility planning.
< PREVIOUS VIEW
44
Chapter 4
MAY 2009
4.3 SUSTAINABLE MINING PRACTICES
Sustainable mining practices have evolved into management of natural resource development in a
manner compatible with the environment and the needs of the community in which mining plays a
signifcant role. By practicing sustainable development, mining companies can improve their access
to land and markets, as well as enhance their reputations and value, and thus promote the coal indus-
try. Yearly (2003) cites the following three tenets for sustainable mining:
Integrated approaches to decision making on a full, life-cycle basis that satisfy obliga-
tions to shareholders and that are balanced and supported by sound science and social,
environmental and economic analysis within a framework of good governance.
Consideration of the needs of current and future generations.
Establishment of meaningful relationships with key constituencies based on mutual
trust and a desire for mutually benefcial outcomes, including those inevitable situa-
tions that require informed trade-ofs.
Inherent in sustainable development is: (1) the commitment to implement efective planning and de-
sign practices and (2) the need for research and development leading to improved methods for har-
nessing natural resources with minimal impacts on the environment. Government regulation creates
the framework for protecting the environment from impacts, although it is the mining professionals
who are responsible for establishing the plans and designs that accomplish the mining operations in
a sustainable manner. This may require the involvement of a diverse professional community. Sus-
tainable mining goals established for engineers and leaders in the mining industry and detailed in
the Milos Statement from the International Conference on Sustainable Development Indicators in the
Mineral Industries held in Milos, Greece are discussed in Karmis (2003).
Project planning for coal refuse disposal should be integrated with overall development and opera-
tions planning for mining activities. In the following paragraphs, economic, environmental and regu-
latory considerations are addressed, along with planning and design sequence.
4.4 ECONOMIC CONSIDERATIONS
To achieve the desired optimum solution, economic factors must be considered during all phases of
disposal facility planning, design and implementation. Major economic decisions are required for: (1)
selecting the disposal facility site, (2) selecting the materials handling systems, and (3) for planning
the entire sequence of disposal operations from initiation through abandonment. The expenditures
required are usually classifed as:
Capital costs
Operation and maintenance costs
Reclamation (abandonment) costs
Potential long-term liability costs
The determination of these costs is part of the planning process and infuences the design, permiting
and site operations. Periodic updates of estimated costs should be part of the planning and design
process. Disposal facility planning and design should be keyed to the fnancial requirements of the
mine, so that cash projections cover the costs associated with operating and maintaining the disposal
facility. Carefully planned investment during initial phases of site development and selection of dis-
posal facility confgurations can have a major benefcial efect on subsequent operating and reclama-
tion (abandonment) costs.
Table 4.1 presents a general breakdown of various cost components associated with planning, design,
and implementation of a coal refuse disposal facility. These components are provided as a general
< PREVIOUS VIEW
45
Project Planning
MAY 2009
TABLE 4.1 COAL REFUSE DISPOSAL FACILITY COST COMPONENTS
I. Capital Cost
Property Acquisition
Permitting and Planning
Engineering and Design
Site Development
Access Roads Erosion and Sedimentation Control
Wetlands and Stream Mitigation Amendment Transport and Handling System
Site Preparation Starter Embankment
Perimeter and Diversion Ditches Internal Drainage System
Liner and Underdrain System Decant Installation Decant Installation
Mine/Entry Stabilization or Protection Spillway Construction
Refuse Transport and Handling System
II. Operating Cost
Equipment Operations
Refuse Transport (haulage & pumping systems)
Refuse Placement and Compaction
Facility Construction
Site Preparation for Facility Expansion Haul/Access Roads and Surfacing
Perimeter and Diversion Ditch Extension Internal Drains Installed in Stages
Mine/Entry Stabilization or Protection Decant Pipe Extension and Capping
Liner and Underdrain System Extension Spillway Extension
Amendment Materials Survey Control
Maintenance
Erosion and Sediment Control
Haul/Access Roads and Surfacing
Liner and Ditch Repairs
Regulatory Compliance and Reporting
Site Inspections and Testing
Monthly, Quarterly, and Annual Reporting
III. Reclamation Costs
Engineering/Permitting Oversight
Facility Construction
Impoundment Backflling and Stabilization
Site Grading
Cap/Soil and Topsoil Placement
Revegetation
IV. Potential Long-Term Liability Costs
Acid Mine Drainage Treatment
Mine Discharge
Erosion and Sedimentation Control
Revegetation
guideline for developing costs. Specifc site conditions and operational requirements may introduce
other elements not refected in the table.
< PREVIOUS VIEW
46
Chapter 4
MAY 2009
4.5 ENVIRONMENTAL AND REGULATORY CONSIDERATIONS
Environmental and regulatory considerations should be incorporated into the planning process,
and can be defned steps performed as part of, or in advance of, the permiting of the coal re-
fuse disposal facility with state and federal agencies. These considerations are addressed jointly,
because frequently the method for addressing environmental issues may be directed by regula-
tory guidance or law. Primary federal statutes and regulations include the Federal Mine Safety
and Health Act (1977) and the Surface Mining and Reclamation Act of 1977 (SMCRA), as well as
more widely applicable statutes and regulations such as the Clean Air Act and Clean Water Act.
SMCRA established the Ofce of Surface Mining (OSM) with authority to implement a national
program to regulate surface efects resulting from coal mining. State regulatory programs and
agencies may take on this responsibility, if approved by the OSM, which will remain in an over-
sight role. For the states that take on this responsibility, most of the regulatory requirements of
their programs are similar to the OSM regulations found in 30 CFR Chapter VII. In response to
local concerns and conditions, some state programs have requirements that are more extensive
than the federal rules. Additionally, some state agencies responsible for other programs, such
as dam safety, impose requirements for coal refuse disposal facilities. Designers should contact
state regulatory agencies to obtain current information on their regulatory programs and re-
quirements.
Environmental issues at some sites can require signifcant time and efort to address, afecting the
entire regulatory approval process. This section discusses environmental and regulatory consider-
ations in the planning process, and the engineering and design of specifc containment structures at
disposal facility sites are discussed in subsequent chapters of this Manual. However, the state and
federal permiting process, including environmental impact studies and mitigation requirements, is
lef for other publications. Because state regulatory input shapes many of the decisions associated
with planning and design of a refuse disposal facility, contact with these agencies and review of their
publications is essential.
Some specifc environmental and regulatory factors that infuence the planning and design of coal
refuse disposal facilities include the following:
Site selection process and permiting submitals prior to facility design. Some states
require a rigorous process for site selection as part of permiting new coal refuse
disposal facilities or the expansion of existing facilities.
Facility confguration (slopes, benches, crest width, etc.). Federal and state regula-
tions include specifc requirements for some facility parameters, either as part of
mining, coal refuse disposal, or dam permiting guidance.
Erosion and sedimentation and stormwater control structures. State regulations and
some local (municipal) entities provide specifc guidance for meeting erosion and
sedimentation control and stormwater requirements.
Liners for containment of refuse materials. Some state regulations provide guidance
for liners for refuse disposal facilities.
Amendments for neutralization of refuse materials. State regulations provide guid-
ance for acid neutralization of refuse materials.
Wetlands and stream encroachments. Federal and state regulations provide guidance
for addressing wetlands and stream encroachments, including mitigation require-
ments where necessary.
Prime farmlands. State regulations provide guidance for addressing the presence
of prime farmlands near proposed refuse facilities and include mitigation mea-
sures where necessary.
< PREVIOUS VIEW
47
Project Planning
MAY 2009
4.6 PLANNING SEQUENCE
Table 4.2 shows a typical sequence of events involved in the planning, design, operation and aban-
donment of coal refuse disposal facilities. Site-specifc factors and the specifc objectives of persons
collectively involved in the process may preclude direct application of the indicated sequence. How-
ever, proceeding generally in the manner shown should aid in the design, permiting and construc-
tion of a coal refuse disposal facility. Particular notation is made of the following items:
Continuing interaction between operations/mine personnel and the designer is
shown. The important relationship between coal production and refuse disposal
dictates this cooperation, if mine operation and facility construction are to proceed
optimally.
Interaction between the designers and the regulatory agencies relative to site selec-
tion and design elements is shown. This interaction will allow identifcation of spe-
cial regional or site-specifc concerns, methods to address concerns, and will gener-
ally facilitate the permiting process.
Engineering should be an integral part of disposal operations, as well as construction
monitoring and inspections, particularly on complex, long-term projects. Involve-
ment of the designer, or engineering personnel thoroughly familiar with the design
requirements, allows review of performance information relative to design param-
eters and identifcation of needed adjustments.
The next to the last step (XI) in Table 4.2 is periodic review of mine operations and
the disposal facility development. Regardless of the accuracy of initial planning, the
typically long active period of use of a disposal facility makes it likely that unantici-
pated changes to the original design will occur. Periodic review of the efects of these
changes must be performed so that continued safe, economical and environmentally
acceptable refuse disposal can be achieved.
4.7 DESIGN SEQUENCE
The design sequence presented in Table 4.3 details the technical aspects of the total planning-imple-
mentation process for disposal facility development in a safe and environmentally acceptable manner.
The primary design emphasis occurs during the third to seventh steps of the planning sequence in
Table 4.2. The elements of coal refuse disposal facility design and their general interrelationship are
presented in Chapter 5. Subsequent chapters of this Manual present detailed technical information
and procedures related to completion of facility design.
Table 4.3 presents a typical sequence of design steps and provides a checklist of the most important
items requiring consideration for a typical refuse disposal project. Reference is made in the table to
sections of the Manual where additional detailed discussion is presented.
The items in Table 4.3 may not be totally applicable in all instances. In the case of small facilities, a
very detailed study may not be appropriate, and portions of the investigation and analyses can pos-
sibly be eliminated by simply using conservative design assumptions. On the other hand, conditions
may be present at other sites that require studies beyond those identifed in Table 4.3. The existence
of such special conditions can only be determined by an experienced designer through careful study
of site conditions.
Item VIII in Table 4.3 summarizes the deliverable products of the design sequence, including the
design report and preparation of facility plans and specifcations. General guidance on the content of
these documents is presented in Chapter 11; however, site-specifc issues may require elements not
identifed in the table.
< PREVIOUS VIEW
48
Chapter 4
MAY 2009
Preparation of a design report is suggested for all facilities, regardless of the required extent of investiga-
tion and analyses. This report: (1) provides the opportunity for all parties (designer, operator and regula-
tory groups) to understand design assumptions and limitations, (2) provides the opportunity for the de-
signer to clearly state validating design assumptions during the construction or operations phases, and (3)
helps to avoid confusion or misunderstanding between the designer, operator and regulatory groups.
Clear and accurate plans and specifcations are essential for operations personnel to properly construct
and develop a disposal facility and for the operators quality control representatives to verify that all
details are completed correctly. Accurate plans and specifcations minimize the potential for misunder-
standings that could cause schedule delays during the review period and during construction. General
guidance for the content of refuse facility plans and specifcations is presented in Chapter 11 of this
Manual; however, there may be site-specifc elements that are not addressed in Chapter 11.
TABLE 4.2 PLANNING STEPS FOR COAL REFUSE DISPOSAL FACILITY DESIGN
Planning Steps Participants
Regulatory
Involvement
I
Gather and Evaluate Mine Development and Disposal
Information
Management
Operations
Engineering
No
II
Perform Initial Siting Studies, Develop Disposal
Concepts and Conduct Alternatives Analyses for
Potential Sites
Operations
Engineering
Yes
III
Perform Detailed Investigations and Prepare
Preliminary Design
Engineering No
IV
Modify to Best Suit Operations and Evaluate
Development Costs
Operations
Engineering
No
V
Confrm Preliminary Design Assumptions and
Regulatory Criteria
Management
Operations
Engineering
Yes
VI
Finalize Plans, Specifcations and Permit Documents;
Refne Development Costs
Engineering No
VII Final Approval
Management
Operations
Engineering
Yes
VIII Implement Site Development
Operations
Engineering
Yes
IX Conduct Disposal Operations
Operations
Engineering
No
X Construction Monitoring, Maintenance
Operations
Engineering
No
XI Periodic Inspections and Review of Operations
Management
Operations
Engineering
Yes
XII Abandonment and Reclamation
Operations
Engineering
Yes
< PREVIOUS VIEW
49
Project Planning
MAY 2009
TABLE 4.3 TYPICAL DESIGN SEQUENCE
(1)
Subject Manual Section
I. Site Contour Survey Data
A. USGS Topographic Maps 6.4.1.1
Hazard potential
Watershed area
B. Aerial Photography 6.4.1.4
Embankment and pond area: fve-foot minimum contour interval (two-foot interval is
preferable)
Spillway and outlet works: fve-foot minimum contour interval (two-foot interval is
preferable)
II. Review of Available Publications and Data
A. Soils and Geology Chapter 6
Government agencies:
United States Geological Survey (USGS)
Natural Resource Conservation Service (NRCS, formerly SCS)
State geologic survey(s)
Other Sources:
Universities
Studies from nearby sites
Coal mine exploration borings
Aerial photographs
B. Hydrology Data Chapter 9
Government agencies:
National Weather Service
USGS gaging station data
NRCS runoff data
C. Mining Status Chapter 8
Coal operators records
Government agencies:
Offce of Surface Mining, U.S. Department of the Interior
State agencies
D. Seismicity 7.7
U.S. Geological Survey (USGS)
Published records of recorded earthquake epicenters
III. Site Reconnaissance
A. Area Upstream and/or Upgradient and Downgradient from the Facility 6.4.2, 14.4.2
B. Reconnaissance of Natural Features 6.4.2
Topography and morphology
Soil conditions, rock outcrops, sinkholes
Vegetation cover
Drainage patterns, springs and streams
Erosion
Stability (sliding and sloughing)
Wetlands
C. Reconnaissance of Man-Made Features 6.4.2
Roads and railroads
< PREVIOUS VIEW
410
Chapter 4
MAY 2009
TABLE 4.3 TYPICAL DESIGN SEQUENCE
(Continued)
Subject Manual Section
Buildings and other structures
Bridges
Stream modifcations and channels
Mine entrances and features (shafts, boreholes, highwalls, auger holes, spoil/refuse,
AMD discharges, mine subsidence features, etc.)
8.3
Other infrastructure (gas wells, pipelines, power lines, etc.)
Water treatment facilities
IV. Site and Facility Confguration Selection
A. General Considerations
Chapters 3, 5
B. Geotechnical Considerations Chapters 6, 7, 8
C. Drainage Considerations Chapter 9
D. Environmental Considerations Chapter 10
E. Equipment and Construction Considerations Chapter 11
V. Field Investigations
A. New Facilities
Borings Locations, depths and types of sampling should be selected by an
experienced designer.
6.4.3
Geophysical surveys Supplemental to borings for exploration and testing. 6.4.4, 8.3.2
Piezometers Normally piezometers are required only in selected borings for
monitoring water levels in abutments, etc. or to determine if water conditions will
create construction diffculties.
6.4, 13.4.2
Test pits Test pits can be used to economically gain additional data in area
planned for facilities and adjacent areas where borrow material may be obtained.
6.4.3.3
Water samples Normally from surface water streams, major springs, and ground
water wells.
10.3
Field mapping Notation should be made of major spring locations, rock outcrop
zones, mine openings, evidence of subsidence, existing landslides and any other
conditions that might affect construction of refuse facility structures.
6.4.2, 8.2
B. Existing Facilities
Borings Locations, depths and types of sampling should be selected by a qualifed
engineer.
6.4.3
Geophysical surveys Supplemental to borings for exploration and testing. 6.4.4, 8.3.2
Piezometers Since phreatic conditions in an existing facility cannot be accurately
estimated, piezometers should be installed in borings critical to stability analyses.
6.4, 13.4.2
Test pits Test pits can be used for the same purpose as for new facilities, plus to
evaluate the nature of weathering with depth, to obtain large samples of existing
embankment materials, and to conduct in-situ density tests.
6.4.3.3
Water samples From stream above and below facility, from the impoundment
(if any) and from all major downstream seeps or springs. Also, from groundwater
wells.
10.3
Field mapping Same as for new facilities, plus observation should be made
to locate any seepage zones on downstream face and to note any evidence of
cracking or movement on any portion of the existing facility.
6.4.2, 8.2
Flow measuring weirs Weirs or calibrated pipes should be installed to monitor the
fow from any major seepage zone noted on the downstream face of the facility.
13.4.3
Survey monitoring If any evidence of movement of the existing facility is noted in
a critical zone, survey monuments can be installed to determine if the movement is
active and the rate of movement.
13.4.1
< PREVIOUS VIEW
411
Project Planning
MAY 2009
TABLE 4.3 TYPICAL DESIGN SEQUENCE
(Continued)
Subject Manual Section
VI. Laboratory Investigations
A. Soils/Rock and Refuse Materials 6.5
General and index properties:
Water contents of all samples
Grain-size analysis of representative samples
Liquid and plastic limits of fne-grained materials
Specifc gravity of refuse materials
Materials behavior:
In-situ properties
On Shelby tube samples or prepared samples for approximating in-situ
conditions
Strength tests with representative pore-pressure conditions
Consolidation of fne-grained materials
Hydraulic conductivity tests
Materials to be used for construction
Compaction tests
Strength tests with representative pore-pressure conditions
Consolidation of fne-grained materials
Hydraulic conductivity tests
Special tests for refuse materials:
Ash content and ignition tests
B. Water Quality Testing Chapter 10
The water quality testing program will be a function of the potential source of
water, environmental conditions, and specifc facility design features (may require
determination of pH, temperature, specifc conductance, suspended and dissolved
solids, sulfates and metals).
VII. Analyses and Design
A. General Considerations
Hazard classifcation Chapter 3
Site development and startup Chapters 4, 5
Erosion and sedimentation control Chapters 5, 6, 9
Staging Chapters 5, 6
Reclamation and abandonment Chapters 5, 6, 10
B. Geotechnical
Parameters Establish material properties from feld/laboratory data and/or other
sources. For new facilities the properties of refuse may have to be estimated
based on coal seam and preparation procedures because the material may not be
available.
6.4, 6,5
Geometry Establish basic embankment cross section confguration for each stage,
including methods for controlling seepage.
Chapters 5, 6
Static stability Estimate pore pressure conditions for critical stages or conditions.
Modify geometry if required to achieve satisfactory static stability conditions.
6.6.4, 6.6.5
Dynamic stability Perform seismic hazard assessment, liquefaction, stability, and
deformation analysis, as required, considering hazard classifcation of embankment
and seismic risk zone of site. Modify geometry if required to achieve satisfactory
dynamic stability.
Chapter 7
Settlement Evaluate if settlement of soft layers could cause loss of freeboard,
embankment cracking or damage to drainage facilities.
6.6.3
< PREVIOUS VIEW
412
Chapter 4
MAY 2009
TABLE 4.3 TYPICAL DESIGN SEQUENCE
(Continued)
Subject Manual Section
Subsidence Evaluate if underground mining could affect stability. Chapter 8
Buried pipe design Analyze pipe stresses and strain, select materials, and design
installation and backfll requirements for conduits and decant pipes.
6.6.6
Special considerations:
Determine special foundation preparation requirements, including treatment of
existing mine spoil or refuse.
Chapters 6, 8
Design special material requirements such as starter embankments, drainage
materials, flters and soil cover or intermediate layers, if any.
Chapter 11
Establish special construction requirements for diversion ditches, spillway cuts,
structure and pipe foundations, etc.
11.7
Specify borrow areas for required soils and rock materials. 6.2.3.4
Specify construction procedures to satisfy design assumptions. 11.1
C. Hydrology and Hydraulics Chapter 9
Determination of design storms 9.5
Establish requirements for critical stages such as facility startup or
abandonment.
Establish basic storage-decant-spillway scheme
Sedimentation Control 9.4.4
Establish sedimentation pond requirements.
Decant and spillway systems
Perform hydrology analyses for storm served only by storage and decant 9.6
Design decant system 9.7.4
Perform hydrology analyses for maximum design storm
Design spillway structures (or cuts) for various stages 9.7
Design erosion protection system and/or stilling mechanisms.
Diversion Systems
Design size of diversion ditches for various stages, including abandonment.
Chapters 9, 10
D. Special Considerations
Environmental protection Evaluate potential acid generation and seepage quality,
and determine containment or neutralization requirements.
Chapter 10
Corrosion Evaluate probable seepage quality and related limitations on
construction materials.
6.5.2.5, 6.6.6.1,
11.7
Vegetation Determine requirements for vegetation of completed surface, including
the need for soil cover or surface preparation and treatment.
Chapter 10
Monitoring Design monitoring and inspection program commensurate with the total
design.
Chapter 12
VIII. Preparation of Design Documents
A. Designers Report Contents Chapter 11
Introduction
History (for existing sites)
Discussion of site conditions and previous mining/site development/refuse disposal
Field investigations
Laboratory testing
Geotechnical analyses
Hydrology and hydraulics analyses
Special considerations
Facility staging
< PREVIOUS VIEW
413
Project Planning
MAY 2009
TABLE 4.3 TYPICAL DESIGN SEQUENCE
(Continued)
Subject Manual Section
Recommended design
Abandonment requirements
Monitoring and inspection
B. Plans Chapter 11
Location map
Location plan on USGS quadrangle base
Plan of borings and feld investigation
Boring and test pit logs
Laboratory data
Hydrology data
Results of stability analyses
Plans of facility at critical stages
Plan and cross sections of hydraulic structures
Details of hydraulic structure components
Capacity curves for control sections of hydraulic structures
Details of monitoring installations
C. Specifcations Chapter 11
Site preparation
Foundation preparation
Embankment construction and internal drainage facilities
Surface drainage facilities
Decant system
Emergency spillway construction
Instrumentation
Reclamation and abandonment
D. Calculation Brief (should include input fles for computer runs) Chapter 11
Coal refuse production rates (by weight and volume)
Starter embankment and staging
Hydrology and hydraulics analyses (sedimentation control, design storm routing,
drainage channel design, dam breach analysis)
Stability analyses (including seismic hazard assessment and stability analysis)
Settlement analyses
Seepage analyses and internal drain design
Surface drainage channel lining design
Buried pipe analyses
Environmental analyses
E. Operation and Maintenance Plan Chapter 11
Note: 1. The design procedure should be modifed appropriately to suit the size, arrangement and specifc
characteristics of the facility being designed.
< PREVIOUS VIEW
Chapter 5
COAL REFUSE DISPOSAL
FACILITY DESIGN
COMPONENTS
5-1 MAY 2009
Coal refuse disposal facility design involves the evaluation of a number of interrelated components
for development of an embankment that can accommodate refuse generation rates, storm water and
environmental control requirements and provide a stable structure consistent with the operational
demands of the mine and post-mining, land-use intentions. The detail and extent of analysis required
for evaluation of each of these components depends upon the type of refuse facility being considered.
The process of designing a disposal facility normally starts with determination of the refuse disposal
volume requirements and, for slurry impoundments, includes consideration of the refuse generation
rates and the design storm runof volume that may need to be temporarily retained. Foundation con-
ditions and possible required treatment, either resulting from existing soil conditions or past mining
practices (underground or surface), may infuence the locations of disposal facility structures and the
overall confguration of the site, resulting in a need for special structures and instrumentation. The
following components are normally evaluated as part of the design of a refuse disposal facility:
Disposal capacity requirements, facility confguration/staging and scheduling
Design storm management and erosion and sediment control
Facility confguration, geometry and stability
Foundation and mine subsidence considerations
Stability of all embankment, channel and other site slopes afected by construction
Internal drainage for embankment/foundation seepage control
Decant and emergency spillway requirements
Surface drainage controls
Instrumentation and monitoring requirements
Reclamation, abandonment and post-mining land use
5.1 DISPOSAL CAPACITY REQUIREMENTS AND SCHEDULING
5.1.1 Refuse Generation Rates and Design Capacity/Life
Refuse generation rates are the basis for determining disposal facility construction scheduling and
the life of a disposal site. The refuse generation rate can be estimated based upon process fow and
coal preparation studies for the new mine development and sampling, testing and volume data from
existing disposal sites. Process fow data typically provide refuse generation rates for coarse refuse
< PREVIOUS VIEW
5-2
Chapter 5
MAY 2009
and fne refuse separately on a dry-weight basis, along with estimates of the moisture content (either
as percent moisture or percent solids for slurry). To estimate the disposal capacity and life of a facil-
ity, a conversion of the process data on a weight basis to in-place disposal data on a volume basis is
required. Where sampling and testing data are not available, estimates of the in-place refuse proper-
ties must be developed, typically by using data from similar processing plants, coal strata, and refuse
disposal sites. The in-place refuse properties required for initial estimation of the life and scheduling
of a facility normally include unit weight/density and moisture content of both coarse and fne refuse.
Ultimately, engineering properties of the refuse will also be required for designing the facility. These
properties should be verifed during the initial phases of the operation of the facility and at periodic
intervals during the facility life. Available test data should be scrutinized to verify that they are rep-
resentative, and the reliability and potential for variability of the refuse material should be evaluated.
Provisions for verifcation of material properties and a schedule for doing so should be indicated in
the design plans.
The design capacity and construction scheduling for a disposal site is also dependent on the planned
confguration of the facility, as well as the type of handling of the fne refuse (slurry, flter cake, or
mixed with coarse refuse). Based on an assumed site confguration, the capacity and life can then be
estimated based upon the computed volume and refuse generation rates.
While the type of handling and processing used for the fne refuse is typically an operational deci-
sion, physical characteristics of fne refuse when disposed may impose certain limitations at the dis-
posal site. For instance, past undermining may preclude the option of a slurry impoundment or may
entail extensive foundation treatment that makes that option infeasible. Therefore, knowledge of the
planned type of handling proposed for the fne refuse prior to selecting a site confguration for the
refuse facility is important.
5.1.2 Slurry Impoundment Staging and Scheduling
The staging and scheduling of the construction of a slurry impoundment must be carefully analyzed
so that the embankment elevation is sufcient to provide storage to contain the slurried fne refuse
generated as well as the runof from a design storm. Additional embankment height to account for
material consolidation or permanent seismic deformations may also be needed. In general, addi-
tional embankment height and width can be added to provide design conservatism and enhance
safety. Although the overall capacity of a slurry impoundment disposal site confguration can be esti-
mated from the refuse generation rates, the rate of incremental construction of embankments or dikes
must be sufcient to accommodate the generation and deposition of fne refuse slurry (referred to as
staging). Additionally, storm runof entering the impoundment will need to be stored temporarily,
increasing capacity requirements. Short-term design storm criteria are applicable during initial site
construction with transition to long-term criteria during subsequent development. Other factors that
may enter into the staging of a slurry impoundment include the following:
Initial construction of the beginning embankment or dike for impounding slurry
(termed starter dam, starter embankment or starter dike) typically requires
borrow material. The size of this embankment is optimized based upon the avail-
able borrow material, topography of the disposal site (which afects the initial slurry
capacity), and the ability to schedule subsequent embankment or dike construction
using solely coarse refuse to impound slurry at the expected refuse generation rates.
Therefore, the reliability and consistency of refuse generation rates is important to
the construction of embankment stages and operation of the impoundment.
Availability and capacity of existing disposal facilities at the mine that could be
used for temporary disposal during site development activities and while the starter
embankment is being constructed.
< PREVIOUS VIEW
5-3
Coal Refuse Disposal Facility Design Components
MAY 2009
Amendments to the refuse or co-disposal of combustion waste with refuse, particu-
larly when such additional material is a signifcant fraction of the refuse.
The type of construction staging (upstream, downstream and centerline) will afect the
capacity and schedule for the facility. Upstream construction utilizes less coarse refuse
to achieve a comparable impoundment capacity because the stages are constructed
partially on the setled fne refuse in the impoundment. Downstream, and to a lesser
extent, centerline construction staging require greater quantities of coarse refuse to
achieve equivalent impoundment capacity in comparison to upstream staging. At
some sites a combination of upstream and downstream staging is employed to meet
site confguration or specifc refuse disposal requirements and to control excess pore
pressures by initiating loading and consolidation of the fnes early in the project life.
In areas subject to seismic loadings, where deformation of the embankment may
occur, additional freeboard between the impoundment surface and embankment
crest may be necessary.
Other special considerations such as impoundment breakthrough potential into
underground mines, mine subsidence, reclamation, and post-mining land use can
also infuence the staging and confguration of the disposal facility.
When determining the type of construction staging (downstream, upstream, centerline, or a com-
bination thereof), the ratio of coarse refuse to fne refuse generated by the preparation plant is a
major factor. When the coarse refuse to fne refuse ratio is relatively high, downstream and centerline
staging is normally preferred. As the coarse to fne ratio decreases, upstream staging is normally
employed so that sufcient embankment height and containment is provided. As indicated above,
both upstream and downstream staging may be employed at a site. Figure 5.1 illustrates a construc-
tion sequence with upstream and downstream staging through Stage II. In this example, an early
stage of upstream construction (Stage II-A) is incorporated in order to initiate consolidation of fne
refuse, while disposal simultaneously occurs in a downstream stage (Stage II-B). The fgure also
shows several of the key components in facility design.
The amount of refuse generated and the coarse to fne ratio must be confrmed from actual produc-
tion measurements and must be rechecked periodically during operation. These parameters should
be periodically evaluated so that the staging of the refuse facility can be modifed, if needed. In
general, downstream construction results in more predictable embankment performance and may
require substantially less testing and engineering to address seismic loading, but this type of staging
is more sensitive to changes in the coarse to fne ratio.
5.1.3 Slurry Cell Staging
If slurry cells are designed and operated as small ponds with size and hazard potential classifcation
that do not require impoundment plans under 30 CFR 77.216, or are arranged such that a low-
hazard-potential impoundment classifcation results, their staging and construction may be more
complex than construction of an impounding refuse facility. A slurry cell facility typically has surface
diversion to prevent runof from collecting in the cells. Thus, the design storm plays a signifcantly
smaller role in slurry cell staging than would be the case with a signifcant- or high-hazard-poten-
tial slurry impoundment. Additionally, when the slurry cells are covered sequentially with layers
of coarse refuse, the fne refuse may be densifed, but this efect typically does not greatly afect the
capacity of a slurry cell facility.
Figure 5.2 illustrates the components of a slurry cell facility. The small size of individual slurry cells
necessitates operation of multiple cells, and in some instances, continuous cell construction, put-
ting greater operational demands on the disposal facility. Thus, staging can be critical with a slurry
< PREVIOUS VIEW
5-4
Chapter 5
MAY 2009
cell facility because the need to divert runof around the cells and to periodically cover the cells can
require as much or more coarse refuse than with slurry impoundment staging. Thus, the consistency
of the refuse generation rate is critical to successful operation of slurry cells. Slurry cells are an atrac-
tive option only when there are limitations to developing an impoundment at a disposal site, or when
the ratio of coarse refuse to fne refuse is quite large. Slurry cells are sometimes used as a contingency
alternative when underground injection is employed.
When slurry cells are operated only to dewater fne refuse slurry, operation and staging difculties
are signifcantly reduced. For this type of operation, slurry is placed in cells for dewatering and sub-
sequent removal and mixing with coarse refuse prior to fnal disposal. Typically, these slurry cells
are utilized for refuse processing plants with lower production rates that are operated over a longer
period of time, thus reducing the burden of continually constructing storm water diversion struc-
tures for new cells.
FIGURE 5.1 TYPICAL SLURRY IMPOUNDMENT COMPONENTS
FIGURE 5.1 TYPICAL SLURRY IMPOUNDMENT COMPONENTS
5.1a PLAN
7
0
0
675
675
700
7
2
5
7
7
5
7
5
0
8
0
0
825
850
875
900
8
5
0
8
7
5
9
0
0
925
9
2
5
9
0
0
8
7
5
8
5
0
8
2
5
8
0
0
750
7
7
5
8
0
0
8
2
5
8
5
0
8
7
5
9
0
0
EMBANKMENT
CREST
EMERGENCY
SPILLWAY
SADDLEBACK
DAM CREST
DOWNSTREAM
STAGE
SEDIMENTATION
POND
IMPOUNDMENT POOL
PERMANENT (GROIN)
DRAINAGE DITCH
UPSTREAM STAGE
SETTLED FINE
REFUSE DELTA
DECANT PIPE INLETS
DECANT PIPE
OUTLET
5.1b LONGITUDINAL CROSS SECTION
STARTER
DAM
STAGE II-B
STAGE II-A
CREST
SEDIMENTATION
POND
INITIAL FINE REFUSE DEPOSITS
SETTLED FINE
REFUSE DELTA
STAGE I
PIEZOMETER
INTERNAL
DRAIN
UNDERDRAIN
IMPOUNDMENT
POOL
CUTOFF KEYWAY
NOTE: 1. ROADS AND BENCHES ARE NOT SHOWN.
2. SLOPES ARE EXAGGERATED FOR ILLUSTRATION PURPOSES.
825
800
725
750
775
650
650
FIGURE 5.1 TYPICAL SLURRY IMPOUNDMENT COMPONENTS
< PREVIOUS VIEW
5-5
Coal Refuse Disposal Facility Design Components
MAY 2009
5.1.4 Non-impounding Refuse Embankment Staging
Staging for non-impounding embankments involves surface runof diversion, drainage provisions for
signifcant springs or groundwater fow, erosion and sedimentation control, construction sequencing
(upstream or downstream construction), and reclamation. When dewatered fne refuse is disposed
within upstream zones of a non-impounding embankment, staging is a function of: (1) the required
quantity of coarse refuse and the rate of construction of the well-compacted downstream shell (some-
times referred to as the structural zone), (2) haul road requirements, and (3) surface runof diversion
requirements. The construction sequence and reclamation planning will also afect the staging.
Figure 5.3 illustrates the components of a coarse refuse disposal facility. Surface runof diversion and
erosion and sedimentation controls infuence staging. Because the working surface and some por-
FIGURE 5.2 TYPICAL SLURRY CELL FACILITY (IMPOUNDING EMBANKMENT)
5.2b LONGITUDINAL CROSS SECTION UPON COMPLETION
FIGURE 5.2 TYPICAL SLURRY CELL FACILITY (NON-IMPOUNDING EMBANKMENT)
5.2a PLAN
REFUSE
CONVEYOR
1500
1450
1300
1350
1400
1550
1500
1250
1200
1200
1250
1350
1300
1400
1500
1550
1600
1650
1550
1600
1600
1700
1750
1650
1450
CELL DRAINAGE
SYSTEM
TEMPORARY
PERIMETER DITCH
PERMANENT (GROIN)
DRAINAGE DITCH
SEDIMENTATION
POND
STRUCTURAL
ZONE
INTERNAL
DRAIN
UNDERDRAIN
EXTENSION
FUTURE CELLS
UNDER CONSTRUCTION
ACTIVE CELL
COVERED
CELL
BENCH
STARTER EMBANKMENT
CREST
FINE COAL
REFUSE CELL
UNDERDRAIN
INTERNAL DRAIN (TYP.)
NOTE: ROADS AND BENCHES NOT SHOWN
ON CROSS SECTION.
FIGURE 5.2 TYPICAL SLURRY CELL FACILITY (IMPOUNDING EMBANKMENT)
< PREVIOUS VIEW
5-6
Chapter 5
MAY 2009
tions of the embankment slopes will be exposed refuse, surface runof from the embankment must
be controlled separately from runof from non-disturbed areas and must be directed to erosion and
sedimentation control structures. This requires that fows from the contributing natural watershed be
diverted around the disposal facility, which may require ditches or channels constructed upstream
from the embankment or integrated into the embankment perimeter.
FIGURE 5.3 TYPICAL REFUSE EMBANKMENT COMPONENTS
5.3a PLAN
5.3b LONGITUDINAL CROSS SECTION
FIGURE 5.3 TYPICAL REFUSE EMBANKMENT COMPONENTS
1800
1675
1650
1625
1575
1800
1775
1750
1725
1700
1675
1775
1500
1750
1725
1
5
2
5
1700
1550
1600
SEDIMENTATION
POND
TEMPORARY
PERIMETER DITCH
PERMANENT (GROIN)
DRAINAGE DITCH
TEMPORARY
PERIMETER DITCH
B
E
N
C
H
B
E
N
C
H
PERMANENT (GROIN)
DRAINAGE DITCH
SEDIMENTATION
POND
UNDERDRAIN
NOTE: 1. ACCESS ROAD NOT SHOWN IN CROSS-SECTION VIEW.
2. SLOPES ARE EXAGGERATED FOR ILLUSTRATION PURPOSES.
FIGURE 5.3 TYPICAL REFUSE EMBANKMENT COMPONENTS
< PREVIOUS VIEW
5-7
Coal Refuse Disposal Facility Design Components
MAY 2009
5.1.5 Amendments and Co-disposal of Combustion Waste
The use of amendments applied to coal refuse and co-disposal of combustion waste and coal refuse
will infuence staging dependent upon the quantity of the materials disposed. Other aspects of
embankment design may be afected because the engineering properties of the amended/combined
materials may be diferent from those of the coal refuse.
5.1.5.1 Amendments for Neutralization and Stabilization of Refuse
Amendments applied to coal refuse have included a variety of materials typically used to neutralize
the potential acidity of coal refuse or to absorb moisture for stabilization of the refuse during place-
ment and compaction. Amendments have included lime and lime products, kiln dust, and combus-
tion waste.
The inclusion of amendments for neutralization or stabilization of refuse may also have the following
infuences on design of the disposal facility:
Neutralizing agents if applied in quantity can afect the strengths of the fll (either
increase or decrease), and may alter handling and placement equipment require-
ments and embankment slopes. The thoroughness of mixing the amendments with
the refuse is important to their application as neutralizing or stabilizing agents and
afects the properties of the combined material.
Runof from neutralizing agents can have a high pH and thus may need to be
retained within treatment ponds prior to discharge into a receiving water body.
The hydraulic conductivity of the amended embankment materials may be lower
than the refuse materials, necessitating additional internal drainage measures and
resulting in precipitates that clog flters and internal drainage structures unless they
are appropriately designed.
Final reclamation of completed surfaces of amended embankment materials may be
facilitated, allowing a thinner soil and topsoil cap.
To assess the behavior of mixed refuse and amendments, physical properties, including strength and
hydraulic conductivity should be evaluated.
5.1.5.2 Co-disposal of Combustion Waste
Several types of combustion waste have been co-disposed with coal refuse at disposal sites, includ-
ing: fy ash and botom ash, fuidized bed combustion (FBC) ash, and fue gas desulfurization (FGD)
sludge (dewatered). Depending on the characteristics of the waste, it has typically been disposed
with coal refuse as: (1) part of contracting opportunities, (2) for structural or flter zones, (3) for sta-
bilization of refuse due to wet conditions, (4) for neutralization of potential acidity of refuse, and (5)
for reclamation improvements.
Co-disposal of combustion waste may have the following efects on the design of a disposal facility:
The extent of mixing of combustion waste with coal refuse will afect the homogeneity
of the embankment, particularly with respect to strength and hydraulic conductivity
because of the infuence of the combustion waste on these characteristics. The presence
of mixed and unmixed layers may result in stratifed conditions within the embankment.
Combustion waste if incorporated within a refuse embankment in quantity may
result in diferent (higher or lower) strength characteristics of the fll, requiring dif-
ferent handling and placement equipment and modifed embankment slopes. FBC
ash, if pozzolanic, will exhibit high strength at low strain similar to soil cement.
< PREVIOUS VIEW
5-8
Chapter 5
MAY 2009
The hydraulic conductivity of the embankment materials mixed with combustion
waste may be diferent (typically lower, but possibly higher) than the refuse materi-
als alone. This may result in precipitates that clog flters and internal drainage struc-
tures necessitating additional internal drainage measures. Combustion waste may
also increase the anisotropy of embankment materials.
FBC may have expansive characteristics that can afect instrumentation per-
formance and monitoring, infuencing the position and elevation of survey
monuments and the verticality and integrity of casings for piezometers and
inclinometers.
Use of combustion waste may increase dust control requirements at the embankment
surface.
Runof from combustion waste can have high pH and thus may need to be retained
within treatment ponds prior to discharge to the receiving water body.
Final reclamation of completed surfaces of embankment materials with co-disposed
combustion waste may be facilitated, allowing a thinner soil and topsoil cap.
Combustion waste may contain metals that can result in leachate and groundwater
impacts.
Research on the benefcial use and co-disposal of combustion waste with coarse coal refuse is reported
in Daniels et al. (2002). Research on co-disposal of combustion waste with fne coal refuse is presented
in Kumar et al. (2001).
5.2 EROSION AND SEDIMENTATION CONTROL AND DESIGN STORM RUNOFF
MANAGEMENT
Erosion and sedimentation control features are required for all disposal facilities, and their
design must be integrated into the overall drainage control plan. Management of runof from
design storms is a function of the type of refuse disposal facility and associated design storm.
Impounding facilities are typically designed to retain a portion of design storm fows within the
impoundment, thus atenuating the peak fow. Non-impounding embankments utilize diver-
sion channels and ditches that are designed to handle the peak discharge. Embankments for
both impounding and non-impounding facilities typically have associated drainage channels
and ditches that are designed to handle storm fows. For larger refuse facilities, erosion and
sedimentation controls can represent a signifcant component in the overall design of the facil-
ity and can occupy a relatively large portion of the site that should be accounted for during the
initial phases of design.
5.2.1 Erosion and Sedimentation Control
Erosion and sedimentation control requirements are generally based on state regulatory criteria.
Their function is to divert runof from undisturbed areas of a site and to collect and treat runof
from disturbed areas. Diversion systems are normally required for all refuse embankments; how-
ever, impoundments can serve to provide sediment control for their contributing watershed. The
design of the erosion and sedimentation control system is based upon the disturbed area at the
disposal facility during operation and as reclamation is performed.
Sediment control is typically provided by ponds that receive drainage from the embankment and
disturbed areas, and these are typically located near the downstream toe of the disposal facility.
As discussed in Chapter 9, the ponds are designed with decant systems and inlet structures that
provide adequate capacity and setling time for removal of sediment and control of storm events
without activating the emergency spillway. The design of the emergency spillway is based on the
ponds hazard-potential classifcation and applicable state criteria.
< PREVIOUS VIEW
5-9
Coal Refuse Disposal Facility Design Components
MAY 2009
5.2.2 Slurry Impoundment Infow Design Storm
Selection of the appropriate design storm (short- and long-term) is a function of the hazard poten-
tial of the planned slurry impoundment. Based on the design storm and site conditions, the runof
rate and volume can be determined. The facility can then be confgured to manage the design storm
through a combination of storage and routing. The runof volume is important because frequently
the size and confguration of the outlet system (decant pipe or spillway) for impoundments are
determined by the volume of water that must be evacuated from the impoundment within a certain
specifed time period afer the storm event (Chapter 9). Similarly, the staging of the facility must be
adequate to retain the runof from the design storm with adequate freeboard.
High-hazard-potential facilities must be able to handle the most severe design storm, while a lesser
magnitude storm will be suitable for a signifcant- or low-hazard-potential classifcation facility. The
determination of the hazard-potential classifcation is discussed in Section 3.1. Dam breach analyses
may be performed as part of hazard-potential determination, as discussed in Chapter 9.
Short- and long-term embankment design criteria must be evaluated in the determination of the
design storm, typically to address conditions during the frst two years of construction and operation,
as the impoundment is being developed, and during the last two years of the facilitys life, when it is
being reclaimed. Short- and long-term design criteria are discussed in Chapter 9, as part of the hydro-
logic and hydraulic engineering analyses, and are primarily a function of the hazard potential and
the size of the planned impoundment. For example, during the initial year of construction and opera-
tion of a large, high-hazard-potential impoundment, the capacity to handle the 100-year storm may
be needed; during the second year the Probable Maximum Flood ( PMF) event; and within two
years the full PMF event. Short-term criteria are intended for unavoidable construction conditions
during initial start-up and abandonment and should only be utilized when absolutely necessary.
Impoundments are typically designed to: (1) store the runof from the design storm with slow release
of the runof through a decant pipe system, (2) release the design storm runof without storage
through an open-channel spillway, or (3) store a portion of the design storm runof and release the
balance through an open-channel spillway. Figure 5.4 illustrates the concepts of 100-percent stor-
age of the design storm runof and the use of an open-channel spillway to route the design storm
runof. An impoundment that is designed to store the design storm runof with release through a
decant (Figure 5.4a) requires greater surcharge storage capacity than an impoundment that employs
an open-channel spillway (Figure 5.4b). The determination of the magnitude of the design storm,
including the short- and long-term criteria upon which the design storm is based and associated
runof rates and volumes, is discussed in Chapter 9.
Drainage channels at other locations at impoundment facilities, such as the downstream embank-
ment face, are typically designed for the 100-year storm, and these structures do not afect the overall
confguration, geometry and staging of embankment construction. However, if impoundment decant
systems discharge to embankment drainage channels, the discharge should be added to the surface
fow collected by the drainage channels. If the drainage channels intercept a substantial amount of
runof, the design storm criteria for the impoundment may need to be applied to the channel design
in order to protect the embankment from erosion during a severe food.
5.2.3 Slurry Cell Facilities
Slurry cell systems are atractive in situations where they can be kept under the size and hazard
thresholds of 30 CFR 77.216, or where they can be classifed as having low hazard potential,
so that design storm management can be based on the 100-year event and not the PMF. A low-
hazard-potential determination requires information to substantiate that, in the event of a failure,
the facility will not release water, sediment, or fowable fne refuse that would cause loss of life
or signifcant property damage, as discussed in Chapter 3. Thus, the sequence of active and open
< PREVIOUS VIEW
5-10
Chapter 5
MAY 2009
cell operation must be designed to minimize impounding capacity. Also, the closed and covered
slurry cells must be oriented and dewatered such that they do not collectively represent a potential
release of fowable material that would severely impact downstream development.
By achieving a low-hazard-potential classifcation or avoiding classifcation as a regulated impound-
ment, slurry cell systems can be designed for lesser design storm criteria (i.e., 100-year storm). Because
the watershed area contributing to the slurry cells is minimal, only a small quantity of runof has to be
managed as part of the slurry disposal operation. While this reduces the infuence of the design storm
runof on the overall confguration and staging of the slurry cell system, runof from adjacent areas, par-
ticularly upstream watershed areas, must be diverted, which requires construction and maintenance of
diversion channels that are incorporated into the embankment or disposal site confguration.
FIGURE 5.4 ILLUSTRATION OF IMPOUNDMENT DESIGN STORM CONTROL
Drainage structures for slurry cells must have the capacity to control runof from the design storm
collecting within the cell, along with clarifed water produced as the slurry is deposited. This drain-
age may be achieved by using a decant pipe with an appropriately sized inlet or riser to facilitate
setling of the fne refuse. An open-channel spillway may also be needed depending on the slurry cell
arrangement and contributing drainage area.
DECANT
INLET
DESIGN STORM FREEBOARD
NORMAL POOL
ELEVATION
NORMAL
FREEBOARD
SURCHARGE STORAGE
SETTLED FINE
REFUSE
EARTH OR
COARSE REFUSE
MINIMUM POOL LEVEL
CONTROLLED BY PUMPING
REFUSE
EMBANKMENT
FIGURE 5.4 ILLUSTRATION OF IMPOUNDMENT DESIGN STORM CONTROL
5.4a IMPOUNDMENT DESIGNED FOR 100-PERCENT STORAGE OF DESIGN STORM
DECANT PIPE
(DIAPHRAGM NOT SHOWN)
5.4b IMPOUNDMENT DESIGNED WITH OPEN-CHANNEL SPILLWAY
DECANT
INLET
DESIGN STORM FREEBOARD
SURCHARGE STORAGE
EARTH OR
COARSE REFUSE
MINIMUM POOL LEVEL
CONTROLLED BY PUMPING
REFUSE
EMBANKMENT
PROJECTED INVERT OF
OPEN-CHANNEL SPILLWAY
DECANT PIPE
(DIAPHRAGM NOT SHOWN)
NOTE: SLOPES ARE EXAGGERATED FOR ILLUSTRATION PURPOSES.
SETTLED FINE
REFUSE
FIGURE 5.4 ILLUSTRATION OF IMPOUNDMENT DESIGN STORM CONTROL
< PREVIOUS VIEW
5-11
Coal Refuse Disposal Facility Design Components
MAY 2009
As with other site drainage not related to impoundments, the channels and ditches on the down-
stream face of a slurry cell embankment must be designed for the 100-year storm and do not afect the
overall confguration or staging. If a slurry cell facility is designed as a signifcant- or high-hazard-
potential impoundment, the associated design storm criteria may need to be used also for the design
of embankment drainage channels that receive decant discharge.
5.2.4 Refuse Embankment Design Storm
The design storm for non-impounding refuse embankments is primarily used for sizing channels and
ditches that divert or collect and convey drainage from the embankment, as part of the erosion and
sedimentation control for the site. Diversion ditches are typically constructed around the perimeter
of refuse embankments so that channel runof from undisturbed portions of the site is conducted
around the disturbed areas to sedimentation ponds. If a diversion ditch must be relocated to a higher
elevation as an embankment is expanded, a plan of ditch relocation that provides for drainage diver-
sion from disturbed areas during the associated construction must be prepared.
Permanent channels and diversion ditches should be designed for the peak discharge from the 100-
year storm. Temporary channels may be designed for a lesser magnitude event, as discussed in
Chapter 9. Long-term drainage control required during reclamation and abandonment should be
incorporated into the drainage channel design.
5.3 FACILITY CONFIGURATION AND GEOMETRY
The confguration and geometry of a refuse disposal facility is a function of staging/construction
methods, capacity requirements, engineering properties of the refuse, initial site topography, storm
water management, regulatory criteria and overall facility stability requirements. Operational fac-
tors, such as refuse transport and handling equipment also infuence the design and overall confgu-
ration of the facility.
5.3.1 Regulatory Criteria
Criteria have been established by federal (Ofce of Surface Mining) and state agencies that address
some parameters for design of coal refuse disposal facilities. These parameters include embankment
slope, which should not be steeper than 2:1 (horizontal to vertical) between benches. Benches are
used to control erosion, retain soil moisture, or facilitate reclamation and are typically integrated into
embankment slopes.
For slurry impoundments that are classifed as dams by state agencies, regulatory criteria for the
width of the crest and maximum slopes should be reviewed. Other regulatory criteria may afect the
design confguration and geometry, as a result of required right-of-ways, bufer zones, and related
of-sets from natural features, structures, utility lines and property lines. State and federal criteria
may change with time.
Federal and state regulatory agencies require engineering analyses for demonstrating that certain
design criteria are satisfed, such as the ability of the facility to safely route the design storm or to pro-
vide a specifed factor of safety relative to stability. State criteria are sometimes diferent from federal
criteria. Both should be considered, and the most stringent criteria should be applied.
5.3.2 Embankment Design Considerations
The following subsections provide a discussion of issues related to embankment design.
5.3.2.1 Stability under Normal Operating Conditions
Stability analyses are performed to determine the embankment confguration and slope. The param-
eters associated with these analyses include the following:
< PREVIOUS VIEW
5-12
Chapter 5
MAY 2009
Embankment material strengths These strengths are dependent upon the types of
refuse, borrow materials, and amended/co-disposed materials used in embankment
construction, as well as the method of placement and compaction. Sufcient explora-
tion and testing must be performed such that all applicable material properties can
be accurately characterized.
Foundation material strengths (as determined by exploration and testing) These
may be improved if necessary by removal and replacement or by ground improve-
ment measures. Sufcient exploration and testing must be performed such that in-
situ material properties can be adequately characterized.
Seepage and phreatic level These are a function of normal pool level, internal
drainage systems and liner systems.
Non-impounding coal refuse disposal facilities are normally developed over several years, and, as
with impounding facilities, the confguration may vary considerably during various stages of the
facilitys life due to upstream and downstream development. Stability analyses should be conducted
at various points or critical stages in the embankment development. A detailed evaluation of the
facility staging and construction should be performed to identify critical confgurations that require
stability analysis. Chapter 6 presents guidance related to stability analyses and other geotechnical
engineering issues.
5.3.2.2 Stability under Extreme Events
Embankment stability under design storm and earthquake loading conditions (assumed to not occur
simultaneously) must be evaluated. For design storm conditions, the embankment stability and con-
fguration may be afected by the following:
Flood level (peak pool level resulting from the design storm event) at an impound-
ment, which may afect the embankment phreatic surface and seepage conditions,
depending on the depth and duration of the elevated reservoir level and the efec-
tiveness of internal drainage measures.
The location and elevation of decant and spillway structures.
Seismic analyses are performed on impounding embankments and sensitive embankments such
as slurry cell facilities. Guidance for these analyses is presented in Chapter 7 and includes evalu-
ation of strength loss or degradation of embankment or foundation materials as a result of seis-
mic loading and evaluation of seismic deformations that may affect embankment performance.
Design features that may be considered for mitigating the adverse effects of earthquake loading
and associated embankment response include providing extra freeboard, enhancing internal
drainage and related consolidation of the fine refuse, widening the embankment, and adding a
buttress.
5.3.2.3 Settlement and Subsidence
Setlement of foundation soils or refuse materials, particularly if upstream construction is employed
for an impounding embankment, may require placement of additional material to maintain crest
elevations and grades of drainage structures. Large diferential setlements may represent a risk
of cracking of strain-sensitive materials (e.g., clay zones) or barriers, thus requiring measures to
limit diferential setlement. Chapter 6 provides guidance for setlement and other geotechnical
analyses.
Subsidence associated with underground mines may afect the design confguration of the embank-
ment and is discussed further in Section 5.4.
< PREVIOUS VIEW
5-13
Coal Refuse Disposal Facility Design Components
MAY 2009
5.3.2.4 Environmental Issues, Impoundment Elimination and Site Reclamation
Amendments to refuse or the presence of a liner system beneath an embankment or zoned embank-
ment (for an impoundment facility) can afect the stability of the embankment, requiring modifca-
tions to confguration or slope. Reclamation, post-mining land use, and abandonment practices can
also afect the embankment confguration and slope.
5.3.3 Impoundment Design Considerations
The impoundment confguration can also be afected by a range of design considerations, as described
in the following paragraphs.
5.3.3.1 Settling and Clarifcation
Facilities such as polishing ponds, bafes, clear water cells, and similar structures may be required in
order to produce water quality that is acceptable for discharge or preparation plant use. Clarifcation
may also be accomplished in ponds downstream of the impoundment facility.
5.3.3.2 Seepage Control and Containment
To mitigate environmental impacts, seepage control and liner systems may be necessary for contain-
ment of the fne coal refuse.
5.3.3.3 Spillway and Decant Systems
Decant systems ofen serve as primary spillways and, if the watershed and impoundment sizes are
compatible, may be sufcient to manage the design storm without an open-channel spillway. This
will require that signifcant freeboard be maintained under normal operating conditions. In some
instances, including during the fnal years of operation of a facility, open-channel spillways sized
to handle the design storm are constructed, thus allowing more of the impoundment to be used for
refuse disposal.
5.3.3.4 Subsidence and Breakthrough Potential
Subsidence analyses may identify potential impacts to an impoundment area, requiring changes in
layout or other design features that afect the site confguration. Where mine workings are in prox-
imity to an impoundment, evaluation of breakthrough potential may indicate that measures such as
constructed barriers (embankment flls) are needed to bufer the zone of impact and maintain inter-
nal stability. These and other measures are addressed in Chapter 8.
5.4 FOUNDATION AND MINE SUBSIDENCE CONSIDERATIONS
5.4.1 Foundation Materials
Foundation materials must be evaluated relative to their strength, compressibility and hydraulic
conductivity. Foundation materials without adequate strength, or with high compressibility, may not
provide adequate support for the disposal facility, leading to detrimental movement or actual failure.
Hydraulic conductivity is an issue primarily with impounding embankments, where high-hydraulic-
conductivity foundation materials can lead to elevated pore pressures afecting local stability and
creating the potential for internal erosion or piping of fne soil particles that can threaten the overall
stability of the disposal facility. High-hydraulic-conductivity foundation materials can also lead to
unacceptable loss of clarifcation water or, in extreme situations, release of impounded materials.
Strength and compressibility are typically concerns for foundation soils, while hydraulic con-
ductivity may be a concern for both soil and bedrock conditions. Bedrock strength conditions can
be an issue where weak layers, such as claystones, underlie an embankment and behave similarly
to a soil (primarily where shallow or surface outcrop conditions occur). Weak or fractured bed-
< PREVIOUS VIEW
5-14
Chapter 5
MAY 2009
rock conditions may also become an issue where underground mine workings are in proximity
to an embankment or impoundment such that subsidence could cause diferential movement and
impose strain on structural components or drainage features of the disposal facility. An addi-
tional concern, as discussed in Section 5.4.3, can be actual breakthrough of the disposal facility
into mine workings.
A variety of measures to address foundation material concerns are available. These may include
accommodations in the disposal embankment and impoundment design or in-situ foundation
treatments ranging from removal and replacement of shallow, weak soils to grouting of fractured
bedrock conditions. Seepage control within foundation materials typically is accomplished by con-
structing cutof trenches through high-hydraulic-conductivity natural soils and weathered bed-
rock and backflling with low-hydraulic-conductivity soils. Cutof trenches may disclose geologic
features such as alluvial deposits, sand dikes, and relief fractures that require treatment and that
would not be revealed by exploration borings.
5.4.2 Abutment and Foundation Geometry
The geometry of the abutments and foundation (steep slopes and bedrock exposures) of an impound-
ing embankment can lead to diferential setlements resulting in cracking of embankment materials,
disruption of drainage control features, and seepage concerns. This situation may be exacerbated at
former surface mine sites where embankment construction is planned in areas of former highwalls.
For non-impounding embankments, such concerns may also need to be addressed in the design of
drainage control systems.
Measures to address steep foundation geometry may include benching and cuting back of slopes,
zoning with select structural materials, and planned compensation flls in the embankment design to
ofset cumulative setlements. These issues are addressed in Chapter 8.
5.4.3 Mine Subsidence Potential
The presence of underground mines or the potential for future underground mining in the vicinity of
a coal refuse disposal facility may infuence the location and design of facility features. Additionally,
auger or highwall mining can lead to subsidence and seepage concerns. Location of an impoundment
facility in the vicinity of mine workings is of concern because of the potential impacts of subsidence
on embankment stability and the related potential for release of impounded materials, as well as the
risk of a breakthrough into the mine, which is discussed in Section 5.4.4. Non-impounding facility
design may also be infuenced by potential mine subsidence, particularly related to impacts on struc-
tural elements and drainage and liner systems.
Depending on the depth to mine workings and proximity to the disposal facility, subsidence analy-
ses may be needed for determination of the potential magnitude, tilt, curvature, and strain associ-
ated with ground movement related to subsidence. These analyses provide a basis for evaluation
of the subsidence impact on the disposal facility and can help to identify potential mitigation mea-
sures. The engineering methodology for performing these analyses is presented in Chapter 8, and
the potential impact of diferential movement on coal refuse materials and structures is discussed
in Chapters 6 and 8.
Similar analyses should be performed for proposed mining in the vicinity of disposal facilities.
Only under very favorable conditions and where development is essential for haulage or ventilation
should mining be performed under impounding embankments. The extent of mining that may be
conducted under an impoundment is dependent upon the overburden and mining conditions. It is
current practice that subsidence analyses are performed for longwall mines in order to establish a
safety zone where extraction is precluded for protection of an impounding embankment.
< PREVIOUS VIEW
5-15
Coal Refuse Disposal Facility Design Components
MAY 2009
Subsidence impacts associated with drainage control systems (liners, internal drains, and surface
drainage systems) may be addressed through evaluation of potential diferential movement and
strain during design. When mine workings are present beneath or in the vicinity of an impound-
ment, the following design measures may be employed to mitigate potential subsidence efects on
embankments:
Maintenance of ample freeboard to compensate for the anticipated subsidence.
Broad embankment confguration design that is less sensitive to subsidence impact
or enhanced zoning of embankments with self-healing characteristics (e.g., cohesion-
less sand) to mitigate potential cracking and internal erosion.
Allowance for potential subsidence efects in the design of internal drainage struc-
tures or implementation of enhanced seepage control measures (barriers and chim-
ney drains) to minimize the efects of potential subsidence.
Backflling or grouting of mine entries in critical support areas to minimize the
amount of movement that can occur.
Installation of monitoring systems for detecting subsidence and associated impacts
such as increased seepage and preparation of contingency plans for mitigating such
impacts.
To address the potential for subsidence and seepage in areas of auger or highwall mining, backflling
and/or protection berms or embankments may be employed, as discussed in Chapter 8.
5.4.4 Impoundment Breakthrough Potential into Mine Workings
In addition to subsidence concerns, if the overburden separating an impoundment from existing
mine workings is relatively thin, there is a potential for sinkholes or even catastrophic breakthrough
of the impoundment into the mine workings. In general, areas where the cover over a mine entry
comprises less than 100 feet of intact rock strata are a concern for sinkhole development. Even greater
amounts of cover are cause for concern in some situations. In addition to possible sinkhole devel-
opment, shearing or punching of a coal outcrop is another potential failure mechanism that must
be prevented. Without proper design, failure can occur through the coal seam itself, through strata
above the coal, or along material interfaces or discontinuities. Internal erosion along these interfaces
and discontinuities must also be prevented.
The presence of auger or highwall mining workings adjacent to an impoundment can also be cause
for concern. The potential for subsidence should be evaluated in design, and measures to control
seepage and to address the potential for breakthrough into neighboring underground mine workings
should be developed, as needed. Figure 5.5 illustrates the presence of mine workings immediately
adjacent to and beneath an impoundment and the use of a berm or barrier and internal drainage
system to mitigate potential seepage and breakthrough potential.
Chapter 8 presents guidance for analyzing breakthrough potential and preventing breakthroughs.
Some basic design measures that may be useful for mitigation of the potential for breakthroughs
include:
Development of a safety zone (based on verifable data related to mining and out-
crop or overburden conditions) that provides sufcient undisturbed ground between
the mine and the impoundment so that the ground movement induced by mining
will not afect the impoundment.
Development of measures to provide adequate support for the impoundment con-
sidering potential failure mechanisms such as some combination of: (1) an engi-
< PREVIOUS VIEW
5-16
Chapter 5
MAY 2009
neered barrier (soil or synthetic barrier) or berm (coarse refuse) to mitigate the
failure mechanisms, (2) an internal drainage system to control seepage and to reduce
pressures in the areas of potential breakthrough, or (3) stabilization of the fne coal
refuse to increase strength and reduce fowability. Any mitigation measure employed
must also be resistant to subsidence.
Development of containment or diversion structures, such as bulkheads, within the
afected mine workings.
Stabilization of the mine workings that may afect the impoundment by mine flling
and/or grouting to improve strength and resistance to failure mechanisms.
As an alternative to the berm illustrated in Figure 5.5, slurry cell confgurations have been employed
at some impoundment sites with breakthrough potential, or the operation has been converted to fne
coal refuse paste in order to restrict the potential release of fowable material.
5.5 UNDERGROUND MINING AT SLURRY IMPOUNDMENT
HIGHWALL
PERIMETER BERM
OR BARRIER
COAL SEAM
DRAIN AND
FILTER SYSTEM
AUGER OR
MINE WORKINGS
MINE WORKINGS
IMPOUNDMENT
POOL
SETTLED
FINE REFUSE
NOTE: SLOPES ARE EXAGGERATED FOR ILLUSTRATION PURPOSES.
FIGURE 5.5 UNDERGROUND MINING AT SLURRY IMPOUNDMENT
5.4.5 Mine Entry Barriers and Bulkheads
Where mining is occurring in the vicinity of a coal refuse disposal facility, barriers and bulkheads
may need to be constructed for the purpose of controlling drainage and mitigating the potential
for internal erosion and piping in the embankment. Mine entries may include mine openings,
ventilation boreholes, auger holes, or highwall mining areas. The design requirements for barri-
ers and bulkheads are dependent on the type of disposal facility, location of the entry, and pres-
ence of drainage entering into or discharging from the mine. For support of non-impounding
embankments, only a barrier may be required, particularly if drainage conditions are not an
issue. Entries beneath impounding embankments or beneath impounded water will probably
require installation of a bulkhead and/or barrier along with additional measures for addressing
drainage at the entry location or within the mine. As an additional line of defense, bulkheads
should be considered at critical discharge locations. Chapter 8 provides guidance for the design
of barriers and bulkheads.
5.5 INTERNAL DRAINAGE AND EMBANKMENT SEEPAGE CONTROL
Internal drainage and seepage control features are designed to accomplish: (1) collection of springs
and hillside subsurface drainage and conveyance to a point downstream of the refuse embankment
(and sedimentation ponds), (2) environmental containment using liner systems to control seepage to
subsurface groundwater beneath the site, (3) collection of seepage and conveyance downstream to
FIGURE 5.5 UNDERGROUND MINING AT SLURRY IMPOUNDMENT
< PREVIOUS VIEW
5-17
Coal Refuse Disposal Facility Design Components
MAY 2009
treatment or sedimentation ponds, (4) control of the phreatic surface or foundation pore pressures
to maintain embankment stability, and (5) prevention of seepage along penetrations (decants, etc.)
through the embankment. Reclamation may also involve construction of a low-hydraulic-conductivity
cap to limit infltration and reduce internal drainage requirements.
Technical guidance for internal drainage and embankment seepage control is provided in Chapter 6
and covers the following:
Seepage analyses for embankments (homogeneous and zoned), including analysis of
the efects of foundation cutofs, liners, and internal drains.
Filter criteria for zoned embankments.
Design of internal drains, including flters.
The environmental aspects of reclamation, internal drainage and seepage control are addressed in
Chapter 10.
5.6 DECANT AND EMERGENCY SPILLWAY SYSTEMS
Hydraulic structures for controlling impoundment levels and providing storm routing are designed
based on design storm runof management, as discussed in Section 5.2, with specifc features infu-
enced by the type of embankment staging employed (upstream versus downstream) and confgura-
tion (height and length of the disposal facility), site terrain, and foundation issues.
5.6.1 Decant Performance Considerations
Decant systems provide for discharge of clarifed water from an impoundment. While a decant system
may be a component of the design-storm water management system, in many cases the impound-
ment level may actually be controlled by pumping when the water surface level is below the decant
invert. In cases where a decant system is part of the design storm management system, it is typically
designed to provide sufcient capacity to evacuate at least 90 percent of the design storm runof
retained within the impoundment within 10 days, assuming that no design-storm runof storage is
available below the lowest ungated decant inlet. The 10-day period is measured from the time of peak
impoundment pool during the design storm.
The design of decant systems involves the following considerations:
Hydraulic capacity evaluation (head versus discharge analysis).
Material selection, as afected by the length and depth of cover, foundation condi-
tions, installation method, and other factors such as strength, fexibility, durability,
corrosion protection, and joint type.
Inlet design features, including riser pipe structural stability and support, trash
guards, anti-vortex devices, and sealing requirements upon discontinuation of use.
Alignment and grade, including variation in grade due to terrain, potential foundation
setlement, the need for reaction blocks, resistance to fotation, and venting systems.
Outfall design, energy dissipation requirements, and erosion protection.
Pipe backfll, drainage diaphragms and seepage control.
Monitoring/testing requirements.
If decant systems are designed to discharge to a groin ditch adjacent to the downstream embank-
ment slope, additional design considerations may apply. In such instances, the peak decant discharge
should be added to the surface fow collected by the groin ditch. If the ditch intercepts a substantial
< PREVIOUS VIEW
5-18
Chapter 5
MAY 2009
amount of runof, the design storm criteria for the impoundment may need to be applied to the ditch
design in order to protect the embankment from erosion during a severe food.
5.6.2 Open-Channel Spillway Performance Considerations
Many impoundments are designed such that the reservoir storage capacity and decant system dis-
charge capacity provide for design storm runof management, but sometimes a separate open-chan-
nel spillway is also required. These should be designed with a capacity that will allow routing of the
design storm fow through the impoundment while maintaining adequate freeboard. Open-channel
spillways are less commonly employed at refuse disposal embankments because of staged develop-
ment. The following should be considered in the design of open-channel spillways:
Hydraulic capacity (head versus discharge capacity) and minimum freeboard for the
approach channel, control section, and discharge channel or conduit so that sufcient capac-
ity is available for releases up to the design peak outfow. Measures to control foating debris
or hillside trees that could cause obstructions may be required at the approach channel.
Channel geometry, including considerations for transition sections (e.g., contraction
section), changes in alignment (e.g., super-elevation at bends), and grade (e.g., suf-
fcient capacity for hydraulic jumps).
Stable channel conditions, considering excavation slopes as well as potential for ero-
sion due to water fow associated with velocity/tractive force/duration on the chan-
nel lining. Channel linings require suitable foundation and drainage systems and
should be designed for tractive and uplif forces.
Energy dissipation structures at the spillway outlet.
If large-diameter conduits are used in conjunction with open channels or in place of an open channel,
material selection, inlet design, alignment and grade, backfll, seepage control, and outlet design issues
are similar to those for decant systems.
5.7 SURFACE DRAINAGE CONTROLS
Surface drainage controls at coal refuse disposal embankments typically consist of drainage ditches
and channels, bench and haul road guters, and culverts that collect and convey runof to down-
stream structures.
5.7.1 Permanent Drainage Controls
Permanent drainage controls are structures that will be in service during operation of the disposal
facility and following reclamation and abandonment of the facility. These structures should be
designed for the 100-year-recurrence-interval storm. Typically, the 100-year, 24-hour-duration storm
is used for design, consistent with state regulatory criteria.
The design of permanent drainage controls should be based upon the following considerations:
Hydraulic capacity considering the peak discharge rate from the design storm for the
contributing drainage area.
Channel geometry, including accommodation for transition sections and changes
in alignment (e.g., super-elevation at bends) and grade (e.g., additional capacity for
hydraulic jumps in subcritical fow sections).
Stable channel conditions, considering fow velocity/tractive force/uplif/duration for
the channel lining, if the channel is not excavated in competent rock.
Energy dissipation structures at channel outfalls.
< PREVIOUS VIEW
5-19
Coal Refuse Disposal Facility Design Components
MAY 2009
5.7.2 Temporary Drainage Controls
Temporary drainage channels function periodically during operation, but are not part of the long-
term surface water controls related to reclamation and abandonment. These may include temporary
ditches for diversion or runof collection within the disposal facility, bench and haul road guters
(before topsoil placement and reclamation), and most culverts. Temporary ditches and culverts may
be designed with a hydraulic capacity lower than for permanent drainage controls, because of shorter
service life and other criteria. For instance, state regulations or recommendations for some drain-
age structures may specify a 10- or 25-year-recurrence-interval storm. Temporary ditches may be
designed without linings provided that: (1) their service life is relatively short and potential erosion
damage can be readily repaired and (2) lack of linings will not cause adverse impacts to the safety of
the disposal facility or compromise downstream sedimentation controls.
5.8 INSTRUMENTATION AND PERFORMANCE MONITORING
Instrumentation is typically a component of coal refuse disposal plans and is recommended for all
high or signifcant-hazard-potential sites. Instrumentation records should be reviewed as they are
accumulated. The data are best evaluated by maintaining a continuous plot of the readings versus
time. The data should be reviewed as part of annual inspections and certifcations and should be
maintained for the life of the facility. The data review provides a basis for: (1) assessment of facility
performance relative to design intentions, (2) detection of trends and problems that may develop,
and (3) operational plan modifcation or facility expansion. Design plans should indicate maximum
acceptable levels for instrument readings or percentage changes that trigger further investigation or
actions. Facility performance will typically be enhanced by monitoring the parameters discussed in
the following paragraphs with appropriate instrumentation:
5.8.1 Seepage
Seepage from the impoundment should be monitored for fow rate and changes in appearance
(discoloration or appearance of fne particulates and precipitates). This monitoring should include
seepage through the embankment, through internal drainage structures, and through underground
mines that receive seepage from the impoundment. Weirs should be installed, preferably with a
staf gauge in the weir approach pool, so that fow rates can be easily and accurately measured. To
evaluate changes in seepage rates, it is important to know the impoundment pool level and to have
data pertaining to rainfall and groundwater levels, so that possible correlations can be evaluated.
5.8.2 Piezometric Levels
Saturation levels and water pressures within an impounding embankment or embankment foun-
dation, as well as within any earthen barriers, should be monitored and recorded to determine
whether hydrostatic pressures are within design limits and whether changes or trends are reason-
able. Selection of the type of piezometer and installation location is based on site-specifc require-
ments and the potential for rapid or sudden changes in pore water pressure (as may occur upstream
of construction areas). Open standpipe piezometers provide direct measurement of groundwater
levels, while vibrating-wire and pneumatic piezometers allow for monitoring of phreatic conditions
in fne-grained deposits where rapid changes in pore pressure may be important. A table indicating
maximum allowable readings for piezometers should be included with the design report.
5.8.3 Pool Levels
Records for the pool level in the impoundment should be maintained for freeboard monitoring and
for determining correlations between piezometric levels and seepage quantities. Pool level can be
monitored by installation and reading of staf gauges. When water in nearby mine workings has the
potential to afect an impoundment or may indicate the performance of a barrier, the mine pool level
should be monitored.
< PREVIOUS VIEW
5-20
Chapter 5
MAY 2009
5.8.4 Rainfall Data
A rain gauge installed near the disposal site can provide site-specifc precipitation data for correla-
tions with piezometric levels and seepage quantities, and these data may be essential in situations
where there is breakthrough potential and where discharges from a mine are related to seepage from
an impoundment. Rainfall data should be routinely collected and recorded so that changes in seep-
age, mine discharge, or water level data can be correlated to rainfall infltration/runof.
5.8.5 Deformation or Movement
Where signifcant embankment setlement can occur or there is a potential for subsidence in the vicin-
ity of an impoundment, movements should be monitored and recorded. Monitoring of the slopes and
crest of an impounding embankment should be conducted at regular intervals during both operat-
ing and dormant phases for evaluating performance and for demonstrating conformance with food
routing assumptions. When subsidence is a concern, both horizontal and vertical movements should
be measured. Movements should also be monitored if evidence of slope displacement is detected. In
situations where deformation is occurring, the rate of movement, and especially any acceleration of
the rate, provides valuable information for assessing its signifcance. Methods for monitoring surface
deformation include survey monuments and extensometers. Movements below the ground surface
can be monitored with inclinometers and extensometers. These types of instruments are discussed in
more detail in Chapter 13.
5.9 RECLAMATION, ABANDONMENT AND POST-MINING LAND USE
A coal refuse disposal plan should address reclamation, abandonment and post-mining land use
requirements.
General provisions for and plans related to abandonment of a coal refuse embankment are part of
the fnal operational stage of the disposal facility and should address elimination of the impound-
ment, unless the impoundment is a component of planned post-mining land use. Reclamation should
incorporate the following provisions:
Stockpiles of soil and topsoil for reclamation should be located near the facility on
stable ground and within sedimentation controls.
Reclamation materials should meet the growth medium and nutrient requirements
of the vegetation plan. Topsoil amendments and alternatives may be necessary.
To protect against erosion and sedimentation and potentially negative environmen-
tal impacts, surface drainage and infltration should be controlled.
Elimination of impoundments should address the following:
Regrading of the impounding embankment and backflling of the impoundment
should be performed in a manner such that proper surface drainage is established
and the fne coal refuse is stabilized. Final backfll elevations should facilitate drain-
age and accommodate setlement of the fne refuse that will occur over time.
During the fnal periods of disposal and progressive elimination of the impound-
ment capacity, the outlet works such as the decant structure or spillway should
remain operational until impoundment regrading is complete.
Decant systems should generally be sealed by grouting.
Post-mining land use will afect the reclamation plan, particularly if existing structures such as the
impoundment or ponds are to be retained.
< PREVIOUS VIEW
Chapter 6
GEOTECHNICAL EXPLORATION,
MATERIAL TESTING, ENGINEERING
ANALYSIS AND DESIGN
6-1 MAY 2009
This chapter addresses important aspects of geotechnical exploration, material testing, engineering
analysis and design for coal refuse disposal embankments and impoundments with consideration
of past, current and future mining practices; characteristics of foundation, coal refuse, and soil and
rock borrow materials; and procedures for material placement and facility construction. Develop-
ment of a subsurface exploration program, implementation of a feld and laboratory testing pro-
gram, and selection of geotechnical parameters are key elements in the design of safe facilities for
coal refuse disposal. The basic design considerations that must be evaluated for both embankments
and foundations are seepage, slope stability and setlement. Another important geotechnical con-
sideration is the analysis of soil-structure interaction for buried pipes (conduits or decant pipes)
that are installed within an embankment. Specifcation, feld control and verifcation of geotechni-
cal properties are essential to the construction of coal refuse embankments that are consistent with
design assumptions.
From a geotechnical perspective, the following steps are normally followed in the design and con-
struction of a new or modifcation of an existing refuse embankment:
Review of available information
Site selection and optimization
Field exploration and in-situ sampling and testing
Laboratory testing
Development of geotechnical design parameters
Analyses and design
Preparation of plans and specifcations
Construction monitoring (quality control and verifcation of feld conditions)
Instrumentation
Embankment and system component performance monitoring
Maintenance
The level of efort and technical scrutiny required during the above steps will vary depending upon
the refuse facility intended use, size and hazard-potential classifcation.
< PREVIOUS VIEW
6-2
Chapter 6
MAY 2009
This chapter describes the scope of geotechnical investigations recommended for support of analyses
and design of a refuse disposal facility. References for additional information are provided herein.
Based upon the technical guidance provided in this chapter and supplemental information available
in the cited references, an experienced geotechnical engineer familiar with the refuse disposal process
and the design of water-retention embankment structures should be able to design an economical,
safe and environmentally acceptable coal refuse disposal facility. Designers should recognize that
investigation programs and studies for specifc projects can not realistically be standardized and will
vary according to site conditions, material properties, embankment geometry, hazard classifcation
and proposed staging scheme.
In addition to refuse disposal embankments, this chapter is also applicable to other types of embank-
ments that are employed at mine sites including fresh water dams and sedimentation or treatment
ponds. Other applications and available design guidance are presented in Section 6.3.6.
In the text that follows, many references are made to ASTM International (ASTM) standards. All
references to ASTM standards in this chapter can be found in three volumes of Section Four Con-
struction of the ASTM standards, which are published annually. The most current versions of the
standards should be used. The applicable volumes and their citations herein are:
Volume 05.06 Gaseous Fuels; Coal and Coke (ASTM, 2008a)
Volume 04.08 Soil and Rock (I): D 420 D 5786 (ASTM, 2008b)
Volume 04.09 Soil and Rock (II): D 5877 latest (ASTM, 2008c)
Volume 04.13 Geosynthetics (ASTM, 2008d)
Full citations for these volumes are provided in the References section at the end of this Manual.
6.1 GEOTECHNICAL DESIGN PROCEDURES
This section outlines the steps normally required for designing the geotechnical aspects of a coal
refuse disposal facility. The geotechnical design of a new refuse disposal facility provides fexibilities
that designers should recognize in their planning. These include:
Ability to determine the optimum location for the facility
Flexibility in the design of the starter dam and embankment staging
Ability to coordinate ongoing and future mining with staging in the vicinity of
refuse disposal footprints
Flexibility in designing the internal drainage system and liner system (if required)
Flexibility in the selection and design of hydraulic structures
Geotechnical design for an expansion of an existing disposal facility typically imposes constraints
that designers must recognize in their planning, particularly limited fexibility in planning embank-
ment staging and design of hydraulic structures while maintaining ongoing disposal operations.
Table 6.1 presents guidance for the geotechnical design of refuse disposal facilities with reference to
applicable sections of this Manual and to supplemental documents.
6.2 GENERAL CONSIDERATIONS
6.2.1 Unique Characteristics of Refuse Disposal Facilities
A typical coal refuse disposal facility has unique characteristics and objectives compared to most
other engineered structures. Some of the basic characteristics of coal refuse disposal facilities and
their related signifcance in geotechnical design are identifed in Table 6.2.
< PREVIOUS VIEW
6-3
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
TABLE 6.1 TYPICAL GEOTECHNICAL DESIGN PROCEDURE FOR
COAL REFUSE DISPOSAL FACILITIES
(1)
Design Considerations
Manual Sections for
Reference
Supplemental
References
I. Obtain and Review Available Information
Topographic Maps
Geologic Maps
Soils Maps
Aerial Photographs
Local Experience
Individual Site Mapping
Seismicity Maps
Mine Maps
6.4 USBR (1992a)
II. Plan Field Exploration
What is probable confguration?
What type of hydraulic structures are likely?
Where might important embankments and structures
be located?
What are signifcant foundation characteristics?
What types of borrow material may be required?
What are probable sources of borrow material?
What types of sampling will be required?
What types of environmental control measures should
be considered?
Is past or present mining in the area being considered?
What changes in the disposal program are likely during
the life of the facility?
6.2, Chapter 3
Chapters 3, 5, 9
6.2, Chapter 3
6.2, 6.3, 6.4, 6.6
6.2, 6.3
6.2, 6.3
6.4, 6.5
6.3, Chapters 4, 10
6.3, Chapter 8
Chapters 4, 10
USBR (1992a)
Sherard et al. (1963)
III. Field Exploration
Surfcial Reconnaissance
Geophysical Surveys
Borings and Sampling
Test Pits
Visual Classifcation
Field Testing
Soils and Water Inventory
6.4
Arman et al. (1997)
Hvorslev (1948)
Legget (1962)
USBR (1998, 1992a)
IV. Laboratory Program
Natural Materials
Index Property Tests
Compaction Tests
Hydraulic Conductivity
Consolidation
Shear Strength
Potential Acidity/Neutralization Potential
Refuse Materials:
Index Property Tests
Compaction
Hydraulic Conductivity and Consolidation
Shear Strength
Leachate Quality
6.5
6.5
ASTM (2008b,c)
Lambe (1951)
Bishop and Henkel
(1962)
USBR (1992a)
MSHA (2007)
< PREVIOUS VIEW
6-4
Chapter 6
MAY 2009
Design Considerations
Manual Sections for
Reference
Supplemental
References
V. Design Considerations and Analyses
Foundation Preparation
Seepage Control
Static and Seismic Stability
Settlement
Rock Excavation
Chapters 6, 7, 8
USBR (1987a,1989)
Leonards (1962)
USACE (1993)
VI. Construction Operations
Refuse Transport and Placement
Foundation Preparation
Borrow Materials
Appurtenant Facility Construction
Materials Selection
Quality Control and Field Testing
Chapter 11
USBR (1987a,1989)
Church (1981)
USBR (1998)
Fell et al. (2005)
VII. Instrumentation and Monitoring
Visual Observations
Movements and Displacements
Pore-Water Pressures
Hydrology and Hydraulics
General Maintenance
Chapters 12, 13
USBR (1998)
Dunnicliff (1993)
USACE (1995c)
Note: 1. This table is presented as a guide to qualifed geotechnical designers. Each site must be evaluated
according to conditions at that site. In some cases, studies beyond those identifed in this table may
be needed.
TABLE 6.1 TYPICAL GEOTECHNICAL DESIGN PROCEDURE FOR
COAL REFUSE DISPOSAL FACILITIES
(1)
(Continued)
6.2.2 Site Conditions
In contrast to site selection for a dam that must be built across a valley to form a reservoir of desig-
nated size, site selection for a coal refuse disposal facility is generally fexible because of the variety of
embankment types that can be constructed. The principal considerations in the selection of a disposal
facility site are typically: (1) the potential disposal capacity of the valley/site, (2) the desired prepara-
tion plant processing output, (3) the potential infuence of previous mining activities, (4) refuse trans-
portation and placement, (5) construction of site development and drainage structures, and (6) other
environmental control and safety factors.
6.2.2.1 Topography
Site terrain slopes are very important in disposal facility site selection because of their impact on stor-
age volume, methods and costs of materials handling, methods and costs of drainage control, and
hazard potential.
In areas of rugged and steep terrain, such as southern West Virginia, southwest Virginia, and eastern
Kentucky, most disposal facilities will be valley-type embankments. Valley slopes are ofen too steep
for side-hill embankments. Ridge tops are generally difcult to access and are limited in area to sup-
port large embankments; however, some ridge-top sites can be used in conjunction with mountain-
top surface mining operations. Major valley botoms are too confning for the construction of large
< PREVIOUS VIEW
6-5
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
heaped or diked embankments. Therefore, the typical disposal facility site is generally a small valley
selected by considering:
Potential efects of past or future mining beneath the site
Proximity to existing preparation plants
Surface land ownership
Cost of establishing a materials-handling system suitable for the topography
Ability to sequence construction to handle the types and volumes of refuse to be
disposed
Potential to discharge storm water through or around the site
Stability of existing slopes when modifed by construction or imposed loads
In areas of less severe topography, such as in southwestern Pennsylvania, the range of potential site
location and embankment types is much greater. However, disposal facilities in these areas are fre-
quently located in small valleys because of their capability to accommodate large volumes of refuse
without extensive modifcation of natural topography. Also, the rolling topography ofen makes con-
struction of ridge, side-hill or heaped embankments practical.
In relatively fat terrain, as found in portions of Illinois and Indiana, refuse disposal in valleys may
not be practical, and the refuse disposal facility will generally be of the heaped (Figure 3.12) or diked-
pond embankment type (Figure 3.13. These types of disposal facilities present unique problems for
fne refuse disposal in slurry form because volume containment by utilizing the natural topography
is not possible. When fne refuse slurry is a small percentage of the total refuse, the most economical
disposal facility is a diked-pond embankment constructed from coarse refuse.
TABLE 6.2 BASIC CHARACTERISTICS OF COAL REFUSE DISPOSAL
AND DESIGN SIGNIFICANCE
Basic Characteristic Design Signifcance
The purpose is safe and economical disposal of
refuse.
Greater fexibility in choosing location, confguration and
sequence of placement.
The total disposal volume of coarse refuse is nor-
mally much greater than that required to serve safety
requirements.
Embankment zones may allow different placement specif-
cations (e.g., structural zone).
The refuse disposal occurs over many years and may
lead to several unforeseen events and con straints not
realized at the time of design.
Construction monitoring and quality control may be speci-
fed for the critical construction items with periodical moni-
toring of routine construction.
The geotechnical properties of refuse may not be
available during the design (particularly for a new facil-
ity), and changes in the material properties are prob-
able during the life of the facility.
For new facilities, geotechnical design parameters may
be estimated based on experience and from facilities with
similar characteristics (e.g., similar seam properties, mining
technique, and cleaning process). The geotechnical design
parameters can be verifed when actual samples are avail-
able and can be re-evaluated if characteristics change.
The refuse being placed can have adverse chemical
characteristics that may lead to undesirable environ-
mental conditions or deterioration of construction
materials.
Geotechnical and leachate characteristics of the refuse
should be evaluated based on experience and/or by labora-
tory test, and appropriate amend ment or containment/pro-
tection requirements should be identifed.
After completion of disposal operations, the facility will
need to be abandoned in a safe, economical and envi-
ronmentally acceptable manner.
Planning and design must allow for an acceptable
abandonment confguration with materials for effective
reclamation.
< PREVIOUS VIEW
6-6
Chapter 6
MAY 2009
6.2.2.2 Climate/Weather
Due to the small surface area of coal refuse impoundments, wind, rainfall and temperature condi-
tions are generally not major factors in selecting the most appropriate site for a refuse disposal facility
for a given mining operation. However, the variation of weather conditions in diferent regions of the
country can signifcantly afect disposal facility confguration and design requirements.
For example, in the Appalachian coal region, rainfall is relatively uniform and abundant throughout
most of the year. Thus, addition of water to coal refuse for controlled placement in an embankment
is seldom a major design and cost consideration. A more important criterion in this region may be
to assure that work in valley botoms, or in critical structural portions of the embankment, can be
accomplished during the summer months when conditions are driest.
At the other extreme, such as in the semi-arid Rocky Mountains and Western plains, there is a gen-
eral lack of precipitation for most of the year. In this situation, the designer must evaluate appropri-
ate control measures and associated costs for adding water to refuse to achieve required placement
criteria. Also, the precipitation that does occur is ofen in the form of high intensity thunderstorms,
increasing the required capacity of food control and diversion structures.
Clearly, the planning of otherwise similar disposal facilities will vary according to geographic loca-
tion. Regardless of the site location, the size of the watershed draining into a disposal area should be
minimized unless other design factors justify the cost of major drainage control systems.
Disposal facilities in regions with extended cold winters and signifcant snowfalls may require spe-
cial atention to confguration, materials handling and refuse placement. For example, it may be that
critical structural portions of an embankment can be constructed most economically during the con-
struction season when drainage and material properties are most easily controlled. Overall efciency
is then accomplished by establishing areas for placement of refuse in non-critical areas during the
remainder of the year.
Blowing dust is generally not considered to be a major design problem with coal refuse disposal.
However, dust has been a problem with other types of industrial and mining waste disposal, particu-
larly for disposal of fne-particle materials such as combustion waste.
6.2.2.3 Geology and Surfcial Soils
Coal refuse disposal facility design is afected by geology and surfcial soil conditions as they relate
to the following:
Extent and efects of past or future mining
Necessary foundation treatment
Available borrow material
Type and size of the starter dam
Efects of refuse disposal on groundwater quality and the efects of groundwater
seepage on embankment design
Stability and hydraulic conductivity of the existing foundation materials under natu-
ral and disturbed conditions
Specifc design factors that may be afected by geology and surfcial soil conditions include:
Acceptable embankment slopes as related to the shear strength of the underlying
foundation materials.
< PREVIOUS VIEW
6-7
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
Limitations on the location and design of the impounding facility posed by active or
inactive landslides.
Seepage cutofs through pervious foundation soils for embankments that impound
water.
Embankment construction to account for setlement of underlying sof foundation
materials.
Selection of the starter dam confguration and zoning.
Starter dam and embankment construction rates to keep excess pore-water pressures
to within acceptable limits.
Seals between foundation rock and the impervious zone of an impounding
embankment.
Situations where badly fractured foundation rock must be grouted to control seep-
age from an impounding embankment.
Liner and underdrain systems to address protection of groundwater or seepage into
the embankment.
Selection of decanting and other hydraulic structures to control impounded storm
runof.
Erosion control measures for surface runof and drainage structures
Potential liquefaction of embankment and foundation materials in regions subject to
moderate or high seismic loading.
The wide variations of geology and surfcial soil conditions that may be present make general-
ized examples of these situations impossible. However, if mining has occurred beneath a disposal
facility, it is particularly important to evaluate the efect of existing or potential subsidence on the
embankment and the potential for the contents of the reservoir to break through into the mine. If it is
determined that subsidence has occurred or that new or continued subsidence may occur, potential
detrimental efects on the structural and hydraulic conductivity characteristics of the refuse embank-
ment must be evaluated. The possible efects on groundwater quality due to leachate infltrations into
fractured foundation bedrock, or due to discharge of fne coal refuse into underground mine voids if
the overburden collapses, should also be considered. If mining has not occurred at the planned refuse
disposal facility site, construction of the disposal facility may limit the extraction of underlying coal
reserves, and this cost factor should be considered in the economic evaluation.
Procedures for investigating and analyzing geology and surfcial soil conditions are discussed in Sec-
tion 6.4, and construction aspects of foundation preparation are discussed in Chapter 11. Designers
must understand that an optimum disposal facility can only be achieved if general site conditions are
well understood prior to the selection of the facility site and confguration. This knowledge will help
to prevent unnecessarily conservative designs as well as designs that may be susceptible to increased
maintenance or environmental problems.
6.2.2.4 Miscellaneous Site Considerations
Other considerations that may not be directly related to site conditions, but may infuence the design
and the volume capacity of the facility include:
Access to public roads
Availability of utilities
Future mine developments
Mine infrastructure (e.g., conveyor belts, access roads and mine entries)
< PREVIOUS VIEW
6-8
Chapter 6
MAY 2009
6.2.3 Embankment Materials
The purpose of a coal refuse disposal impoundment is to provide a means for safe and economical
disposal of coal refuse. However, due to safety and operational considerations, other materials are
required for development of the facility. For example, since coarse coal refuse is generally susceptible
to weathering, crushing and degradation, more resistant granular rock is normally needed for con-
struction of drainage zones and for providing erosion protection. Also, at new facilities, disposal of
fne coal refuse requires initial construction of a starter embankment or dam for slurry retention and
setling when sufcient coarse refuse is unavailable during disposal facility startup. Starter dams are
generally constructed with borrow material, unless suitable coarse coal refuse is available on site.
The cost of imported borrow materials is much greater than the coarse refuse, so their use should be
limited to addressing specifc design requirements. Borrow materials can be obtained directly from
mine spoils, from processing of mine spoils, or from a suitable area at or near the site. Embankment
materials that must be purchased from a commercial quarry and transported to the site are the most
costly alternative.
Other examples where site borrow or imported materials may be needed to supplement the refuse
include: (1) fne-grained soils for creating an embankment impervious zone when the material avail-
able will not adequately limit seepage, (2) cover soils suitable for revegetation, and (3) soil and rock
for supplementing embankment construction when coarse refuse is unavailable or inadequate for
slurry retention and storm routing.
The most convenient and economical source for borrow material is typically the area that will eventu-
ally be covered by refuse. Use of this material will minimize transportation costs and will increase the
capacity of the disposal facility, although it can lead to more extensive seepage control and groundwa-
ter protection requirements. Another economical source of borrow material is mine spoil with a matrix
of soil and rock. To avoid unnecessary double handling and stockpiling and to assure that the required
quantities of materials are excavated from the disposal area, careful planning is essential. This precau-
tion is especially important when the borrow material will be used for fnal cover.
When selecting borrow materials from the disposal area, it is important that removal of the material
will not be detrimental to the long-term performance of the facility. For example, if the natural soils
form a desirable impermeable boundary between the refuse and underlying rock, borrow activities
should be restricted to areas of thickest soil cover in order to leave a continuous layer of soil.
The required characteristics for materials used in the construction of refuse embankments are dis-
cussed in Sections 6.4 and 6.6; transport and placement procedures are discussed in Chapter 11. The
following brief discussion of materials generally available for embankment construction is presented
to aid initial planning and design.
6.2.3.1 Coarse Coal Refuse
Non-impounding embankments (refuse piles) are commonly constructed entirely with coarse coal
refuse. Portions of a non-impounding embankment where supplemental borrow material may be
needed include: (1) granular underdrain zones for collecting and discharging groundwater seepage
away from the refuse, (2) granular zones for controlling seepage or collecting leachates, (3) cover
material for promoting vegetation of the embankment surface, and (4) durable, weather-resistant
rock for erosion protection in swales, ditches and channels.
Coarse refuse is typically the predominant material used to construct embankments for impounding
fne coal refuse slurry. However, materials for impervious zones, flters, and drainage zones for these
structures must have specifc characteristics and normally must be obtained from suitable borrow
areas or from commercial sources.
< PREVIOUS VIEW
6-9
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
A summary of published grain-size, specifc-gravity, and strength-testing data for coarse coal
refuse samples, including their geographic source, is presented in Table 6.3. While coarse coal
refuse is generally a well-graded material, signifcant variation, particularly with respect to
the clay-, silt- and fne-sand-size fraction, has been reported depending on geographic loca-
tion (which generally relates to geologic conditions) or coal mining or preparation processes.
Advances in coal mining and preparation processes over the last 15 years have resulted in a trend
toward greater percentages of fnes in coarse coal refuse. While recent published data (Hegazy et
al., 2004) provide evidence of the increase in fnes content of coarse refuse in the northern Appa-
lachian area, resulting in classifcation as silty, clayey sand with gravel to clayey, silty sand with
gravel, other regions typically exhibit lower fnes content with corresponding classifcation as a
well-graded to silty gravel. Reported strength data are consistent with soil and rock content and
gradation.
TABLE 6.3 COARSE COAL REFUSE CHARACTERIZATION SUMMARY
OF AVERAGE/RANGE OF VALUES
Reference Location
Grain Size
Specifc
Gravity
G
s
(gm/cm
3
)
Effective Shear Strength
D
30
(mm)
D
50
(mm)
D
60
(mm)
Passing
No. 200
Sieve
(%)
(degrees)
c
(psf)
Almes and
Butail (1976)
PA, WV,
KY, VA
0.7 2.5 4.5 10 1.8-2.4 33-39 0
McCutcheon
(1981)
OH 1.9 4.5 7 7 2.0 36 NR
Saxena et al.
(1984)
WV 12 16 22 2 2.6 27-40 0-450
Albuquerque
(1994)
VA 3.5 7.5 12 1.5 NR
(1)
39 0
Hegazy et al.
(2004)
PA 0.35 1.23 2.02 19.8 2.0 34 250
Busch et al.
(1974)
WV 0.2-4 1-10 3-20 2-19 1.7-2.3 NR NR
Backer et al.
(1977)
UT, NM 1-6 3-15 6-20 4-15 1.7-2.3 NR NR
Stewart and
Atkins (1983)
Eastern PA
(anthracite)
1-8 5-16 7-22 1-7 2.2-2.4 NR NR
Zeng and
Goble (2008)
Appalachian
region
2 6 9 12 2.5 NR NR
Note: 1. NR = not reported
Compaction, equipment trafc and weathering cause degradation of coarse coal refuse. Larger pieces
of shale will generally crumble to small particles afer exposure to the atmosphere for only a short
period of time. Thus, the percentage of fnes in aged refuse can be noticeably greater than in fresh
or recently placed coarse coal refuse. Based upon sieve analyses results for fresh and compacted
samples, Hegazy et al. (2004) have reported an average increase of fnes of about 4 percent due to
compaction alone.
< PREVIOUS VIEW
6-10
Chapter 6
MAY 2009
6.2.3.2 Fine Coal Refuse
Fine coal refuse, when very wet or in slurry form, is not generally suitable for construction of the
structural portion of an embankment. However, fne refuse can be used as the foundation for por-
tions of an embankment when it has had sufcient time to setle and excess pore-water pressures
have adequately dissipated. Designers are cautioned that embankments that have fne refuse as a
foundation material require comprehensive evaluations and analyses of the following:
Construction schedule
Geotechnical properties and strength
Setlement and seepage properties
Placement procedures
Measures for equipment operator safety
Seismicity, dynamic properties and potential for associated strength loss
Fine coal refuse is the product of extracting, crushing, and cleaning raw coal. The fne coal refuse
slurry is typically pumped upstream of an impounding embankment. The coarser material setles
out more quickly nearer the discharge location (customarily near the upstream slope of the embank-
ment), forming a fnes delta or beach. The fner materials migrate throughout the impoundment,
because they take longer to setle. Thus, samples collected from or near the delta are predominantly
sand and silt-sized particles, whereas samples collected away from the delta are predominantly silt
and clay-sized particles. A summary of published geotechnical data (average and range of values)
for fne coal refuse samples, including their source location, is presented in Table 6.4. This table is
based on samples collected from slurry impoundments and may refect the efect of segregation
that occurs with setling and deposition. Variations in grain size and plasticity occur due to rock
strata, coal extraction and processing, and impoundment depositional characteristics, and thus
properties may vary from those reported in Table 6.4. Site-specifc testing has characterized fne
refuse as plastic clay/silt, low plasticity sandy silt or clay, or low to non-plastic silty/clayey sand
typically exhibiting a lower specifc gravity (and dry density) and lower peak shear strength than
coarse coal refuse.
The grain-size distribution of thickened and dewatered fne coal refuse can be anticipated to be
similar to the averages provided in Table 6.4, as this material does not signifcantly segregate with
placement.
6.2.3.3 Combined Refuse
Some coal preparation plants produce combined refuse that does not require impoundments for
disposal of fne refuse. At these preparation plants, partially dewatered fne refuse flter cake is pro-
duced in addition to coarse refuse. These materials are normally combined, transported and dis-
posed in a non-impounding disposal facility.
Properties of this combined refuse depend on the initial moisture content of the flter cake, the ratio
of fne to coarse refuse, and the particle size and moisture content of the coarse refuse. Ofen, these
properties make it difcult to place combined refuse in a controlled manner, and they may limit its
potential for use within the structural portion of an embankment. Combined refuse is sometimes
mixed with combustion ash for construction of a homogeneous embankment or, in some cases, a
zoned embankment (for construction of the downstream structural shell). Table 6.5 presents pub-
lished geotechnical test data for combined coal refuse from several locations for a range of fnes con-
tent. The strength data are based on remolded samples compacted to 95 percent standard Proctor
maximum dry density, which can be difcult to achieve if the fne portion of the refuse has a high
water content.
< PREVIOUS VIEW
6-11
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
6.2.3.4 Borrow Materials
Borrow materials are those soil and rock materials used in an embankment to meet specifc design
criteria. Borrow materials are used principally for:
Starter dam construction
Filters and drainage zones
Impervious zones
Sedimentation pond embankments
Erosion protection
Butresses
Reclamation cover
For economical designs, most borrow materials are obtained at the site or from suitable nearby
mine spoil. Materials for flters, drains, and erosion protection are typically obtained from com-
mercial sources.
TABLE 6.4 FINE COAL REFUSE CHARACTERIZATION SUMMARY
OF AVERAGE/RANGE OF VALUES
Reference Location
Grain Size
Atterberg
Limits
Specifc
Gravity
G
s
(gm/cm
3
)
Effective Shear
Strength
Passing
No. 40
Sieve
(%)
Passing
No. 200
Sieve
(%)
LL
(%)
PL
(%)
PI
(%)
(degrees)
c
(psf)
Almes and
Butail (1976)
PA, WV,
KY, VA
64-100 36-47 20-40 NR
(1)
<10 1.55-1.65 29-34 0
McCutcheon
(1983)
OH 81 46 29 22 7 1.85 36 0
Qiu and
Sego (2001)
Western
Canada
90 66 40 24 16 1.94 32 200
Hegazy et al.
(2004)
PA 65-100 58 31 20 11 1.52 33 230
Genes et al.
(2000)
WV NR 16-90 NR NR <12 1.44-2.37 23-36 0
Cowherd and
Corda (1998)
NR NR 24-91 23-39 NR 0-9 1.4-2.1 NR NR
Huang et al.
(1987)
KY, OH, PA
TN, VA, WV
NR 27-95 22-44 NR 0-12 1.52-2.14 NR NR
Busch et al.
(1974,1975)
WV 50-98 10-60 34-51 NR 0-13 1.45-2.07 NR NR
Backer et al.
(1977)
UT, NM 60-100 16-98 NR NR NR 1.33-2.07 NR NR
Ullrich et al.
(1991)
KY, TN, OH 45-95 25-85 31-44 NR 0-31 1.8-2.5 NR NR
Zeng and
Goble (2008)
Appalachian
Region
75-85 40-62 27-36 21-26 3-11 2.02-2.16 NR NR
Note: 1. NR = not reported
< PREVIOUS VIEW
6-12
Chapter 6
MAY 2009
The available mine spoil and borrow materials in most coal mining regions of the U.S. consist either
of bedrock or soils derived from bedrock. Typical bedrocks include: (1) sof shales, siltstones and clay-
stones that weather rapidly when excavated and break down to a soil when compacted, (2) harder
limestones that usually resist weathering except when exposed to acidic waters from leachates pass-
ing through pyritic coal and coal refuse, and (3) hard sandstones that ofen are resistant to natural
weathering and atack from leachates. The soil components are typically present as alluvial deposits
in valley botoms, colluvial deposits that have accumulated toward the base of slopes, residual soils
derived from surface weathering of the rock, or partially decomposed weathered rock.
Sof rock is normally suitable for the downstream portion of starter dams not critical to seepage con-
trol. The initial particle size of sof rock may prevent its use for constructing impervious zones, and
its weathering characteristics usually prevent its use for drainage or erosion protection purposes.
When sof rock is used, its design strength characteristics should be based on predicted future com-
pacted or weathered condition, ofen as a soil. The use of limestone should usually be avoided due to
its susceptibility to deterioration from leachates. Limestone can be used in situations where: (1) acidic
conditions will not occur, (2) it is arranged within an embankment in a manner that assures separa-
tion from acidic leachates, or (3) when it is used in a manner that does not depend upon continued
integrity as a granular material. In most coal mining regions, hard sandstone is the best material for
erosion protection, flters and drainage zones.
Mine spoil tends to be highly variable in soil and rock content and particle size, and processing may
be necessary. This can be accomplished by segregating over-sized fractions or fnes, which can be
used for other project applications. Sometimes mine spoil can be used for construction without pro-
cessing of the materials (e.g., in a zoned embankment).
Recent alluvial and older river terrace deposits can vary from relatively clean sands and gravels to
dirty soils with high contents of fne-grained soils and organic material. Material from clean sand
and gravel deposits may be suitable for constructing flter or drainage zones in an embankment,
but only afer feld investigation has determined the quantity of clean material available and labora-
tory testing has verifed the suitability of the grain-size distribution and mineral composition for the
intended purposes. Otherwise, alluvial soils typically are not desirable for use in a coal refuse dis-
posal embankment because of the expense associated with excavation and preparation.
Colluvial soils generally consist of a combination of soils derived from fne- and coarse-grained rocks
and vary from clays to primarily sandy material. Fine-grained colluvial soils may be suitable for con-
structing an impervious zone, while all types of colluvium are normally suitable for use as structural
TABLE 6.5 COMBINED COAL REFUSE CHARACTERIZATION
Location
Grain Size
Specifc Gravity
G
s
(gm/cm
3
)
Total Shear Strength
Passing
No. 4 Sieve
(percent)
Passing
No. 200 Sieve
(percent)
(degrees)
c
(psf)
Pennsylvania 25-60 7-26 1.9-2.0 26-28 NR
Ohio 25-58 5-11 1.8 36-38 NR
West Virginia 18-58 4-12 2.1 32 NR
Colorado 18-54 3-18 1.8-1.9 33-38 NR
(STEWART AND ATKINS, 1982)
< PREVIOUS VIEW
6-13
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
components for stability or as cover material to support vegetation. Because they generally have
a wide grain-size distribution, colluvial soils normally are not suitable for either flter or drainage
zones in an embankment.
Residual soils derived from sof rocks (e.g., shale) are normally fne-grained and suitable for either
an impervious zone or the structural portion of an embankment. When available in thick deposits,
residual soils can generally be excavated economically. Soils derived from sandstone are too coarse
for use in impervious zones, but may be suitable for structural fll or use in drainage zones.
For any borrow material, the deposit must be explored to verify that it is available in adequate quanti-
ties to meet the design requirements of the disposal facility. This must be followed by laboratory testing
to evaluate the suitability of the borrow materials/mine spoils for embankment construction. Although
practically any inorganic, insoluble soil can be incorporated into an embankment when modern com-
paction equipment and control standards are employed, the following problems may arise:
Fine-grained soils may have insufcient shear strength or excessive compressibility.
Clays of medium to high plasticity may expand if placed under low confning pres-
sures and/or at low moisture contents.
Plastic soils with high natural moisture content may be difcult to adjust for proper
moisture for compaction.
Dispersive clays are not suitable for use in dam embankments.
Silts may have insufcient erosion resistance.
Stratifed soils may require extensive mixing of borrow material.
Table 6.6 shows a correlation between soil classifcation and the engineering and design properties
of compacted soils (DOD, 2005). This table can be used for evaluation of borrow materials and pre-
liminary design of starter dams. Table 6.7 provides a correlation between soil classifcation and the
relative desirability of soil as compacted fll material for various types of starter embankments. Table
6.8 (Sherard et al., 1963) illustrates a correlation between soil classifcation and engineering properties
related to embankment design and constructability.
6.2.3.5 Coal Combustion Products from Power Plants
The combustion of coal at fossil fuel power plants produces fy ash and botom ash as residual waste
products. Two other products of coal combustion air pollution control technology are fuidized-bed
combustion (FBC) waste and fue-gas-desulfurization (FGD) sludge. While no detailed assessment
of these and other wastes from the power plants is provided in this Manual, embankments at some
refuse disposal sites have been constructed by integrating power plant waste products with coal
refuse. The important issues that designers should consider if these waste products are disposed at
coal refuse facilities are discussed in this section.
Power plant wastes have certain properties that can be very benefcial if these materials are judiciously
integrated at coal refuse disposal sites. The economic viability of disposing dissimilar refuse materi-
als can be especially benefcial when the mining operation is near the coal burning plant. With this
in view, a brief description of the engineering properties of power plant waste materials is provided
in this section. For more detailed discussion and design considerations, the designer should refer
to the Electric Power Research Institute (EPRI) Coal Ash Disposal Manual (DiGioia et al., 1995) and
other references such as McLaren and DiGioia (1987), DiGioia and Gray (1979), Gray and Lin (1972).
Additionally, research has been published (Daniel et al., 2002) on the material properties of mixtures
of fy ash and coal refuse in southwest Virginia. A thorough review of the overall environmental
implications of fy ash use is provided in Carlson and Adriano (1993). The USEPA (2000) performed a
< PREVIOUS VIEW
6-14
Chapter 6
MAY 2009
T
A
B
L
E
6
.
6
C
O
R
R
E
L
A
T
I
O
N
B
E
T
W
E
E
N
U
S
C
S
C
L
A
S
S
I
F
I
C
A
T
I
O
N
A
N
D
P
R
O
P
E
R
T
I
E
S
O
F
C
O
M
P
A
C
T
E
D
S
O
I
L
S
G
r
o
u
p
S
y
m
b
o
l
S
o
i
l
T
y
p
e
R
a
n
g
e
o
f
M
a
x
.
D
r
y
W
e
i
g
h
t
(
p
c
f
)
R
a
n
g
e
o
f
O
p
t
i
m
u
m
M
o
i
s
t
u
r
e
(
%
)
T
y
p
i
c
a
l
V
a
l
u
e
o
f
C
o
m
p
r
e
s
s
i
o
n
T
y
p
i
c
a
l
S
t
r
e
n
g
t
h
C
h
a
r
a
c
t
e
r
i
s
t
i
c
s
T
y
p
i
c
a
l
H
y
d
.
C
o
n
d
.
(
f
t
/
m
i
n
)
R
a
n
g
e
o
f
C
B
R
V
a
l
u
e
R
a
n
g
e
o
f
S
u
b
g
r
a
d
e
M
o
d
u
l
u
s
k
(
l
b
s
/
i
n
3
)
1
.
4
t
s
f
=
2
0
p
s
i
3
.
6
t
s
f
=
5
0
p
s
i
C
o
m
p
a
c
t
e
d
C
o
h
e
s
i
o
n
(
p
s
f
)
S
a
t
u
r
a
t
e
d
C
o
h
e
s
i
o
n
(
p
s
f
)
(
d
e
g
)
t
a
n
(
%
o
f
o
r
i
g
i
n
a
l
h
e
i
g
h
t
)
G
W
W
e
l
l
g
r
a
d
e
d
c
l
e
a
n
g
r
a
v
e
l
s
,
g
r
a
v
e
l
-
s
a
n
d
m
i
x
t
u
r
e
s
1
2
5
-
1
3
5
1
1
-
8
0
.
3
0
.
6
0
0
>
3
8
>
0
.
7
9
5
1
0
-
2
4
0
-
8
0
3
0
0
-
5
0
0
G
P
P
o
o
r
l
y
-
g
r
a
d
e
d
c
l
e
a
n
g
r
a
v
e
l
s
,
g
r
a
v
e
l
-
s
a
n
d
m
i
x
1
1
5
-
1
2
5
1
4
-
1
1
0
.
4
0
.
9
0
0
>
3
7
>
0
.
7
4
1
0
-
1
3
0
-
6
0
2
5
0
-
4
0
0
G
M
S
i
l
t
y
g
r
a
v
e
l
s
,
p
o
o
r
l
y
-
g
r
a
d
e
d
g
r
a
v
e
l
-
s
a
n
d
m
i
x
1
2
0
-
1
3
5
1
2
-
8
0
.
5
1
.
1
>
3
4
>
0
.
6
7
>
1
0
-
6
2
0
-
6
0
1
0
0
-
4
0
0
G
C
C
l
a
y
e
y
g
r
a
v
e
l
s
,
p
o
o
r
l
y
g
r
a
d
e
d
g
r
a
v
e
l
-
s
a
n
d
-
c
l
a
y
1
1
5
-
1
3
0
1
4
-
9
0
.
7
1
.
6
>
3
1
>
0
.
6
0
>
1
0
-
7
2
0
-
4
0
1
0
0
-
3
0
0
S
W
W
e
l
l
-
g
r
a
d
e
d
c
l
e
a
n
s
a
n
d
s
,
g
r
a
v
e
l
l
y
s
a
n
d
s
1
1
0
-
1
3
0
1
6
-
9
0
.
6
1
.
2
0
0
3
8
0
.
7
9
>
1
0
-
3
2
0
-
4
0
2
0
0
-
3
0
0
S
P
P
o
o
r
l
y
-
g
r
a
d
e
d
c
l
e
a
n
s
a
n
d
s
,
s
a
n
d
-
g
r
a
v
e
l
m
i
x
1
0
0
-
1
2
0
2
1
-
1
2
0
.
8
1
.
4
0
0
3
7
0
.
7
4
>
1
0
-
3
1
0
-
4
0
2
0
0
-
3
0
0
S
M
S
i
l
t
y
s
a
n
d
s
,
p
o
o
r
l
y
-
g
r
a
d
e
d
s
a
n
d
-
s
i
l
t
m
i
x
1
1
0
-
1
2
5
1
6
-
1
1
0
.
8
1
.
6
1
0
5
0
4
2
0
3
4
0
.
6
7
5
1
0
-
5
1
0
-
4
0
1
0
0
-
3
0
0
S
M
-
S
C
S
a
n
d
-
s
i
l
t
-
c
l
a
y
m
i
x
w
i
t
h
s
l
i
g
h
t
l
y
p
l
a
s
t
i
c
f
n
e
s
1
1
0
-
1
3
0
1
5
-
1
1
0
.
8
1
.
4
1
0
5
0
3
0
0
3
3
0
.
6
6
2
1
0
-
6
5
-
3
0
1
0
0
-
3
0
0
S
C
C
l
a
y
e
y
s
a
n
d
s
,
p
o
o
r
l
y
g
r
a
d
e
d
s
a
n
d
-
c
l
a
y
m
i
x
1
0
5
-
1
2
5
1
9
-
1
1
1
.
1
2
.
2
1
5
5
0
2
3
0
3
1
0
.
6
0
5
1
0
-
7
5
-
2
0
1
0
0
-
3
0
0
M
L
I
n
o
r
g
a
n
i
c
s
i
l
t
s
a
n
d
c
l
a
y
e
y
s
i
l
t
s
9
5
-
1
2
0
2
4
-
1
2
0
.
9
1
.
7
1
4
0
0
1
9
0
3
2
0
.
6
2
>
1
0
-
5
1
5
1
0
0
-
2
0
0
M
L
-
C
L
M
i
x
t
u
r
e
o
f
i
n
o
r
g
a
n
i
c
s
i
l
t
a
n
d
c
l
a
y
1
0
0
-
1
2
0
2
2
-
1
2
1
.
0
2
.
2
1
3
5
0
4
6
0
3
2
0
.
6
2
5
1
0
-
7
C
L
I
n
o
r
g
a
n
i
c
c
l
a
y
s
o
f
l
o
w
t
o
m
e
d
i
u
m
p
l
a
s
t
i
c
i
t
y
9
5
-
1
2
0
2
4
-
1
2
1
.
3
2
.
5
1
8
0
0
2
7
0
2
8
0
.
5
4
>
1
0
-
7
1
5
5
0
-
2
0
0
O
L
O
r
g
a
n
i
c
s
i
l
t
s
a
n
d
s
i
l
t
-
c
l
a
y
s
,
l
o
w
p
l
a
s
t
i
c
i
t
y
8
0
-
1
0
0
3
3
-
2
1
5
5
0
-
1
0
0
M
H
I
n
o
r
g
a
n
i
c
c
l
a
y
e
y
s
i
l
t
s
,
e
l
a
s
t
i
c
s
i
l
t
s
7
0
-
9
5
4
0
-
2
4
2
.
0
3
.
8
1
5
0
0
4
2
0
2
5
0
.
4
7
5
1
0
-
7
1
0
5
0
-
1
0
0
C
H
I
n
o
r
g
a
n
i
c
c
l
a
y
s
o
f
h
i
g
h
p
l
a
s
t
i
c
i
t
y
7
5
-
1
0
5
3
6
-
1
9
2
.
6
3
.
9
2
1
5
0
2
3
0
1
9
0
.
3
5
>
1
0
-
7
1
5
5
0
-
1
5
0
O
H
O
r
g
a
n
i
c
c
l
a
y
s
a
n
d
s
i
l
t
y
c
l
a
y
s
6
5
-
1
0
0
4
5
-
2
1
-
-
-
-
-
-
-
5
2
5
-
1
0
0
N
o
t
e
:
1
.
A
l
l
p
r
o
p
e
r
t
i
e
s
a
r
e
f
o
r
c
o
n
d
i
t
i
o
n
o
f
S
t
a
n
d
a
r
d
P
r
o
c
t
o
r
m
a
x
i
m
u
m
d
e
n
s
i
t
y
,
e
x
c
e
p
t
v
a
l
u
e
s
o
f
k
a
n
d
C
B
R
w
h
i
c
h
a
r
e
f
o
r
M
o
d
i
f
e
d
P
r
o
c
t
o
r
m
a
x
i
m
u
m
d
e
n
s
i
t
y
.
2
.
T
y
p
i
c
a
l
s
t
r
e
n
g
t
h
c
h
a
r
a
c
t
e
r
i
s
t
i
c
s
a
r
e
f
o
r
e
f
f
e
c
t
i
v
e
s
t
r
e
n
g
t
h
e
n
v
e
l
o
p
e
s
a
n
d
a
r
e
o
b
t
a
i
n
e
d
f
r
o
m
U
S
B
R
d
a
t
a
.
3
.
C
o
m
p
r
e
s
s
i
o
n
v
a
l
u
e
s
a
r
e
f
o
r
v
e
r
t
i
c
a
l
l
o
a
d
i
n
g
w
i
t
h
c
o
m
p
l
e
t
e
l
a
t
e
r
a
l
c
o
n
f
n
e
m
e
n
t
.
4
.
(
-
)
i
n
d
i
c
a
t
e
s
i
n
s
u
f
f
c
i
e
n
t
d
a
t
a
a
v
a
i
l
a
b
l
e
f
o
r
a
n
e
s
t
i
m
a
t
e
.
(
D
O
D
,
2
0
0
5
)
< PREVIOUS VIEW
6-15
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
T
A
B
L
E
6
.
7
C
O
R
R
E
L
A
T
I
O
N
B
E
T
W
E
E
N
U
S
C
S
C
L
A
S
S
I
F
I
C
A
T
I
O
N
A
N
D
R
E
L
A
T
I
V
E
D
E
S
I
R
A
B
I
L
I
T
Y
O
F
S
O
I
L
S
A
S
C
O
M
P
A
C
T
E
D
F
I
L
L
G
r
o
u
p
S
y
m
b
o
l
S
o
i
l
T
y
p
e
R
e
l
a
t
i
v
e
D
e
s
i
r
a
b
i
l
i
t
y
f
o
r
V
a
r
i
o
u
s
U
s
e
s
(
1
i
s
m
o
s
t
d
e
s
i
r
a
b
l
e
.
1
4
i
s
l
e
a
s
t
d
e
s
i
r
a
b
l
e
)
R
o
l
l
e
d
E
a
r
t
h
F
i
l
l
D
a
m
s
L
i
n
i
n
g
F
o
u
n
d
a
t
i
o
n
H
o
m
o
g
e
n
o
u
s
E
m
b
a
n
k
m
e
n
t
C
o
r
e
S
h
e
l
l
E
r
o
s
i
o
n
R
e
s
i
s
t
a
n
c
e
C
o
m
p
a
c
t
e
d
E
a
r
t
h
L
i
n
i
n
g
R
o
a
d
w
a
y
S
u
r
f
a
c
i
n
g
S
e
e
p
a
g
e
I
m
p
o
r
t
a
n
t
S
e
e
p
a
g
e
N
o
t
I
m
p
o
r
t
a
n
t
G
W
W
e
l
l
g
r
a
d
e
d
c
l
e
a
n
g
r
a
v
e
l
s
,
g
r
a
v
e
l
-
s
a
n
d
m
i
x
t
u
r
e
s
-
-
1
1
-
3
-
1
G
P
P
o
o
r
l
y
-
g
r
a
d
e
d
c
l
e
a
n
g
r
a
v
e
l
s
,
g
r
a
v
e
l
-
s
a
n
d
m
i
x
-
-
2
2
-
-
-
3
G
M
S
i
l
t
y
g
r
a
v
e
l
s
,
p
o
o
r
l
y
-
g
r
a
d
e
d
g
r
a
v
e
l
-
s
a
n
d
m
i
x
2
4
-
4
4
5
1
4
G
C
C
l
a
y
e
y
g
r
a
v
e
l
s
,
p
o
o
r
l
y
g
r
a
d
e
d
g
r
a
v
e
l
-
s
a
n
d
-
c
l
a
y
1
1
-
3
1
1
2
6
S
W
W
e
l
l
g
r
a
d
e
d
c
l
e
a
n
s
a
n
d
s
,
g
r
a
v
e
l
l
y
s
a
n
d
s
-
-
3
,
i
f
g
r
a
v
e
l
l
y
6
-
4
-
2
S
P
P
o
o
r
l
y
g
r
a
d
e
d
c
l
e
a
n
s
a
n
d
s
,
s
a
n
d
-
g
r
a
v
e
l
m
i
x
-
-
4
,
i
f
g
r
a
v
e
l
l
y
7
,
i
f
g
r
a
v
e
l
l
y
-
-
-
5
S
M
S
i
l
t
y
s
a
n
d
s
,
p
o
o
r
l
y
g
r
a
d
e
d
s
a
n
d
-
s
i
l
t
m
i
x
4
5
-
8
,
i
f
g
r
a
v
e
l
l
y
5
,
e
r
o
s
i
o
n
c
r
i
t
i
c
a
l
6
3
7
S
C
C
l
a
y
e
y
s
a
n
d
s
,
p
o
o
r
l
y
g
r
a
d
e
d
s
a
n
d
-
c
l
a
y
m
i
x
3
2
-
5
2
2
4
8
M
L
I
n
o
r
g
a
n
i
c
s
i
l
t
s
a
n
d
c
l
a
y
e
y
s
i
l
t
s
6
6
-
-
6
,
e
r
o
s
i
o
n
c
r
i
t
i
c
a
l
-
6
9
C
L
I
n
o
r
g
a
n
i
c
c
l
a
y
s
o
f
l
o
w
t
o
m
e
d
i
u
m
p
l
a
s
t
i
c
i
t
y
5
3
-
9
3
7
5
1
0
O
L
O
r
g
a
n
i
c
s
i
l
t
s
a
n
d
s
i
l
t
-
c
l
a
y
s
,
l
o
w
p
l
a
s
t
i
c
i
t
y
8
8
-
-
7
,
e
r
o
s
i
o
n
c
r
i
t
i
c
a
l
-
7
1
1
M
H
I
n
o
r
g
a
n
i
c
c
l
a
y
e
y
s
i
l
t
s
,
e
l
a
s
t
i
c
s
i
l
t
s
9
9
-
-
-
-
8
1
2
C
H
I
n
o
r
g
a
n
i
c
c
l
a
y
s
o
f
h
i
g
h
p
l
a
s
t
i
c
i
t
y
7
7
-
1
0
8
,
v
o
l
.
-
c
h
a
n
g
e
c
r
i
t
i
c
a
l
-
9
1
3
O
H
O
r
g
a
n
i
c
c
l
a
y
s
a
n
d
s
i
l
t
y
c
l
a
y
s
1
0
1
0
-
-
-
-
1
0
1
4
(
D
O
D
,
2
0
0
5
)
< PREVIOUS VIEW
6-16
Chapter 6
MAY 2009
detailed review of the use of coal combustion products in mining environments that supported their
classifcation as residual waste, although they recommended continued study of disposal in deep
mines and in mine backfll situations where the materials may contact groundwater.
Fly ash is a fne, silt-sized material usually ranging in diameter from 0.5 to 100 microns and consist-
ing largely of spherical, sometimes hollow, glassy particles. Botom ash consists of primarily coarser
material with heavier particles than fy ash. It is generally angular with a porous surface. For fy ash,
hydraulic conductivities have been reported in the range of 10
-7
to 10
-4
cm/sec and for botom ash in
the range of 10
-3
to 10
-1
cm/sec (DiGioia et al., 1995). Depending on the actual material, fne-grained
fy ash and coarse-grained botom ash can be used as an impervious liner and drainage flter,
respectively, in conjunction with coarse coal refuse and other borrow materials. If fy ash is enriched
with nitrogen compounds, it can sometimes be used as a supplement for vegetation growth.
There are two general types of fy ash as defned by ASTM. Class F fy ash contains less than 20 per-
cent calcium oxide and is produced by burning bituminous or anthracite coal. Class C fy ash is pro-
duced by burning subbituminous coal or lignite. Both have pozzolanic properties, but Class F fy ash
is not appreciably self-cementing. Because of the geographical distribution of coal types, Class F fy
ash is principally produced in the eastern U.S., while most Class C fy ash is produced in the western
U.S. Class C fy ash is self-cementing due to presence of lime and other chemical compounds. How-
ever, much of the fy ash returned to Appalachian mines does not meet either Class C or F criteria
(Daniels et al., 2002).
TABLE 6.8 APPROXIMATE CORRELATION BETWEEN ENGINEERING PROPERTIES
AND SOIL CLASSIFICATION GROUPS
USCS
Group
Symbol
Relative
Hydraulic
Conductivity
Probable
Range of k
(ft/yr)
Relative Piping
Resistance
Relative Shear
Strength
Relative
Workability
(1)
GW Pervious 1,000 100,000 High Very High Very Good
GP
Pervious to Very
Pervious
5,000
10,000,000
High to Medium High Very Good
GM Semi-pervious 0.1 100 High to Medium High Very Good
GC Impervious 0.01 10 Very High High Very Good
SW Pervious 500 50,000 High to Medium Very High Very Good
SP
Pervious to Semi-
pervious
50 500,000 Low to Very Low High Good to Fair
SM
Semi-pervious to
Impervious
0.1 500 Medium to Low High Good to Fair
SC Impervious 0.01 50 High High to Medium Good to Fair
ML Impervious 0.01 50 Low to Very Low Medium to Low Fair to Very Poor
CL Impervious 0.01 1 High Medium Good to Fair
OL Impervious 0.01 10 Medium Low Fair to Poor
MH Very Impervious 0.001 0.1 Medium to High Low Poor to Very Poor
CH Very Impervious 0.0001 0.01 Very High Low to Medium Very Poor
Note: 1. Relative workability = ease of moisture-density control.
(ADAPTED FROM SHERARD ET AL., 1963)
< PREVIOUS VIEW
6-17
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
Power plants may also generate combustion waste from FBC systems and SO
2
scrubbers by utilizing
limestone. FBC waste may be pozzolanic or cementitious.
The physical and engineering properties of coal ash that could be of importance when it is used in
combination with coal refuse disposal are grain size, specifc gravity, density, optimum moisture con-
tent, hydraulic conductivity, shear strength, and compressibility.
Tables 6.9 and 6.10 show summaries of particle-size testing results for Class F and Class C fy ash,
respectively.
TABLE 6.9 CLASS F FLY ASH
Gradation
Property
Number of
Samples
Mean Value
(mm)
Standard Deviation
(mm)
Coeffcient of
Variation
D
85
84 0.079 0.063 0.800
D
50
84 0.023 0.015 0.669
D
15
84 0.0075 0.0048 0.648
(MCLAREN AND DIGIOIA, 1987)
TABLE 6.10 CLASS C FLY ASH
Gradation
Property
Number of
Samples
Mean Value
(mm)
Standard Deviation
(mm)
Coeffcient of
Variation
D
85
17 0.063 0.020 0.317
D
50
17 0.022 0.011 0.500
D
15
17 0.0084 0.0082 0.976
(McCLAREN AND DIGIOIA, 1987)
A review of available data (McLaren and DiGioia, 1987) shows that fy ash is a relatively uniform, silt-
sized material with a specifc gravity slightly lower than most natural soils. Compaction of fy ash is
moisture dependent, but the range of optimum moisture contents is greater than that of natural silts
and silty clays. The maximum dry and wet densities of compacted fy ash are somewhat less than
typical values for natural soils, which makes fy ash useful as a light-weight structural fll.
Of importance when fy ash is used in structural zones of a disposal facility is shear strength. The
efective angle of internal friction of fy ash and FBC combustion waste varies with the degree of
compaction, but generally ranges from 25 to 40 degrees. Class F fy ash is non-cohesive, and while
it may appear to be cohesive when partially saturated, this efect is completely lost when the mate-
rial is either dried or saturated. In contrast, Class C fy ash can develop considerable cohesive shear
strength due to cementitious reactions. This cohesion is the dominant factor in the shear strength of
Class C fy ash. Similar to fy ash, the shear strength of botom ash varies with the degree of compac-
tion. The efective angle of friction for botom ash in a loose state can vary from 38 to 42.5 degrees,
with an average of about 41 degrees. The shear strength of SO
2
sludge varies signifcantly with the
solids content and the amount of stabilizing agent added. The strength of a stabilized sludge is
comparable to dense sand and gravel, while the strength of unstabilized sludge is similar to that of
loose sand. The angle of friction for stabilized sludge ranges from 38 to 51 degrees depending on
the solids content and amount of additives. The angle of friction ranges from 20 to 30 degrees for
unstabilized sludge.
< PREVIOUS VIEW
6-18
Chapter 6
MAY 2009
Research into the benefcial reuse of fy ash mixed with coarse coal refuse in southwest Virginia
was performed by Daniels et al. (2002). Table 6.11 shows variations of maximum compacted dry
density, shear strength and hydraulic conductivity reported by them for mixtures of fy ash and
coal refuse.
TABLE 6.11 FLY ASH/COAL REFUSE MIXTURE PROPERTIES
Fly Ash Mix Ratio
(%)
Maximum
Dry Density
dmax
(lb/ft
3
)
Effective Angle of
Internal Friction
(degrees)
Effective
Cohesion
c
(lb/ft
2
)
Hydraulic
Conductivity
k
(cm/sec)
0 125 39 0 2.86 x 10
-3
8 123 37.7 0 1.01 x 10
-3
16 120 37 0 2.56 x 10
-4
24 119 37 0 1.71 x 10
-4
32 117 37 0 7.88 x 10
-5
100 85 37 0 5.78 x 10
-5
(DANIELS ET AL., 2002)
Research conducted on coal refuse and Type F fy ash from southern Illinois (Kumar et al., 2001) indi-
cated an increase in strength for mixtures containing up to 15 percent ash. There was a tendency for
strength to decrease for mixtures with a higher percentage of ash.
Mixing procedures to blend fy ash with refuse material should be developed based on the strength
or stabilization requirements for the embankment. In many cases, the primary benefcial use is asso-
ciated with moisture control, and for this usage blending with spreading equipment on the embank-
ment surface is acceptable. Where enhancement of the strength of the refuse material is a requirement,
or in cases where fy ash amendments are introduced to control acid generation, greater efort to mix
or blend the materials may be required or benefcial. Additional discussion concerning blending of
amendments is presented in Section 11.5.6.
6.2.4 Scheduling
The procedure for scheduling construction of a disposal facility embankment difers from that used
for most other types of constructed embankments for several reasons:
Refuse disposal occurs continually over many years.
Disposal must occur on a year-round basis, regardless of weather conditions.
The rate of refuse disposal is not only determined by the needs of embankment
construction, but also by the rate of mining, the quality of the seam, market changes,
and scheduled and unscheduled work interruptions.
The many variables afecting refuse disposal scheduling limit the ability to provide a detailed fow
chart that accounts for all construction events that could occur during the operational period of a
disposal facility. However, the following are questions to be answered in developing the design and
establishing a general construction schedule for existing and new disposal facilities:
What is the expected operational period of the disposal facility?
What volumes of coarse, fne and combined refuse are anticipated from the prepara-
< PREVIOUS VIEW
6-19
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
tion plant on an annual basis, and are there reasons to believe that the relative pro-
portions may change?
What types of borrow material are available, and must they be developed prior to
covering the borrow site with refuse?
What embankment confgurations would enable critical sections to be constructed
from available materials during periods of the year most conducive to controlled
construction?
What measures to address potential environmental impacts and protection of surface
and groundwater will be required?
Should hydraulic structures (e.g., spillways, decants, diversion systems) be con-
structed in stages corresponding to disposal rates?
Can soil and rock materials excavated for hydraulic structures be used as borrow for
embankment construction?
What measures will be required to economically abandon the disposal facility at the
end of its expected operational period?
If for some reason the disposal facility must be abandoned earlier than expected, can
that abandonment be reasonably accomplished?
Will the disposal rate allow portions of the embankment to be completed in stages so
that exposed refuse and eventual abandonment costs are minimized?
If the rate of refuse production or the operational period of the disposal facility is
extended beyond initial projections, can the facility be reasonably expanded?
Will the timing of future surface or underground mining in the vicinity of the facility
impact the design?
The designer must recognize that these schedule-related questions may not address all of the poten-
tial issues. However, it is prudent to consider each of the above questions when designing a coal
refuse disposal facility.
6.3 DESIGN CONSIDERATIONS
This section presents the basic design considerations for coal refuse embankments and earthen dams.
Embankments with and without impoundments are considered separately for two reasons. First, an
impounding embankment generally has a greater hazard potential because the impounded water can
cause damage for a substantial distance downstream from the disposal facility in the event of a fail-
ure. Second, impounded water can contribute directly to a failure if the embankment is improperly
designed and/or constructed.
The examples discussed in this section are based on the assumption that coarse refuse is the major
structural component of embankments for disposal of coarse and fne coal refuse. Mine spoil, soil
and rock borrow materials, which are obtainable at most disposal facility sites, could also be used
for the starter dam and in the structural portions of the embankment. The designer should refer to
Sherard, et al. (1963) and USBR (1992a; 1998) for additional discussion of design principles for dams
and embankments composed primarily of soil and rock materials.
6.3.1 Impounding Embankments
A coal refuse impounding embankment is generally designed based upon the same principles as
earthen dams, with the exception that refuse materials are used to the maximum extent possible. Two
major aspects of coal refuse impounding embankment construction are: (1) the starter dam for initial
disposal of fne refuse and (2) the embankment raising methodology used for long-term refuse dis-
posal. The starter dam is typically constructed with borrow materials, while subsequent crest raisings
generally utilize coarse coal refuse.
< PREVIOUS VIEW
6-20
Chapter 6
MAY 2009
The borrow material available at a site may range from fne to coarse soils to coarse coal refuse, or
rocks and soil from mine spoils and highwall cutings. For economy, suitable materials at or near the
disposal site should be used for starter dam construction. Depending on the type of material avail-
able at the site, the following types of starter dam may be designed:
Homogeneous Embankment
Zoned Embankment
A homogeneous embankment is generally constructed in situations where the borrow materials vary
litle in hydraulic conductivity or soil type. A zoned embankment is constructed where two or more
types of materials are available for embankment construction. Rockfll, when large quantities of rocks
are available from the mine spoil or from spillway construction, can also be used as the stability por-
tion of a zoned embankment.
The principal geotechnical considerations in designing an impounding embankment are:
Seepage control internal drainage system, impervious zone, and foundation
treatment
Slope and foundation stability static slope stability, end-of-construction stability,
seismic slope stability and deformation, sloughing and erosion
Drainage structures principal and emergency spillways, conduits and surface
drainage structures
Underground mine workings stability, subsidence and breakthrough potential,
sealing of mine openings and boreholes, and potential infltration into underground
mine workings
For earth dams, foundation support, stability and seepage control are important design consider-
ations. Geologic and geotechnical investigations should be performed to identify potentially unstable
soils that are incapable of sustaining embankment loadings or susceptible to adverse impacts from
seepage. Slope failure and piping (internal erosion) are the most common types of embankment fail-
ure associated with seepage. In addition to embankment slope and material strength and unit weight,
the stability of an embankment is a function of the depth of the saturation level (seepage phreatic
surface) below the embankment face, regardless of the volume of seepage through the embankment.
If the phreatic surface rises above the level assumed in the design, embankment stability can decrease
to the point where failure occurs. Piping is a process in which particles are carried out of an embank-
ment with seepage, creating voids. This can lead to failure as the voids progressively extend farther
back into the embankment as more material is removed and more concentrated seepage fow occurs.
Installation of an internal drainage system within the embankment is specifcally intended to address
seepage-type failures and is fundamental to the selection of the embankment confguration. The sub-
sequent discussions of basic embankment types repeatedly emphasize the importance of seepage
control. Cedergren (1989) is an important reference for analyzing seepage conditions in embank-
ments and foundations.
In addition to controlling seepage for embankment stabilization, liner systems may also be needed
for mitigation of potential environmental impacts to the groundwater and surface water. Impervious
liners composed of fne-grained borrow materials, geomembranes or geosynthetic clay liners can be
a critical component of disposal facility design.
Uncontrolled seepage into underground mine workings can afect embankment safety. If the water
pressure in an impoundment has the potential for breaking through the overburden into abandoned
entries or through zones of sof, weathered rock, the rapid release of water or slurry into a mine could
< PREVIOUS VIEW
6-21
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
trap mine personnel and equipment and lead to undesirable environmental conditions. Such a situ-
ation must be prevented. Even in abandoned workings, the resulting water fow may endanger the
population downstream of uncontrolled mine discharge points and cause undesirable environmental
conditions. Guidance related to evaluation of breakthrough potential is presented in Section 8.5.
Analysis of seepage, slope stability and structure foundations is generally performed in detail only
afer the embankment confguration has been selected and the subsurface exploration and labora-
tory testing programs have been completed. These analyses are introduced in this section and are
discussed in more detail in Section 6.6.
6.3.1.1 Homogeneous Embankments
A homogeneous embankment is constructed of only one material; however, to control seepage and sat-
uration of the embankment, a granular internal drain is typically incorporated into the cross section.
Homogeneous embankments constructed for coal refuse sites with various types of internal drains are
illustrated in Figure 6.1. In all cases, the purpose of the drain is: (1) to maintain stability by keeping the
phreatic surface low and (2) to control seepage as it leaves the embankment to minimize the potential
for piping. The drainage system may consist of one or more flter zones of intermediate grain-size mate-
rial to mitigate potential conveyance of embankment material into the collection zone. Selection of the
relative gradations of adjacent zones of material and the use of geotextiles to prevent fner material from
piping into a coarser downstream zone is discussed in Section 6.6.2. If possible, drainage material should
be selected to act as both the flter and the collection zone to avoid the higher costs and more difcult
construction associated with placing multiple layers. Selection of the type of internal drainage system
is normally based upon the fne refuse and food pool levels, embankment confguration, and material
characteristics, including anisotropy. Table 6.12 summarizes the advantages and disadvantages of the
internal drains that are illustrated in Figure 6.1.
When layers of coal refuse are placed during refuse embankment construction, the top surface is
ofen broken into smaller-grained, less-permeable material by the movement of equipment and the
efects of weathering. The materials beneath the top surface retain their original grain-size distribu-
tion and greater hydraulic conductivity. As a result, anisotropic conditions can develop, leading to
an embankment hydraulic conductivity that is greater in the horizontal direction than in the verti-
cal direction. Thus, the efectiveness of a horizontal drainage blanket or toe drain is reduced if the
anisotropy is large and the height of the embankment is signifcant. A chimney drain, or other types
of drains that intercept horizontal seepage planes, may be more efective. Section 6.6.2 presents guid-
ance for seepage analyses for the design of internal drains, including the determination of drain
dimensions.
At many coarse refuse embankment dams, the fne coal refuse deposited in the impoundment, cre-
ates a delta or beach on the upstream slope that typically restricts seepage, provided that the
normal pool is maintained upstream of the delta. Economies in internal drain construction can be
achieved by evaluating the hydraulic conductivity characteristics of the fne refuse material, as well
as available borrow material, and using zoned embankment design. For homogeneous embankment
dams used for fresh water supply, some savings can also be gained by using outlet drains for dis-
charging the water from the base of a chimney drain, which will eliminate the need for a granular
blanket extending beneath the entire downstream portion of the embankment. This alternative is
illustrated in Figure 6.1d.
6.3.1.2 Zoned Embankments
Coarse coal refuse and mine spoil are generally not sufciently fne grained to keep seepage at a low
level. If the design of an embankment requires that seepage be minimized or it is desired to lower
the saturation level in the downstream portion of the embankment, a less pervious zone within the
< PREVIOUS VIEW
6-22
Chapter 6
MAY 2009
COMPACTED COARSE
COAL REFUSE
DRAINAGE BLANKET
(FILTERS NOT SHOWN)
6.1a VERTICAL OR STEEPLY SLOPING CHIMNEY DRAIN
FIGURE 6.1 DRAIN CONFIGURATIONS FOR HOMOGENEOUS EMBANKMENTS
WITH IMPOUNDMENTS
SETTLED FINE
COAL REFUSE
6.1b HORIZONTAL BLANKET DRAIN
COMPACTED COARSE
COAL REFUSE
DRAINAGE BLANKET
(FILTERS NOT SHOWN)
SETTLED FINE
COAL REFUSE
6.1d CHIMNEY DRAIN WITH OUTLET DRAIN EXITS
PERVIOUS DRAIN OUTLET
DISCHARGING TO DOWN-
STREAM TOE (FILTERS NOT
SHOWN)
A
A
SECTION A - A
DRAINAGE
MATERIAL
PERFORATED
PIPE
COMPACTED COARSE
COAL REFUSE
FLOOD POOL LEVEL
NORMAL POOL
LEVEL
SETTLED FINE
COAL REFUSE
FLOOD POOL LEVEL
FLOOD POOL LEVEL
6.1c DISCRETE INTERNAL DRAIN
DISCRETE INTERNAL DRAINS
(FILTERS NOT SHOWN)
COMPACTED COARSE
COAL REFUSE
A
A
SETTLED FINE
COAL REFUSE
FLOOD POOL LEVEL
NORMAL POOL
LEVEL
NORMAL POOL
LEVEL
NORMAL POOL
LEVEL
PHREATIC
SURFACE
PHREATIC
SURFACE
PHREATIC
SURFACE
PHREATIC
SURFACE
FIGURE 6.1 DRAIN CONFIGURATIONS FOR HOMOGENEOUS EMBANKMENTS
WITH IMPOUNDMENTS
FIGURE 6.1 DRAIN CONFIGURATIONS FOR HOMOGENEOUS EMBANKMENTS
WITH IMPOUNDMENTS
< PREVIOUS VIEW
6-23
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
TABLE 6.12 ADVANTAGES AND DISADVANTAGES OF INTERNAL DRAIN TYPES
Internal Drain Type Figure No. Advantage(s) Disadvantage(s)
Steeply Sloping
Chimney Drain
6.1a
Positive seepage interception
and collection system
Expensive and diffcult to construct;
requires careful planning and stringent
construction control to connect with
future stages.
Horizontal Blanket
Drain
6.1b Simple construction
Ineffective in high anisotropy
conditions.
Discrete Internal
Drain
6.1c
Relatively inexpensive
and independent of future
embankment raising
Partial seepage interruption;
effectiveness depends on anisotropy.
Chimney Drain and
Outlet Sections
6.1d Positive seepage interception Expensive and diffcult to construct.
embankment may be needed. A zoned embankment consists of multiple material zones, generally
including an impervious (or low-hydraulic-conductivity, fne-grained material) core or upstream
zone of limited width and additional zones of coarse material that provide strength and erosion resis-
tance. Employing such a zoned embankment concept can reduce material requirements for internal
drainage structures and structural embankment zones.
At some coal refuse disposal sites, two zones are incorporated into the embankment cross section a
relatively impervious soil zone in the upstream section and coarse material such as mine spoil, granu-
lar borrow material or coarse coal refuse in the downstream section. If controlling the volume of seep-
age is not of primary importance, zoned embankments can be designed with the object of lowering
the phreatic surface in the downstream face of the embankment. The primary design consideration in
such cases is that the hydraulic conductivity of the downstream zone be sufciently high to discharge
the water seeping through the core or upstream zone without an elevated phreatic surface.
Figure 6.2 illustrates confgurations of zoned embankments most commonly used in dam construc-
tion. Figures 6.2a and 6.2b show a vertical core and a sloping core, respectively. Figure 6.2c illustrates a
zoned embankment consisting of fner soil in upstream portion of the dam and coarse material in the
downstream portion. An upstream zone may be preferred for slurry impoundments, provided erosion
from pool-level fuctuations or runof is not signifcant or can be controlled. Alternately, zoned coarse
refuse dams, in which the upstream zone receives more compactive efort to increase material break-
down and to lower hydraulic conductivity, may be specifed. However, many coarse refuse dams are
not zoned, taking advantage of the fne coal refuse deposited in the impoundment that may be as efec-
tive as zoning in limiting seepage, provided that: (1) the impoundment pool is maintained at a low level
and does not surcharge the upstream embankment face and (2) the response of the phreatic surface to
increased pool levels as a result of storm runof will not compromise embankment stability.
A description of important features related to impervious zone and shell design for zoned embank-
ments is provided in the following text.
Impervious Zone Design
Although small amounts of seepage are typically present in the low-hydraulic-conductivity portion
of an embankment, these embankment components are generally referred to as impervious zones.
For design of impervious zones, the primary considerations are the type of impervious zone and the
thickness and material used for their construction. The upstream zone of a zoned embankment may
serve a similar purpose, particularly for coal refuse impoundments.
< PREVIOUS VIEW
6-24
Chapter 6
MAY 2009
Impervious Zone Type
Both vertical and sloping impervious zones have advantages, as described in Table 6.13. The selec-
tion of the type of zoning of the material must be determined on an individual site basis. Valuable
additional discussion of the design of embankments with impervious zones is provided in USBR
(1992a).
PERVIOUS FOUNDATION ZONE PERVIOUS FOUNDATION ZONE
TRANSITION
ZONE (IF NECESSARY)
6.2a VERTICAL CORE
FILTER ZONES
(IF NECESSARY)
IMPERVIOUS CORE
FIGURE 6.2 ZONED EMBANKMENTS WITH IMPOUNDMENTS
FLOOD POOL LEVEL
NORMAL POOL
LEVEL
IMPERVIOUS FOUNDATION MATERIAL
PERVIOUS FOUNDATION ZONE
6.2b SLOPING CORE
IMPERVIOUS CORE
TRANSITION ZONE
(IF NECESSARY)
FLOOD POOL LEVEL
FILTER ZONES
(IF NECESSARY)
NORMAL POOL
LEVEL
IMPERVIOUS FOUNDATION MATERIAL
6.2c UPSTREAM IMPERVIOUS ZONE
FILTER/TRANSITION ZONE
(IF NECESSARY)
UPSTREAM IMPERVIOUS ZONE
PERVIOUS FOUNDATION ZONE
IMPERVIOUS FOUNDATION MATERIAL
FLOOD POOL LEVEL
NORMAL POOL
LEVEL
SETTLED FINE
COAL REFUSE
EMBANKMENT MATERIAL
(INTERNAL DRAINAGE NOT
SHOWN)
EMBANKMENT MATERIAL
(INTERNAL DRAINAGE NOT
SHOWN)
EMBANKMENT MATERIAL
(INTERNAL DRAINAGE NOT
SHOWN)
FIGURE 6.2 ZONED EMBANKMENTS WITH IMPOUNDMENTS
FIGURE 6.2 ZONED EMBANKMENTS WITH IMPOUNDMENTS
< PREVIOUS VIEW
6-25
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
Impervious Zone Thickness
The impervious zone thickness is normally governed by practicalities (Sherard et al, 1963), including:
(1) tolerable seepage volume, (2) thickness that will permit proper construction, (3) type, quality and
cost of available low-hydraulic-conductivity material, (4) type, quality and cost of available material
for any flter layers between the impervious zone and the adjoining downstream soil, and (5) the
quantity and quality of available soil/rock for the shell. While subsequent research and testing has
provided refnement in the design of impervious zone thickness, the following criteria developed for
water retention dams (USBR, 1992a; McCook, 2002) may serve as preliminary guidelines for accept-
able impervious zone thickness for other impounding embankments, recognizing that other dimen-
sions/confgurations may be suitable pending seepage and stability analyses:
Cores with a thickness greater than 30 percent of the depth of water head have
proven satisfactory for many dams under diverse conditions; for a starter dam, a
core of this thickness will probably be adequate for many types of impervious core
material and embankment heights.
Cores with a thickness of 15 to 20 percent of the depth of water head are considered
thin, but if adequately designed and constructed flter layers are used, they will
probably be satisfactory under most circumstances.
Cores with a thickness of less than 10 percent of the depth of water head should be
considered only in circumstances where a large leak through the core would not lead
to embankment failure or unacceptable environmental conditions; very carefully
designed and constructed flter layers should be considered.
Impervious zones of limited width, governed by material properties and availability as well as con-
struction practicalities, have proven advantageous in limiting seepage at starter dams and slurry
TABLE 6.13 RELATIVE ADVANTAGES OF VERTICAL AND SLOPING EMBANKMENT CORES
Vertical Core Sloping Core
1. Higher confnement pressure, more uniform settlement
with increased embankment load, and shear strength
is not as critical as with a sloping core.
2. Higher pressure is present between the impervious
core and the foundation, providing additional pro-
tection against leakage along the contact surface.
3. For a given volume of impervious soil material, the
thickness of the vertical core will be slightly greater
than that of sloping core.
4. If it becomes necessary to grout the foundation after
the embankment has been raised to a signifcant
height, the grouting can be con ducted from the crest
of the embankment through the core and directly
into the foundation, as opposed to a sloping core,
where the tie between the core and the foundation is
beneath the impoundment.
5. A vertical core is not subject to damage from sloughing
or erosion of the upstream toe.
1. The main downstream portion of the embank ment can
be constructed frst and the imper vious core placed
later without disrupting construction operations. This
advantage allows the downstream portion of the
embankment to be constructed year round, even if
controlled construction of the core can be done only
during short, good-weather periods.
2. Filter layers between the core and the upstream and
downstream portions of the embankment can be
made thinner and constructed more easily.
3. The core construction can be staged easily if the
embankment is expanded by the down stream staged
construction method.
< PREVIOUS VIEW
6-26
Chapter 6
MAY 2009
impoundments. As discussed above, the deposition of fne refuse slurry may also be efective at
limiting seepage provided that: (1) the impoundment pool is maintained at a low level and (2) the
response of the phreatic surface to increased pool levels as a result of storm runof will not compro-
mise embankment stability.
Impervious Zone Material
Impervious zone material is usually selected based on specifc requirements for controlling seep-
age and the availability of suitable material at the site. Typically, fne-grained soil is used for such
construction. The potential for failure resulting from loss of impervious zone material and leakage
caused by cracking or diferential slippage within the impervious zone will infuence the design,
the materials used and the construction procedure. The possibility of excessive leakage due to
cracking is a particularly important consideration for embankments on sof foundation material,
in areas susceptible to subsidence, or in regions of high seismic activity. Britle soil behavior and
cracking problems ofen can be minimized by placing the impervious zone material at a higher
than optimum moisture content. Figure 6.3 provides a classifcation of materials according to
resistance to piping and cracking. If foundation setlements are expected to be high, a suitable
internal drainage layer should be placed immediately downstream of the impervious zone to con-
trol seepage resulting from possible cracking.
Dispersive clay soils have a preponderance of sodium cations in their pore water in contrast to most
clays, which have a preponderance of calcium and magnesium cations in their pore water. A hole
through a dispersive clay will increase in size as water fows through (due to the breakdown of
the soil structure), whereas a hole in a non-dispersive clay will remain essentially constant in size.
Dispersive clays should not be used in dam construction because they are extremely susceptible to
piping. The crumb test (ASTM D 6572) can be conducted in the feld or laboratory and may indicate
if soils are dispersive. The dispersion potential can most reliably be determined using the pinhole
test (ASTM D 4647). The dispersion potential of clay for several ranges of measured dispersion is
provided in Table 6.14 (Sherard et al., 1976).
Design of flters for impervious soils used for the core can be critical for the downstream interface
with the shell, but is generally less critical for the upstream interface for a coal refuse impoundment
because reservoir fuctuations are minimal. Sherard et al. (1985) and McCook (2002) address flter
criteria and hydraulic gradient issues associated with impervious soils in dams; this topic is further
discussed in Section 6.6.2.
Shell Design
Embankment shell zone material typically is selected from granular material at the site. Coarse coal
refuse, mine spoil and rock excavated from the construction of water drainage systems or haul roads
are generally suitable for shell construction. The earthfll portion (upstream zone or core) may be
zoned or protected by graded flter zones, as discussed in Section 6.6. The use of rockfll is governed
by economy and structural stability in addition to slope protection. If a plentiful supply of suitable
rock is available at low cost, it is ofen possible to steepen the slopes of the adjoining earthfll portion
by providing additional free-draining rock on the downstream face, resulting in added stability. With
careful planning and design, the rockfll section of the starter dam can be utilized as a toe drain for
future expansion of the facility by the upstream method of construction. However, this may pose a
design constraint if the expansion is by the downstream method.
6.3.1.3 Foundation Seepage Control
A very important consideration in the design of an impounding embankment is the control of seep-
age through the underlying foundation materials. Unlike a water storage reservoir, there is litle need
to retain water in a slurry disposal facility. Therefore, minimizing seepage through the foundation
< PREVIOUS VIEW
6-27
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
FIGURE 6.3 RESISTANCE OF CORE MATERIALS TO PIPING AND CRACKING
100 8 6 4 3 2 10 8 6 4 2 1 3
10
20
0
30
40
50
60
70
80
90
100
8 6 4 3 2 0.1 8 6 4 3 2 0.01 8 6 4 3 2 0.001
P
E
R
C
E
N
T
F
I
N
E
R
B
Y
W
E
I
G
H
T
4
"
3
/
8
"
1
/
2
"
3
/
4
"
1
"
3
"
2
1
/
2
"
2
"
1
1
/
2
"
2
0
0
1
0
0
5
0
7
0
4
0
3
0
1
6
81
0
4
CLEAR SQUARE OPENINGS U.S. STANDARD SIEVE NUMBERS HYDROMETER ANALYSIS
GRAIN SIZE (mm)
COBBLES
GRAVEL SAND
SILT OR CLAY (PLASTIC OR NON-PLASTIC)
FINE MEDIUM COARSE FINE COARSE
A LEAST CHANCE
FOR CRACKING
C
LITTLE CHANCE
FOR CRACKING
VULNERABLE TO
CRACKING
PIPING
RESISTANCE 1
PIPING RESISTANCE 2
3
B
TYPICAL GRADATION ASSOCIATED
WITH PIPING RESISTANCE
CATEGORY MATERIAL CHARACTERISTICS
PIPING
RESISTANCE
CL AND CH WITH PI>15, WELL
GRADED SC WITH PI>15.
GREATEST RESISTANCE TO PIPING. SMALL AND MEDIUM CONCENTRATED LEAKS WILL HEAL THEMSELVES.
EMBANKMENT MAY FAIL AS A RESULT OF SLOWLY PROGRESSIVE PIPING CAUSED BY LEAK OF ABOUT ONE-
HALF CFS.
1
PIPING
RESISTANCE
INTERMEDIATE RESISTANCE TO PIPING. SAFELY RESISTS SATURATION OF LOWER PORTION OF DOWN-
STREAM SLOPE INDEFINITELY. MAY FAIL EVENTUALLY AS A RESULT OF EROSION CAUSED BY A SMALL
CONCENTRATED LEAK OR BY PROGRESSIVE SLOUGHING. IF A LARGE LEAK DEVELOPS, PIPING CAUSES
FAILURE IN A SHORT TIME.
CL AND ML WITH PI<15, WELL
GRADED SC AND GC WITH
7<PI<15.
2
PIPING
RESISTANCE 3
SP AND UNIFORM SM AND ML
WITH PI<7.
LEAST RESISTANCE TO PIPING. USUALLY FAILS IN A FEW YEARS AFTER FIRST RESERVOIR FILLING IF
SEEPAGE IS ABLE TO BREAK OUT ON DOWNSTREAM SLOPE. SMALL CONCENTRATED LEAK ON DOWN-
STREAM SLOPE CAN CAUSE FAILURE IN A SHORT PERIOD OF TIME. HIGH DENSITY FROM COMPACTION
INCREASES RESISTANCE SIGNIFICANTLY.
CRACKING
RESISTANCE A
CH WITH D50<0.02MM AND
PI>20.
HIGH POST-CONSTRUCTION SETTLEMENT, PARTICULARLY IF COMPACTED DRY. HAS SUFFICIENT DEFORM-
ABILITY TO UNDERGO LARGE SHEAR STRAINS FROM DIFFERENTIAL SETTLEMENT WITHOUT CRACKING.
CRACKING
RESISTANCE B
GC, SC, SM, SP WITH
D50>0.15MM.
SMALL POST-CONSTRUCTION SETTLEMENT. LITTLE CHANCE FOR CRACKING UNLESS POORLY COM-
PACTED AND LARGE SETTLEMENT IS IMPOSED ON EMBANKMENT BY CONSOLIDATION OF THE
FOUNDATION.
CRACKING
RESISTANCE C
CL, ML AND SM WITH PI<20,
0.15MM>D50>0.02MM.
MEDIUM TO HIGH POST-CONSTRUCTION SETTLEMENT AND VULNERABLE TO CRACKING. SHOULD BE
COMPACTED AS WET AS POSSIBLE CONSISTENT WITH STRENGTH REQUIREMENTS.
FIGURE 6.3 RESISTANCE OF CORE MATERIALS TO PIPING AND CRACKING
(DOD, 2005)
is not as critical to the design as it is for storage reservoirs. Where seepage occurs, it must be con-
trolled such that it does not adversely afect the safety of the embankment or result in environmental
impacts. The foundation conditions likely to be encountered beneath the starter dam are: (1) pervious
foundation, (2) impervious foundation, or (3) impervious stratum at the surface underlain by a pervi-
ous stratum. An additional foundation seepage concern at some coal refuse facilities is the potential
for fracturing due to subsidence of the ground surface above underlying mines.
Pervious foundations may consist of boulders, gravels, sands or mixtures thereof. For such foun-
dations, measures to minimize seepage quantity and to provide controlled seepage discharge are
FIGURE 6.3 RESISTANCE OF CORE MATERIALS TO PIPING AND CRACKING
< PREVIOUS VIEW
6-28
Chapter 6
MAY 2009
normally required. Control measures may include low-hydraulic conductivity barriers (e.g., cutof
trench and backfll) to decrease or virtually stop seepage, or a collection system can be provided
beneath the downstream portion of the embankment to control the discharge of seepage. If seepage
control and/or collection systems are not intended to be employed, the safety (including the potential
for piping of foundation or embankment materials) and environmental ramifcations should be care-
fully evaluated, and suitable measures should be employed to monitor pore pressures, if necessary.
The most common methods used for controlling foundation seepage are construction of a low-hydrau-
lic-conductivity cutof through the pervious foundation material and construction of an impervious
blanket extending far enough upstream to sufciently restrict the fow. In some situations, construc-
tion of an impermeable liner beneath the impoundment may also be used to address foundation seep-
age control. These methods are illustrated in Figure 6.4. Normally, if the pervious foundation material
is thin and excessive groundwater problems due to excavating in the valley botom are not anticipated,
the low-hydraulic-conductivity cutof is the least expensive method. Important considerations in the
design and construction of seepage cutofs are presented by Sherard et al. (1963) and USBR (1987a,
1992a). Procedures for designing an impervious blanket are presented by Cedergren (1989).
Where seepage through the foundation is allowed to occur, a collection system is almost always pro-
vided. The two major negative efects of allowing seepage to occur are: (1) a decrease in the factor
of safety against instability of the embankment due to high pore pressures in the foundation and
(2) the potential for piping in the foundation. If the foundation soil is not stratifed in the horizontal
direction, seepage control can be provided by a horizontal blanket drain (Figure 6.1a). This method
normally requires analysis of the path of the seepage and assurance during placement of the blanket
that it is directly tied to underlying pervious material. Seepage is collected at the downstream end of
the blanket and discharged at a predetermined location near the valley botom.
Other methods of controlling seepage through a pervious foundation include a deep drainage
trench constructed near the toe of the embankment and a relief well system, as discussed in greater
detail in Section 6.6.2.3.4. Measures associated with the design and management of the impound-
ment and clarifed water level can also aid in controlling seepage. In some cases, impoundment
cells for clarifed water are developed in the upstream portion of the impoundment. The deposi-
tion of fne refuse and resulting longer seepage path aids in restricting seepage beneath the down-
stream embankment.
Impervious foundations typically consist of massive rock and predominantly clayey soils. When
an impoundment is to be constructed upon impervious foundation materials, seepage beneath the
embankment is not a major design consideration if a proper seal is placed between the embankment
and the foundation. Methods of foundation preparation that efectively create a seal and references to
supplemental technical publications on grouting and rock preparation are presented in Chapter 11.
TABLE 6.14 CLAY DISPERSION POTENTIAL
Percent Dispersion
(1)
Dispersive Tendency
Over 40 Highly Dispersive (do not use)
15 to 40 Moderately Dispersive
0 to 15 Resistant to Dispersion
Note: 1. The ratio between the fraction fner than 0.005 mm in a soil-water suspension that
has been subjected to a minimum of mechanical agitation and the total fraction fner
than 0.005 mm determined from a regular hydrometer test times 100.
(SHERARD ET AL., 1976)
< PREVIOUS VIEW
6-29
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
FIGURE 6.4 SEEPAGE CUTOFFS FOR PERVIOUS FOUNDATIONS
FIGURE 6.4 SEEPAGE CUTOFFS FOR PERVIOUS FOUNDATIONS
6.4a PERVIOUS FOUNDATION CUTOFF WITH ZONED EMBANKMENT
COMPACTED COARSE
COAL REFUSE
FILTER/TRANSITION ZONE
(IF NECESSARY)
PERVIOUS FOUNDATION ZONE
UPSTREAM IMPERVIOUS ZONE
FLOOD POOL LEVEL
NORMAL POOL
LEVEL
SETTLED FINE
COAL REFUSE
IMPERVIOUS FOUNDATION MATERIAL
PERVIOUS FOUNDATION ZONE
IMPERVIOUS FOUNDATION MATERIAL
6.4b IMPERVIOUS BLANKET
COMPACTED COARSE
COAL REFUSE
IMPERVIOUS CORE
FILTER ZONES
(IF NECESSARY)
PARTIAL IMPERVIOUS BLANKET
SETTLED FINE
COAL REFUSE
FLOOD POOL LEVEL
NORMAL POOL
LEVEL
TRANSITION ZONE
(IF NECESSARY)
NOTE: PARTIAL IMPERVIOUS BLANKET EXTENDS UPSTREAM
SUFFICIENTLY FAR TO CONTROL SEEPAGE.
6.4c IMPERVIOUS LINER WITH HOMOGENEOUS EMBANKMENT
COMPACTED COARSE
COAL REFUSE
PERVIOUS FOUNDATION ZONE
FLOOD POOL LEVEL
NORMAL POOL
LEVEL
SETTLED FINE
COAL REFUSE
LINER SYSTEM
PHREATIC
SURFACE
IMPERVIOUS FOUNDATION MATERIAL
DISCRETE INTERNAL
DRAINS
FIGURE 6.4 SEEPAGE CUTOFFS FOR PERVIOUS FOUNDATIONS
< PREVIOUS VIEW
6-30
Chapter 6
MAY 2009
For disposal facilities where impoundments are constructed upon an impervious foundation, the
National Coal Board (1970 and 1972) has suggested an efective concept for allowing controlled
drainage from setled slurry. Granular drainage and flter material is placed beneath the impound-
ment area prior to flling. The drainage and flter material transition to pipes that pass under the
embankment to a downstream collection system or to a discharge outlet. In some cases, to assure
complete drainage, this zone is constructed over the upstream face of the embankment. DAppolonia
(1988) conducted OSM-sponsored research and implemented a design for an impoundment internal
drain structure to improve consolidation of the fne coal refuse and to intercept seepage before it
enters the coarse refuse embankment, thus mitigating acid mine drainage potential. This concept is
illustrated in Figure 6.5. Another method to improve drainage and consolidation of fne coal refuse
is installation of wick drains.
S
T
A
G
E
I
S
T
A
G
E
II
S
T
A
G
E
III
S
T
A
G
E
I
D
R
A
IN
P
IP
E
F
IN
E
C
O
A
L
R
E
F
U
S
E
CONDUIT SEEPAGE
INTERCEPTION
GRANULAR INTERNAL DRAIN
S
T
A
G
E
II
S
T
A
G
E
III
IN
T
E
R
N
A
L
D
R
A
IN
(T
Y
P.)
FIGURE 6.5 IMPOUNDMENT INTERNAL DRAIN CONCEPT
For the case of an impervious stratum at the surface overlying more pervious strata below, there is
potential for high pore-water pressure to occur downstream of the dam. This may cause blow outs,
boiling, piping or instability at or beyond the downstream toe. An upstream cutof trench, deep
drainage using relief wells or construction of berms should be considered in such cases.
6.3.2 Non-impounding Embankments
Non-impounding embankments require many of the same design and construction considerations
as impoundments. Careful geotechnical investigation can: (1) identify potentially unstable soils that
are incapable of sustaining embankment loadings and (2) decrease the probability of groundwater
impacts by identifying certain site characteristics that should either be avoided or recognized during
the design phase. Safety may become a major design factor when:
The disposal facility is located in areas where failure would have a high possibility of
taking lives and could seriously damage infrastructure or buildings.
The facility is located immediately above a stream and signifcant movement of the
embankment could block or restrict the fow, possibly creating a temporary impound-
ment that could release a food wave upon breaching of the sloughed material.
FIGURE 6.5 IMPOUNDMENT INTERNAL DRAIN CONCEPT
< PREVIOUS VIEW
6-31
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
The facility is constructed across a large valley with a signifcant watershed or it may
temporarily impound water during some stage of development, despite the presence
of drainage systems.
Embankments should not impede drainage, and cross-valley confgurations should be avoided
because they can retain water and be subject to classifcation by MSHA as an impoundment. MSHA
(2007) presents factors to be considered in this regard and also cautions that reclamation at abandon-
ment can require signifcant regrading to address drainage concerns.
The following discussion of important design considerations for non-impounding embankments
does not specifcally address cases where a high safety hazard may exist. The designer must rec-
ognize special safety hazards and formulate the investigation program and analyses accordingly.
To properly plan and design a non-impounding coal refuse embankment, the following questions
should normally be answered:
Considering the required volume and the disposal facility geometry, what is the
smallest quantity of constructed material that could form the critical downstream
structural portion of the embankment with assured stability?
If the embankment were to temporarily impound water, could it lead to pore-water
pressure conditions that are hazardous to the stability of the embankment?
Is an impervious liner system needed in order to limit seepage or infltration of the
groundwater?
Should an internal drainage system be placed under critical portions or all of the
disposal facility to control saturation for stability purposes or to collect seepage for
treatment before discharge?
Non-impounding coal refuse embankments typically accommodate disposal of coarse, combined,
and dewatered fne coal refuse. Well-graded coarse refuse may be generated without signifcant
excess moisture, such that internal drainage provisions are primarily directed at control of natural
springs or infltration. Consequently, seepage control and slope stability can readily be addressed
with minimal internal drainage requirements and normal placement and compaction of materi-
als. Combined refuse and dewatered fne refuse generally contain excess moisture when they are
generated, and greater measures are required for seepage control and slope stability. Such embank-
ments may be zoned to establish structural downstream zones for stability and may incorporate
internal drainage control structures, allowing for upstream zones where the material consistency
and strength are less important. However, applicable regulations for the construction of refuse
piles must be followed.
Several examples of designs for non-impounding embankments, including provisions for seepage
control, are illustrated in Figure 6.6. Figures 6.6a and 6.6b show drainage system concepts previously
discussed in Section 6.3.1, where seepage is controlled by internal drainage systems. The previous
discussion of internal drain materials and their gradation, and of the need for flters, is also applicable
to this type of construction. The full extent of seepage control and drainage systems must be deter-
mined on a project-by-project basis considering site conditions, the economics of various methods of
providing embankment stability, the cost of leachate treatment, and appropriate regulations.
For some disposal facilities, usually depending on site location and condition, seepage impacts on the
groundwater regime may need to be mitigated. In such cases, a relatively impervious liner composed
of fne-grained borrow materials should be constructed beneath the disposal facility. This can be accom-
plished as the facility is being constructed by rearranging and compacting native soils or by using borrow
soils, as illustrated in Figure 6.6c. At some sites, a geomembrane or geosynthetic clay liner (GCL) may
< PREVIOUS VIEW
6-32
Chapter 6
MAY 2009
6.6a BOTTOM INTERNAL DRAIN
6.6b BOTTOM AND DISCRETE INTERNAL DRAINS
6.6c BOTTOM INTERNAL DRAIN WITH LINER SYSTEM
FIGURE 6.6 NON-IMPOUNDING EMBANKMENTS
STRUCTURAL ZONE
(COARSE OR COMBINED REFUSE)
INTERNAL DRAIN
PLACED REFUSE ZONE
(COARSE OR COMBINED REFUSE)
STRUCTURAL ZONE
(COARSE OR COMBINED REFUSE)
INTERNAL DRAIN
PLACED REFUSE ZONE
(COMBINED OR DEWATERED
FINE REFUSE)
DISCRETE INTERNAL
DRAINS
STRUCTURAL ZONE
(COARSE OR COMBINED REFUSE)
INTERNAL DRAIN
PLACED REFUSE ZONE
(COARSE OR COMBINED REFUSE)
LINER SYSTEM
(MAY INCLUDE UNDERDRAIN)
FIGURE 6.6 NON-IMPOUNDING EMBANKMENTS
be used as an impervious liner if satisfactory borrow materials are not available. Selected portions of the
impervious zone may be covered by drainage materials to collect and transport leachates to a common
point for treatment and/or discharge. The introduction of liner materials may afect the stability of the
embankment and thus potentially impact the design slopes or confguration of a disposal facility.
Typically, groundwater seepage or mine discharge is collected and conveyed in a manner such that it
cannot enter the refuse. Figure 6.7a shows a seepage collector drain comprising a collection zone and a
FIGURE 6.6 NON-IMPOUNDING EMBANKMENTS
< PREVIOUS VIEW
6-33
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
FIGURE 6.7 EXAMPLE SPRING COLLECTION ZONES AND ROCK UNDERDRAIN
FIGURE 6.7 EXAMPLE SPRING COLLECTION ZONES AND ROCK UNDERDRAIN
E
X
T
E
N
T
O
F
S
E
E
P
A
G
E
Z
O
N
E
PROTECTIVE COVER
MATERIAL
FILTER SAND
G
R
A
N
U
L
A
R
D
R
A
IN
M
A
T
E
R
IA
L
EXTENDS TO
UNDERDRAIN
OR DITCH
GEOTEXTILE WRAP
6.7a SPRING COLLECTION ZONE WITHOUT LINER SYSTEM
6.7b SPRING COLLECTION ZONE AS PART OF LINER SYSTEM
EXTENT OF
SEEPAGE ZONE
FILTER SAND
LINER MATERIAL
PROTECTIVE COVER MATERIAL
GEOTEXTILE
WRAP
GEOTEXTILE
GRANULAR DRAIN MATERIAL
6.7c ROCK UNDERDRAIN AS PART OF LINER SYSTEM
LINER MATERIAL
PROTECTIVE COVER MATERIAL
FILTER SAND
GEOTEXTILE
GEOTEXTILE
WRAP
ROCK UNDERDRAIN
FIGURE 6.7 EXAMPLE SPRING COLLECTION ZONES AND ROCK UNDERDRAIN
< PREVIOUS VIEW
6-34
Chapter 6
MAY 2009
conveyance system. The collection zone consists of granular material covered with a granular flter layer
and/or geotextile. The collected water is discharged through the granular drain or pipe to the nearest
internal drainage system or beyond the downstream toe. The granular flter layer or geotextile situated
between the water collecting zone and the overlying refuse and perforated pipes incorporated into the
drainage system (if present) should be designed using procedures discussed in Section 6.6. Inclusion
of an impervious liner over the collection zone, as shown in Figure 6.7b, may be appropriate for pre-
venting embankment seepage from entering the collected groundwater and to minimize the volume of
poor-quality water that may require treatment. Rock underdrains may be placed in valley botoms to
collect outfows from springs. Figures 6.7b and 6.7c show examples of spring collection zones and rock
underdrains for a lined facility. Spring collection zones and rock underdrains should be designed for
compatibility with surrounding materials using the flter criteria presented in Section 6.6.2.
Underground mine voids may cause subsidence that can impact an overlying embankment, poten-
tially disrupting liners and internal drainage structures. Measures for evaluating and addressing
potential subsidence problems are discussed in Chapter 8.
6.3.3 Slurry Cell Embankments
Slurry cell embankments may be classifed as impounding embankments or non-impounding
embankments depending on the confguration and storage capacity of active, uncovered cells and
the potential for multiple cells to be involved in a failure. For facilities with signifcant coal refuse
production rates, it is difcult to keep the active cell capacity below the impoundment classifcation
limit, and thus the aim is to design the facility so that it can be classifed as having low hazard poten-
tial. Such a system requires a design and construction sequence that minimizes the volume of active
cells and provides for timely drainage, covering, and consolidation of completed cells so that the fne
coal refuse is not fowable.
For slurry cell systems that are designed and classifed as impoundments, the following geotechnical
considerations are applicable:
Seepage control An internal drainage system and foundation cutof system
designed to intercept seepage and prevent it from impacting the structural zone or
toe of the embankment and a foundation treatment and liner system if necessary
to address leachate migration from the disposal facility. Slurry cells have limited
water and slurry storage and thus represent less signifcant sources of seepage than
conventional impoundments for fne coal refuse slurry disposal.
Slope stability Static slope stability is maintained by a structural zone con-
structed from coarse coal refuse or borrow material with sufficient width to
effectively contain and isolate the slurry cells from affecting potential failure
surfaces. Seismic stability and deformations may be of less concern because
individual cells tend to be shallower deposits of fine coal refuse that are better
drained and consolidated by layers of coarse refuse or borrow materials. Slough-
ing and erosion considerations are essentially the same as for other refuse
embankments. Slurry cell disposal requires coarse refuse or borrow material for
the cell structures and covering of completed cells and for any structural zones
that may be part of a valley-fill configuration. Thus, to address slope stability,
slurry cell systems may require more coarse refuse or borrow material for struc-
tural elements and cell construction than more traditional types of impounding
embankments.
Drainage structures Structural foundations and excavation slopes for diversion
channels, principal and emergency spillways, conduits and other auxiliary struc-
tures associated with impoundments.
< PREVIOUS VIEW
6-35
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
Underground mines Stability, sealing of mine openings, and infltration into
underground mine workings. Slurry cell systems provide a means to mitigate break-
through impacts of an impoundment into underground mine workings.
Slurry cell embankments that are designed with limited impounding capacity for water, sediment
or slurry so that they do not meet the impoundment size criterion (as discussed in Section 3.4.1) are
less sensitive to the impacts of seepage on stability, provided that efective drainage measures are
incorporated into the cells. Such embankments may not need principal and emergency spillways and
conduits associated with impoundments, thus reducing geotechnical considerations.
6.3.4 Embankment Construction Staging
Embankments are raised in stages using one or more modifcations of three basic approaches: (1)
upstream method, (2) downstream method, or (3) centerline method. In fact, many disposal plans use
a combination of upstream and downstream methods for staging. Figures 6.8 and 6.9 show example
confgurations of embankments constructed using a combination of the upstream and downstream
methods. This section describes common procedures for staging the development of a new embank-
ment or extending the life of an existing embankment. In the previous two sections, general concepts
for controlling groundwater and seepage are discussed; they are repeated here only as needed for
explaining staging methods. The inclusion of drainage systems in the overall embankment and its
staged parts must be designed on a site-by-site basis.
6.3.4.1 Upstream Method
In the upstream method of construction, the crest of the embankment is shifed progressively
upstream from the starter dam, as shown in Figure 6.8. The upstream method has several advan-
tages and disadvantages from an operational and safety standpoint, but one important drawback
is the engineering and design requirements necessary to address structural performance under
earthquake loading conditions. While there has not been a reported failure of a fne coal refuse
impoundment due to earthquake loadings in the U.S., other similar tailings impoundments and
hydraulic-fll dams have failed during or following seismic activity. The engineering analyses pre-
sented in Section 6.6 and Chapter 7 employ current methods for evaluating material properties and
loadings for design.
FIGURE 6.8 UPSTREAM CONSTRUCTION FOLLOWING INITIAL DEVELOPMENT
FIGURE 6.8 UPSTREAM CONSTRUCTION METHOD FOLLOWING
INITIAL DEVELOPMENT
NOTE: TAKEN FROM ACTUAL STAGING FOR VALLEY FILL SITE IN PENNSYLVANIA. THE FIGURE SHOWS UPSTREAM
CONSTRUCTION METHOD (STAGES IV - VII) FOLLOWING INITIAL DEVELOPMENT (STAGES I - III).
STAGE I
STAGE II
STAGE III
STAGE IV
STAGE V
STAGE VI
STAGE VII
FOUNDATION SOILS
FINE COAL REFUSE
TYPICAL INTERNAL
DRAIN
CONSTRUCTION KEY
STARTER
EMBANKMENT
FIGURE 6.8 UPSTREAM CONSTRUCTION FOLLOWING INITIAL DEVELOPMENT
< PREVIOUS VIEW
6-36
Chapter 6
MAY 2009
The major advantages of the upstream method are:
The structural fll volume is minimized because part of the new embankment is con-
structed in stages on top of the existing embankment and part is constructed over the
deposited fne refuse slurry.
The downstream face of the constructed embankment is the fnal face of the com-
pleted embankment, and vegetation and other environmental control measures can
be performed on a permanent basis.
The operational requirements such as haul and access roads, culverts, diversion and
perimeter ditches can be constructed to serve the entire useful life of the facility.
The impoundment watershed area decreases with the progress of embankment con-
struction, requiring less volume for storm runof handling with time.
The starter dam, if properly designed, can provide support and become part of the
internal drainage control for the subsequent embankment development.
When fne coal refuse slurry is deposited in an impoundment, grain-size sorting and layering occurs
in both the horizontal and vertical planes during the depositional process. Peripheral discharge of
slurry, either by single-point discharge or by multiple discharges, results in the formation of a beach
around the discharge point. The coarser particles setle close to the discharge point and the fner par-
ticles concentrate in the upstream portion of the impoundment. The depositional process may also
result in the formation of horizontal layers of coarser and fner fractions of the refuse and afect the
engineering properties of the setled fne refuse.
The fne coal refuse particles setling from the slurry accumulate in a very loose state initially and
have a high void ratio and moisture content. These loose deposits consolidate with time as water
is expelled from the voids between the particles. Consolidation is infuenced by such factors as the
efective pressure of the overlying material, the hydraulic conductivity of the deposit and surround-
ing soil, the distance that water must travel to drain and the time during which vertical pressure is
applied. Thus, the consolidation of fne coal refuse deposits varies spatially within the impoundment.
Even afer consolidation under self weight for years, such deposits may have high void ratios and
high moisture contents, afecting the shear strength of the fne refuse. This should be borne in mind
by the designer, if the upstream method of construction is utilized for embankment raising.
The disadvantage of the upstream method is that construction occurs atop previously deposited,
unconsolidated tailings. Under static loading conditions, there is a limiting height to which mate-
rials can be placed without the risk of a shear failure. In the initial placement of materials for the
embankment stage, normal spreading and compaction procedures cannot be followed until a frm or
stabilized base is prepared. Safety of operating personnel during the initial placement of material for
an upstream construction stage is very important. Additionally, the height and confguration of an
upstream constructed stage will depend on the strength of the material within the zone of shearing,
the downstream slope of the embankment, and the location of the phreatic surface within the facil-
ity. Under earthquake loading, this type of embankment may be susceptible to failure by strength
degradation. However, a stable impoundment can be constructed by following the basic principles
of embankment design and through judicious handling of material. A disposal facility constructed
using the upstream method must be designed by an engineer experienced in the behavior of soils,
coal refuse materials and embankments. Important considerations include:
The width of the embankment stages or structural fll zones must be adequate to
provide downstream slope stability. Upstream slope stability must be demonstrated
by adequate factors of safety for failure surfaces that could compromise the crest of
the embankment.
< PREVIOUS VIEW
6-37
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
As the embankment is raised, coarse refuse should be placed over setled slurry
according to an established schedule, considering the potential for pore pressure
buildup and dissipation in the saturated material resulting from the applied loads.
To the extent practical, the initial push out for upstream stages should not be placed
on submerged fne coal refuse.
Procedures for placement of materials for the embankment stage should address
equipment and operator activity and safety during the initial push-out onto the
setled fne coal refuse.
Reasonable estimates of movement and diferential setlements should be used in
designing the portion of the embankment constructed over setled slurry.
The constructed structural portion of each stage of the disposal facility should
include a system for controlling seepage, as discussed in Section 6.3.1.
If the starter dam is intended to become the toe portion of the overall embankment, the
downstream zone should be constructed of well-compacted, free-draining material so
that it will facilitate the dissipation of pore pressures within the overall embankment.
In disposal facilities retaining fne refuse slurry, the slurry discharge points should be
located adjacent to the embankment and above the pool level. To the extent practical,
the discharge point should be periodically moved along the length of the embank-
ment. This will concentrate the coarsest slurry particles adjacent to the embankment,
ofering the advantage of their greater strength and reducing setlement when the
next embankment stage is constructed.
Analyses of embankment stability should consider dynamic loads and potential for
strength degradation. Seismic design and deformation analyses are discussed in
Chapter 7.
Exploration, testing, engineering analyses and regulatory review associated with addressing the
above considerations are typically more complex and lengthy as compared to designs that employ
the downstream method.
6.3.4.2 Downstream Method
In the downstream method of construction, the crest of the embankment is shifed progressively
downstream from the starter dam, as designed for the early stages shown in Figure 6.9. This
method can be used for embankments with or without impoundments. The major advantages of
the method are:
The embankment is not built on hydraulically-deposited refuse.
Placement and compaction control can be exercised as required over the entire fll
operation.
The embankment can generally be raised above its initial design height without seri-
ous limitations and complicated design modifcations.
Internal drainage systems can be installed, as required, as the construction
progresses.
The embankment can be designed with minimal concern for strength loss of the fne
coal refuse under earthquake loadings. This can result in less complex and lengthy
exploration, testing, engineering analyses, and regulatory review.
The major disadvantage of the downstream method is that it requires relatively large volumes of fll
for raising the embankment. In the early stages of construction, it may not be possible for the mine to
< PREVIOUS VIEW
6-38
Chapter 6
MAY 2009
produce a sufcient volume of coarse refuse to maintain the crest of the embankment above the level
required for disposal of slurry and routing the design storm. Features such as haul and access roads,
culverts, benches, guters, diversion ditches and perimeter ditches may have to be reconstructed at
each stage, resulting in higher cost. Also, the fnal external face of the embankment is not created
until the construction of last stage of disposal, which typically ranges from 5 to 25 years afer the
start of construction. Consequently, interim faces are exposed to the weather prior to covering by the
next stage. These issues may have a considerable fnancial impact, if the facility has a relatively small
drainage area and a decant system is the primary spillway. However, when the construction mate-
rial is readily available, the downstream method normally results in simpler construction than the
upstream method and is also more easily implemented because of the many special requirements of
the upstream method.
The arrangement of the drainage and impervious zones within an embankment constructed using
the downstream method is usually similar to that of the embankments discussed in Section 6.3.1.1.
The zones can be placed in stages corresponding to the height of the embankment.
6.3.4.3 Centerline Method
The centerline method of embankment construction is essentially a variation of the downstream
method where the crest of the embankment is not shifed in the downstream direction, but instead
is raised vertically above the crest of the starter dam. A major advantage of this method is that the
downstream portion of the embankment is built on a frm foundation, and therefore placement and
compaction control can be exercised as required over that portion of the embankment. An important
consideration in using this method is to maintain an adequate width of structural fll in order to
FIGURE 6.9 DOWNSTREAM EMBANKMENT CONSTRUCTION FOLLOWED BY
EXPANSION USING DOWNSTREAM AND UPSTREAM METHODS
6.9a EMBANKMENT STAGES I - V CONSTRUCTED USING DOWNSTREAM METHOD
6.9b EMBANKMENT STAGES VI - XI-B CONSTRUCTED USING DOWNSTREAM
AND UPSTREAM METHODS
TYPICAL INTERNAL
DRAIN
STAGE V
STAGE IV
STAGE III
STARTER EMBANKMENT
STAGE II
IMPERVIOUS
ZONE
FINE COAL REFUSE
FOUNDATION SOILS
STAGE I
INTERNAL DRAIN
STAGES I - V
STAGE VI-A
STAGE VII-A
STAGE VII-B
STAGE VIII-A STAGE VIII-B
STAGE IX
STAGE X
STAGE XI-B
STAGE XI-A
FINE COAL REFUSE
LOWER INTERNAL
DRAIN
FOUNDATION SOILS
STAGE VI-B
FIGURE 6.9 DOWNSTREAM EMBANKMENT CONSTRUCTION FOLLOWED BY EXPANSION
USING THE DOWNSTREAM AND UPSTREAM METHODS
FIGURE 6.9 DOWNSTREAM EMBANKMENT CONSTRUCTION FOLLOWED BY EXPANSION
USING THE DOWNSTREAM AND UPSTREAM METHODS
< PREVIOUS VIEW
6-39
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
achieve stability. Other design considerations, as discussed in Section 6.3.4.1 for upstream construc-
tion, are also applicable to the centerline method. The centerline method also has all the disadvan-
tages listed in Section 6.3.4.2 for downstream method of construction.
The arrangement of the drainage and impervious zones within an embankment constructed using
the centerline method is usually similar to that of the embankments discussed in Section 6.3.1.1. The
zones can be placed in stages corresponding to the height of the embankment. The centerline method
poses many design, construction, environmental and operational problems, and thus is not generally
a preferred method of refuse disposal in the coal industry. This method is primarily used in tailings
dam construction where cyclones are used for separating the coarser fraction of the tailings. In the
coal industry, a combination of the upstream and downstream methods is typically employed to
meet storm water and slope stability criteria and material requirements.
6.3.4.4 Embankments Supplemented with Borrow Material
When the production of coarse refuse is not sufcient to construct the planned height of an
impounding embankment, or in other situations dictated by mining conditions, borrow material
or mine spoil may be used to supplement the planned construction. The embankment can be con-
structed from borrow material or mine spoil using the upstream or downstream methods, or these
materials can be used in combination with coarse refuse to meet the specifed gradation in a zoned
embankment, provided that specifc zones are designated for such materials. The guidance for
seepage control and drainage systems discussed in this Manual is appropriate for construction of
embankments of this type.
6.3.5 Special Considerations for Existing Embankments
The engineering principles and procedures for analyzing and designing modifcations to existing
refuse embankments are the same as those for new facilities. If an existing embankment is not per-
forming adequately, is not consistent with current engineering practice, or involves re-activation of
an older site, the following issues may need to be considered:
Current or additional loadings that may impact stability (static and/or seismic load-
ing conditions).
Potential for liquefaction and foundation failure of the embankment if constructed
over fne refuse.
Potential for piping and/or excessive uncontrolled seepage discharge from the
embankment or around the decant discharge pipe.
Potential safety problems to miners in nearby underground active workings due to
overburden breakthrough.
Potential for excessive erosion along sloped surfaces with inadequate drainage con-
trol and vegetative cover.
For impounding embankments, an inadequate combination of runof storage and
discharge capacity to accommodate the design storm.
Specifc conditions that need to be evaluated relative to the modifcation of existing embankments
include:
Material Placement The structural portions of embankment should be evaluated
relative to shear strength and seepage characteristics.
Foundation Preparation The foundation preparation work undertaken prior to con-
struction of the embankment and the potential for sliding along the embankment-
foundation interface zone should be evaluated.
< PREVIOUS VIEW
6-40
Chapter 6
MAY 2009
Foundation Condition It should be determined whether the embankment overlies
sof natural soils or setled fne refuse that must be considered in static and seismic
stability analyses.
Piping Potential It should be determined if existing seepage could lead to a piping
failure.
Underground Mining It should be determined if the facility overlies abandoned
or active mines, particularly ones with openings previously covered by coal refuse
that could afect the safety of miners or cause environmental or property damage
downstream.
Drainage Facility It should be determined if the existing drainage facilities will be
structurally safe for expanded operation and will meet design-storm criteria.
Site Boundary Constraints It should be determined if physical restrictions (e.g.,
streams, rivers, mines, utilities, roads, railroads) and property boundary restrictions
will limit the extent of any future modifcations.
Operations Capabilities Existing equipment and manpower capabilities or engi-
neering services (e.g., surveying or engineering support) should be evaluated for
adequacy to implement expansion or abandonment plans or facility modifcations
If an existing embankment exhibits unacceptable performance or is determined to be inadequate
against a stability failure, the following modifcations are appropriate:
Sliding or inadequate factor of safety against slope failure The slope can be stabi-
lized by: (1) constructing a downstream butress, (2) removing material from the top,
or (3) fatening the slope. Of these options, constructing a downstream butress is
most efective in stabilizing a slope.
Downstream Butress A downstream butress can be constructed using coarse
refuse from normal disposal operations, and borrowed granular soil can
be used for internal drainage control. Major limitations in constructing the
butress are physical or property limitations, poor access to the base of the
embankment, inadequate supply of coarse refuse to accomplish butress con-
struction in a reasonable period, and extremely poor foundation conditions in
the area of the butress requiring excavation. Inclusion of drains within the but-
tress section of the embankment can be very efective in improving the stability
of existing embankments. As illustrated in Figure 6.10a, a drain can be placed
as a granular blanket over the downstream face of an existing embankment
and then covered with coarse refuse to form a butress.
Removal of Material from Top Although this alternative improves slope sta-
bility, it is ofen not practical to remove material from the top of the slope
because of the resulting efect of the overall embankment confguration. Also,
the cost of removing the material and disposing it at a diferent location ofen
makes this alternative unatractive. This disadvantage can be partially over-
come if the excavated material can be used to construct a butress at the toe of
the slope, additionally enhancing stability.
Flatening of the Slope This alternative increases the stability of a non-
impounding facility, but it is ofen not suitable for an impounding embank-
ment constructed by the upstream method. The removal of material reduces
the thickness of the structural section of the embankment and may lower the
factor of safety. Also, the cost of removing material and disposing it at a dif-
ferent location ofen makes this alternative unatractive. This disadvantage
< PREVIOUS VIEW
6-41
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
can be partially overcome if the excavated material can be used to construct a
butress at the toe of slope.
Phreatic Level Reduction If high phreatic surface conditions in the embankment
contribute to an undesirable factor of safety, it is normally possible to lower the phre-
atic surface by: (1) installing horizontal drains and (2) reducing seepage.
Horizontal Drains One means for providing internal seepage control in an
existing embankment is by drilling horizontal drains beyond the phreatic sur-
face to intercept the seepage, as shown in Figure 6.10b. The design require-
ments and construction techniques for such drains are discussed in Section
6.6.2.3.5. The installation of horizontal drains is ofen very efective. How-
ever, the unknown and potentially variable nature of coal refuse introduces
greater risk that the drains will not meet expectations. If horizontal drains are
employed, concurrent monitoring of their efectiveness should be performed.
Extra drains should be added, or the existing drains should be supplemented
with other drainage improvement methods, if conditions warrant. Normally,
considerable monitoring of pore pressures conditions is required for evalua-
tion of drain system efectiveness.
Reducing Seepage It is normally possible to reduce the level of seepage by
sealing the surface of the upstream face of an impounding embankment.
The major disadvantage of this alternative is that the magnitude and rate
of improvement is difcult to predict. A wait-and-see approach can be
taken if the existing factor of safety is not critical on a short-term basis. If the
impoundment confguration permits, shifing of the fne refuse discharge
point to aid in sealing of the upstream face of the embankment and to force
the pool toward the back of the impoundment has been found to be efective
in reducing seepage.
Potential for Piping The most common measure for reducing the potential for
piping of embankment material in a seepage zone is to provide a granular flter
around the discharge location so that water can escape without carrying additional
fnes. Ofen this measure is coupled with construction of a butress with a flter zone
between the existing embankment and new butress material, as discussed above
and illustrated in Figure 6.10a. The way to minimize the potential for piping is to
prevent any past piping from extending into the new construction. Geophysical
exploration may be useful for identifying voids. If a void is present, the designer
must determine whether a flter system will be adequate and, if not, eliminate the
void by a technique such as grouting.
Erosion Control Erosion control can ofen be achieved through relatively minor
regrading and planting of vegetation or use of vegetative mats (erosion control
blankets). If an existing slope is too steep, regrading the surface to provide horizontal
benches can be efective.
6.3.6 Other Impounding Embankments
Other impounding embankments at mine sites include fresh water impoundments, sedimentation
ponds, and treatment ponds. The primary distinctions between these structures and slurry impound-
ments are: (1) the size of the embankment is usually smaller, with the width and height designed to
make efcient use of borrow material, typical of an earthen dam; (2) earthen borrow materials that
are not part of the on-going disposal operation are typically used for embankment construction; (3)
a permanent water pool level is maintained for fresh water impoundments and treatment ponds,
< PREVIOUS VIEW
6-42
Chapter 6
MAY 2009
resulting in steady state seepage conditions; (4) water pool level fuctuation within a sedimentation
pond is usually over a limited range, with normal levels maintained at a low level; and (5) without
the presence of fne coal refuse slurry, the pool level is typically in contact with the upstream face of
the dam and thus is a source of seepage through the embankment. These geotechnical design con-
siderations are important in the application of this Manual to projects with impoundments used for
other than the disposal of coal refuse slurry.
The following geotechnical considerations apply to fresh water impoundments, sedimentation ponds,
and treatment ponds:
Seepage control Internal drainage systems and foundation cutof systems to pre-
vent seepage from impacting the structural zone or toe of the embankment may be
needed. Foundation treatment and a liner system may be needed. Impoundments
that maintain a signifcant water pool depth impose greater hydraulic gradients
on the embankment, potentially requiring more seepage control. Internal drains
for these structures should have aggregate flters in order to comply with the
federal dam safety practices discussed in Section 6.6.2.3. Additionally, the initial
flling of such impoundments should be monitored closely for evidence of seepage
or other distress, as many dam incidents and failures have occurred under these
conditions.
Slope stability Static slope stability is maintained by a structural embankment
constructed of borrow material, with signifcant atention given to foundation condi-
tions. Seismic stability and deformations are most critical where loose foundation
conditions are present, as the embankment materials are usually designed to be well
FIGURE 6.10 INTERNAL DRAINAGE SYSTEMS FOR EXISTING IMPOUNDING EMBANKMENTS
6.10a BLANKET DRAIN ON EXISTING SLOPE
FIGURE 6.10 INTERNAL DRAINAGE SYSTEMS FOR EXISTING
IMPOUNDING EMBANKMENTS
EXISTING EMBANKMENT
BUTTRESS OF COMPACTED
COARSE COAL REFUSE OR
BORROW MATERIAL
BLANKET DRAIN
(FILTERS NOT SHOWN)
PHREATIC SURFACE
DRILLED HORIZONTAL
DRAINS WITH SLOTTED
PIPE
6.10b DRILLED HORIZONTAL DRAINS
EXISTING EMBANKMENT
ORIGINAL PHREATIC
SURFACE
LOWERED PHREATIC
SURFACE
FIGURE 6.10 INTERNAL DRAINAGE SYSTEMS FOR EXISTING IMPOUNDING EMBANKMENTS
< PREVIOUS VIEW
6-43
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
compacted and not subject to strength degradation. Sloughing and erosion consid-
erations may be more critical than encountered for slurry impoundments because
slopes may be steeper and the water pool level may fuctuate signifcantly. Fresh
water impoundments may impose stresses on the impounding embankment due to
rapid drawdown of the reservoir.
Drainage and outlet structures Structural foundations and excavation slopes
are required for diversion channels, principal and emergency spillways, conduits
and other auxiliary structures. Most fresh water impoundments incorporate outlet
structures that allow emergency drawdown of the reservoir. Control of seep-
age along conduits should be a point of emphasis. Many dam failures have been
caused by internal erosion due to excessive seepage through poorly compacted
backfll around conduits.
Underground mines Stability, sealing of mine openings, and infltration into under-
ground mine workings are concerns. Considering the hydraulic pressures imposed
by a fresh water dam, sites with shallow underground mine workings may not be
feasible or may require signifcant remedial measures.
Several design references are available for earthen dams used for fresh water, sedimentation con-
trol, and treatment: USBR (1987a); USBR (1992a); Bigatel et al. (1999); NRCS (2005b).
6.4 SITE GEOTECHNICAL/GEOLOGICAL EXPLORATION
Although the characteristics of foundation materials have been discussed in general terms, specifc
properties required for the design of a disposal facility can be determined only by conducting sur-
face and subsurface geological explorations at the site. The type and extent of explorations to be
conducted will depend upon the size of the planned embankment, the complexity of the site geology,
the nature of the foundation materials, the specifc function of the facility, and, most importantly,
whether or not the embankment will impound water. In general, a comprehensive site evaluation and
exploration program for a coal refuse disposal facility comprises the following tasks:
1. Review of available topographic maps, geologic soil survey and mine maps, satellite
imagery, and aerial photographs of the site and the surrounding area.
2. Surfcial geological and geotechnical reconnaissance of the site.
3. Identifcation of data needs for design, such as topographic maps, mine void loca-
tions, and soil, rock and refuse properties.
4. Preparation of a site investigation plan using compiled information to develop a site-
specifc strategy addressing exploration methods and locations, sampling and in-situ
testing requirements, and contingency activities based on anticipated and possible
unanticipated subsurface conditions.
5. Conducting a subsurface exploration program consisting of a combination of bor-
ings, test pits, in-situ testing and geophysical surveys.
6. Installation of monitoring systems such as piezometers, monitoring wells, and sur-
veying monuments, particularly for exploration at existing facilities.
7. Comparison of sample descriptions to anticipated conditions. Laboratory index test-
ing (i.e., particle-size distribution, Aterberg limits and moisture content) should be
performed on the samples to confrm visual classifcations.
8. Preparation of subsurface profles based upon results from the feld exploration and
laboratory index tests and review of these profles relative to the initial site investiga-
tion objectives and expectations.
< PREVIOUS VIEW
6-44
Chapter 6
MAY 2009
9. Selection of samples for performance testing and development of engineering prop-
erties for facility design from test results.
10. Laboratory performance testing and verifcation of the results using correlations
with the index test results. If performance test results are inconsistent with the
index test data, the inconsistency should be resolved (e.g., performance testing of
reserve samples or additional feld exploration to obtain replacement samples for
testing).
11. Interpretation of performance test results, comparison to anticipated conditions, and
selection of engineering properties needed for design.
12. Preparation of facility design, considering constructability issues, with identifcation
of construction-phase exploration and testing requirements to confrm critical perfor-
mance parameters.
The sequence and potentially iterative nature of these steps are summarized in the fow chart pre-
sented in Figure 6.11.
The preceding 12-step procedure is a suggested guideline for developing a thorough and cost-efec-
tive feld exploration and testing program. The efort required for each of the listed steps will vary.
Experience with similar geotechnical conditions will facilitate the development of the site exploration
program. Each site exploration program must be sufciently complete to provide the data required
for a geotechnical evaluation of the site and design of the coal refuse disposal facility.
6.4.1 Background Data Sources
In planning a refuse disposal facility and in the initial site explorations, much useful information can
be obtained from topographic, geologic and mine maps; agricultural soil surveys; satellite imagery
and aerial photographs; and publications available from various government agencies. In addition,
information from past investigations and area mining plans may also be searched to augment the
information from public agencies. While typically not of sufcient detail for design purposes, these
resources are readily available and represent an inexpensive source of valuable planning information.
6.4.1.1 Topographic Maps
Topographic maps are a vital source of information for planning a disposal facility and many
published and on-line sources are available. The most important source is the series of standard
topographic maps of the U.S. published by the U.S. Geological Survey (USGS). These maps cover
quadrangles of 7.5 minutes of latitude and longitude at a scale of 1:24,000 (1 inch = 2,000 feet).
More than 55,000 7.5-minute maps are available covering the 48 contiguous states. Other scales
and areas of coverage are available from older map series or from other agencies. These topo-
graphic maps show the variation of ground surface elevation using contours (i.e., lines of constant
elevation). The contour interval (elevation diference between adjacent contour lines) depends
upon the scale of the map and the steepness of the terrain. In the Eastern United States the contour
interval is usually 10 to 20 feet. In the western mountains, the contour interval is more commonly
50 feet. USGS topographic maps also show cultural and man-made features (roads, dams, build-
ings and political boundaries), water features (lakes, rivers and canals), wooded areas and areas
of past mining activity.
Topographic maps are useful for investigating an existing disposal facility, even though it may have
been developed subsequent to the topographic mapping. Comparison of the original topography
with the existing condition can provide an insight to the history and development of the site. Knowl-
edge of the topography of the surrounding area is important in planning the expansion or abandon-
ment of an existing disposal facility.
< PREVIOUS VIEW
6-45
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
FIGURE 6.11 FLOW CHART FOR SITE EXPLORATION, MATERIAL
PROPERTY TESTING
AND FACILITY DESIGN
Summarize basic soil/rock data and develop subsurface profile
Perform design, consider constructability issues, and identify construction
exploration and testing requirements
Select material properties and finalize subsurface profile
Review laboratory test data
Conduct laboratory testing
Review design objectives and initial results
Select representative soil/rock samples and plan laboratory testing
Review available site data and develop preliminary subsurface profile
Identify material properties needed for design and estimate scope of field program
Plan site exploration and field testing
Conduct site exploration and field testing
Prepare sample descriptions and perform laboratory index tests
P
H
A
S
E
I
I
S
T
U
D
Y
,
I
F
N
E
E
D
E
D
NO
YES
ENGINEERING
DESIGN
LABORATORY
TESTING
AND DATA
INTERPRETATION
SITE
EXPLORATION
AND FIELD
TESTING
FIGURE 6.11 FLOW CHART FOR SITE EXPLORATION, MATERIAL
Are results consistent with
preliminary profile?
Are there additional
data needs?
YES
Are results consistent and
representative of site
conditions?
Is a Phase II study
necessary?
NO
NO
YES
YES
NO
(SABATINI ET AL., 2002)
FIGURE 6.11 FLOW CHART FOR SITE EXPLORATION, MATERIAL PROPERTY TESTING
AND FACILITY DESIGN
< PREVIOUS VIEW
6-46
Chapter 6
MAY 2009
Another resource is USGSs The National Map web site, a framework for geographic knowledge of
the U.S. that provides public access to high-quality, geospatial data and information. The site allows
users to access, integrate, and apply geospatial data at global, national, and local scales. It includes a
variety of information layers such as boundaries, elevation, geographic names, geology (global seis-
mic networks, and real-time earthquakes), hydrography (real-time gauging stations and wetlands),
imagery, land use and land cover, natural hazards (climate and hurricanes), and topographic maps
that may be useful to designers.
6.4.1.2 Geologic Maps and Publications
Geologic maps of a proposed disposal facility site and surrounding area can provide valuable engi-
neering information. These maps are prepared by the USGS or by state geological surveys and gen-
erally show ground surface outcrops of various rock units using a color code and leter symbols.
Geologic maps usually have a column identifying formations and corresponding symbols and one
or more geologic sections depicting the regional structure of the rock, identifying the rock units and
providing a description of the units and their characteristics. To be most useful for evaluation of indi-
vidual disposal facility sites, a geologic map should preferably have a scale no larger than the USGS
topographic map of the area. Only a small part of the United States has been mapped in this detail,
however, and it is ofen necessary to use maps covering up to several hundred square miles. Geologic
maps can be valuable in the initial investigation and evaluation of a site, but proper interpretation of
these maps requires knowledge of the fundamentals of geology and an understanding of how geo-
logic information can be used in planning and design.
Specifc regional and local information on geologic conditions should also be considered. For exam-
ple, in steep Appalachian Valleys, joints and fractures from stress relief can afect excavation and
abutment stability and impoundment seepage control. Reports on physiography published by the
USGS (e.g., Wyrick and Borchers, 1981) can provide insight for planning exploration programs. Fur-
thermore, weathered joints and fractures encountered in eastern Kentucky and southern West Vir-
ginia, sometimes referred to as hillseams, can represent critical foundation or abutment features,
and mining publications may be helpful in planning associated with site preparation (e.g., Sames and
Moeb, 1989). Additionally, local mining and highway construction experience should be sought, as
it can disclose information on bedrock structure and fracture conditions that may infuence develop-
ment plans. Perin (2000) presents a case study that demonstrates the use and limitations of geologic
publications, mapping, and exploration for a refuse disposal site in eastern Kentucky.
The USGS provides access to a wealth of information resources including maps, reports, publica-
tions, and links to related web sites. Resources of possible interest for refuse disposal planning and
design at the time of publication include:
USGS Library Access to over 300,000 book, map, and serial records in the USGS
Library online catalog.
USGS Store Source for USGS maps and books, as well as products from other
agencies.
Publications Warehouse Search engine for 67,000 bibliographic citations.
Geologic Information National clearinghouse for geologic maps, datasets, and
related geoscience information with links to USGS geoscience databases and pro-
grams and resources for creating digital geologic maps.
National Water Data NWISWeb Comprehensive gateway to water-resources data
throughout the U.S.
National Atlas of the United States Comprehensive collection of small-scale geo-
spatial data from federal agencies.
< PREVIOUS VIEW
6-47
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
geodata.gov Web-based portal for access to maps, data, and other geospatial ser-
vices from across all levels of government.
6.4.1.3 Agricultural Soil Surveys
Much more widely available than large-scale geologic maps are soil surveys prepared by the Natural
Resources Conservation Service (NRCS). A soil survey is a detailed report on the surfcial (i.e., upper
5 to 6 feet) soils of a specifc area. Soil surveys typically have maps showing soil-type boundaries and
photographs, descriptions of soil characteristics, and tables of soil properties and features. The tables
section of a soil survey report provides information on soil properties including engineering index
properties, physical and chemical properties, and soil and water features. The tables section also has
information on soil use, such as crops and pasture, recreation, and engineering. Although data from
these surveys are generally not suitable for design analyses, the surveys are valuable tools for initial
site reconnaissance studies and for planning detailed feld explorations. Printed soil surveys can be
obtained from NRCS regional ofces or local soil conservation district ofces. Surveys are also avail-
able from the NRCS web site.
6.4.1.4 Satellite Imagery, Aerial Photographs and Other Imagery
Aerial and satellite photographs and other imagery can be extremely valuable in the investigation
and evaluation of a proposed or existing disposal facility site because they reveal much natural and
man-made detail that may not be apparent from the ground, no mater how carefully ground recon-
naissance is carried out. Also, these data can be used with Computer-Aided Design and Drafing
(CADD) and Geographic Information System (GIS) sofware to provide informative representations
of site conditions.
Data and imagery from satellites and aerial reconnaissance fights are increasingly being utilized for
site exploration and characterization purposes. These data are available from commercial vendors
as well as state and federal government agencies. At the time of publication of this Manual, satellite
images are available to the public at a resolution of as small as 2 feet and aerial photographs can be
obtained with a resolution of as small as 6 inches. An important use of satellite imagery and aerial
photographs is the performance of terrain analysis and lineament studies for identifcation and
location of surface features that are expressions of discontinuities in the underlying bedrock. Such
features may refect bedrock joint/fracture zones that warrant additional focus during exploration
programs for impoundments or that require assessment of the potential for breakthrough potential
to underground mines.
Photographs represent only a portion of the information available. Sophisticated sensors on satel-
lites and sensory equipment that can be mounted on airplanes can provide a spectrum of light band
information that when analyzed in combinations with sophisticated sofware allow identifcation of
various surface characteristics. Interpretation of these data is referred to as remote sensing, and some
applications include classifcation of land usage, classifcation of surface cover types, identifcation of
stressed vegetation or tree canopy, and delineation of surface thermal variations.
Some satellite systems that currently (2009) generate data that might be used in site exploration and
reconnaissance studies and for preparation of site drawings and fgures include:
Landsat QuickBird
IKONOS ASTER
SPOT EO-1
OrbView-3
< PREVIOUS VIEW
6-48
Chapter 6
MAY 2009
There are many forms of satellite data and many vendors that can provide satellite data packages.
These data providers can be identifed most easily from an Internet search. Also, a wide range of
information can be gathered from custom-designed aerial surveys. The types and quality of data
available and the associated costs should be carefully reviewed prior to purchase of data or contract-
ing for such data to be obtained.
Generally aerial and satellite imagery data are available or can be obtained in an orthorectifed format,
which means that the positions of all the data in the photograph or image are accurately located with
respect to a known coordinate system. Thus the data can be input to a GIS-based system and auto-
matically shown in true relationship to other orthorectifed data such as site boundaries, features,
structure, and infrastructure. The USGS through its Earth Resources Observation and Science (EROS)
center provides a wide range of such data, some of which can be obtained without charge.
Digital Raster Grids (DRGs) for some parts of the country can be downloaded free from the Internet
and printed, and these are generally identical to USGS topographic sheets. They are frequently used
as bases for CADD drawings, but they are not atractive for GIS applications because the various
types of data shown (e.g., contours, roads, structures, shadings) are combined into a single layer.
Aerial photographs corresponding to USGS quadrangles (referred to as Digital Orthophoto Quad-
rangles or DOQs) are also frequently used as bases for CADD drawings where proposed construc-
tion or property boundaries are shown over a photographic base. They can be used in a similar
manner in GIS-based sofware where individual layers representing roads, structures, utilities, etc.
are displayed over a photographic base. Also available are Digital Orthophoto Quarter Quadrangles
(DOQQs), which are orthorectifed quarter-quadrangle (7.5-minute coverage) photographs.
Digital Elevation Model (DEM) data can be used to generate elevation contour maps of a site. DRG
and DEM data can be combined in GIS-based sofware to generate three-dimensional models of USGS
topographic sheets.
The above data can be obtained from the USGS, and some states provide extensive data free over the
Internet. Pennsylvania, for example, provides 7.5-minute DRGs (also available with the surrounding
border cropped of), DOQs, DOQQs and DEMs for the entire state. Other information such as coal
mine maps and environmental data may be available from state web sites.
Soil-type data are available for some parts of the country in digital, orthorectifed form suitable for
use in GIS-based sofware.
These sources of aerial and satellite photographic data are increasingly available in an orthorectifed
(also referred to as georeferenced) format for input to GIS-based sofware where they can be auto-
matically displayed in accurate relationship to other georeferenced site data. The reliability, avail-
ability and costs of various types of data should be carefully evaluated when planning site drawings
and pictorial displays.
6.4.1.5 Past Investigations and Area Mining Plans
Information from past investigations and old mine plans in the vicinity of a planned or existing oper-
ation can provide valuable information for planning and preliminary design of a new facility. Particu-
larly valuable are maps of past or planned future mining, as well as geologic information on bedrock
structure, jointing and fracturing. As discussed in Section 8.2.1, mine maps are available at state
agencies (e.g., Virginia Department of Mines, Minerals and Energy; Kentucky Department of Mines
and Minerals; West Virginia Ofce of Miners Health, Safety and Training; Illinois State Geological
Survey) and from MSHA and the Ofce of Surface Mining (OSM). MSHA district ofces maintain
< PREVIOUS VIEW
6-49
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
maps until a mine closes and OSM stores a copy of the fnal map afer closure at the OSM National
Mine Map Repository in Pitsburgh, Pennsylvania. Additional information related to sources and
availability of mine maps is presented in Chapter 8.
Underground mine maps can be used to locate mine features with respect to the surface or other
underground mines and to determine the dimensions of pillars and mine openings (Shackleford,
2000). Information typically provided on underground mine maps includes (NRC, 2002):
Pillared, worked-out, and abandoned areas, pillar locations, sealed areas, future
projections, adjacent mine workings within 1,000 feet, surface or auger mines, mined
areas of the coalbed, and the extent of pooled water.
Dates of mining, coal seam sections, and survey data and markers.
Surface features, coal outcrop, and 100-foot-overburden contour or other prescribed
mining limit; mineral lease boundaries, surface property or mine boundary lines,
and identifcation of coal ownership.
The accuracy and completeness of underground mine maps varies due to the age and non-uni-
form standards followed in their development. For example, there are signifcant limitations to
some maps, particularly those for abandoned mines and mines operating before 1969. In addi-
tion, the horizontal and vertical (overburden) distances between mined barriers and an impound-
ment may not be accurately shown. Designers must consider these factors and the resulting
impacts in using mine maps for refuse disposal planning and design in the vicinity of active or
abandoned mines. Compounding the problem, some maps and records of older mines have been
lost or destroyed. Therefore, to confrm map accuracy, site exploration using the geotechnical and
geophysical exploration methods described in this chapter may be needed. Sections 8.2 and 8.3
present references and information to assist in locating available mapping and confrming its
accuracy.
6.4.1.6 Individual Site Mapping
For most disposal facility sites, site mapping should be performed before planning reaches the
design phase. Usually mapping is accomplished using low-altitude, large-scale aerial photography
to develop detailed, large-scale topographic maps. The topographic maps can be produced at the
scale and contour interval required for fnal site planning and design. Aerial topographic maps typi-
cally have contour intervals of one or two feet for relatively fat terrain and as much as 5 to 10 feet in
steeper terrain.
For geographic areas of the size associated with most disposal facility sites, the major cost of obtain-
ing aerial photographs is that of the aircraf. Therefore, it costs very litle more to photograph areas
adjacent to the anticipated site. This allows fexibility in making fnal plans and provides additional
data for interpreting site conditions that may afect the facility design.
6.4.2 Surfcial Reconnaissance and Geologic Mapping
The available topographic and geologic maps, aerial photography, and other documentation that per-
tain to the site should be supplemented by a surfcial reconnaissance and geologic mapping, which
consists of walking the disposal facility site and vicinity and observing topography, rock outcrops, mine
openings, soil types, vegetative cover, spring discharges, perennial and intermitent watercourses, and
any other features that may be relevant to the planned use of the site. This type of site reconnais-
sance generally requires a geologist or engineer who is familiar with refuse disposal and embankment
design and who can recognize the signifcance of the observed features. If possible, reconnaissance
should be conducted during times when vegetation is dormant so that site features are more visible.
< PREVIOUS VIEW
6-50
Chapter 6
MAY 2009
Consideration should also be given to conducting the reconnaissance shortly afer rainfall so that
spring and fow channel conditions that may be relevant to the design can be observed.
The elevation of the site should be compared to known or correlated elevations of mineable coal
seams in the vicinity. The site reconnaissance should include a search for evidence of past mining,
including but not limited to mine entries, sinkholes, highwalls, haul roads, spoil piles, discolored
seepage or watercourses, and areas with no vegetation or distressed vegetation. The presence of any
oil and gas wells, pipelines, and other underground or overhead utilities should be noted.
Rock outcrops should be observed for lithology, bedding, and structure. Structural observations
include strike and dip of strata, fracture orientation and spacing, and observable folds and faults.
Where weathered joints and fractures are encountered, in-fll materials and widths should be
recorded. Relatively recent fracturing should be noted as it may be indicative of subsidence. Bedding
observations include thickness, sedimentary structures (e.g., planar bedding, cross bedding), and lat-
eral continuity. Distinctive or known marker beds, such as coal seams or other persistent strata such
as limestone or dolomite should be noted. Lithologic observations (e.g., color, grain size, mineral
presence, weathering, and hardness) should also be noted.
Soils should be observed with respect to density, grain size, stifness, color, motling, structure, organic
mater, and depth. Slopes should be observed for evidence of recent or older landslides, including
slumps, scarps, and bent tree trunks. Mine rock waste (spoil) piles or other conditions related to pre-
vious mining such as clifs, strip pits or ponds should be noted. General vegetation type and density
should also be observed and documented.
Surfcial reconnaissance may be supplemented by test pits or shallow, hand-augered borings to pro-
vide samples for basic laboratory tests for soil classifcation. Test pits excavated to rock may help
disclose bedrock types, coal seam outcrops, and overburden jointing or fracturing. A careful surfcial
reconnaissance and accompanying tests can produce data sufcient for a preliminary surfcial map
of a refuse disposal facility site. However, as with any exploration program, it should be recognized
that undetected subsurface conditions are a risk, and reasonable contingencies should be considered
in the development of designs.
6.4.3 Subsurface Exploration and In-Situ Test Planning
Afer available information has been collected and evaluated, the designer can begin planning a
program for subsurface exploration and in-situ testing. The feld exploration methods, sampling
requirements, and types and frequency of feld tests to be performed should be determined based
on project design requirements, geologic conditions, the availability of existing subsurface informa-
tion, the availability of equipment resources, and local practice. ASTM D420, Standard Guide to Site
Characterization for Engineering Design and Construction Purposes, provides general guidelines
for site reconnaissance, exploration planning, equipment and procedures, geophysical exploration,
sampling, material classifcation, in-situ testing, interpretation and reporting.
An overall feld exploration and in-situ testing program for obtaining the data needed to defne sub-
surface conditions and perform engineering analyses and design should be developed. Once the
feld exploration and testing begins, the program may need to be modifed in response to site access
constraints (e.g., steep terrain may not be accessible to the available drilling equipment) or to address
variations in subsurface conditions not anticipated during exploration planning.
Site exploration programs are ofen conducted in phases. To obtain an overview of the geological
issues that can afect a facility, remote sensing, geophysical exploration and widely-spaced geotechni-
cal sampling and testing may be conducted as part of an initial phase. During a second or subsequent
< PREVIOUS VIEW
6-51
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
phase, localized disturbed and undisturbed sampling and in-situ testing may be conducted to obtain
more detailed information for defning geologic features and for determining geotechnical engineer-
ing properties for design. The types of subsurface exploration and testing activities in the typical
sequence in which they are conducted are:
Remote sensing
Geophysical investigations
Test pits
Disturbed sampling
In-situ testing
Undisturbed sampling
Remote sensing data can be used to identify terrain conditions, geologic formations, escarpments and
surface refection of faults or highly-jointed bedrock zones, buried stream beds, site access conditions
and general soil and rock formations. Remote sensing data from satellites (e.g., Landsat images from
NASA), aerial photographs from the USGS or state geologic surveys, and data from commercial
aerial mapping services may be useful. The designer should be familiar with the use of such data,
including limitations.
Geophysical methods ofer another means for characterizing subsurface conditions. Geophysical
methods can be used for general site characterization, mapping abandoned mine workings and mea-
suring physical properties in boreholes at coal refuse disposal facilities. For general site characteriza-
tion, geophysical methods can be used to determine the depth to bedrock, map ground stratigraphy,
detect sudden changes in subsurface formations, assess the rippability of bedrock, and map varia-
tions of physical properties within coal refuse and groundwater. Surface techniques such as electrical
resistivity, ground penetrating radar (GPR), electromagnetic conductivity (EM) or seismic refraction
can be applied. These techniques may be useful in defning broad variations in the subsurface, but
boreholes are needed for verifcation and interpretation.
Some geophysical methods can be used for detection of abandoned workings or cavities in karst for-
mations from the surface. For this purpose, the most commonly considered geophysical techniques
(although only recently applied at refuse impoundment sites) include electrical resistivity, seismic
refection and gravity. The success of these techniques depends on the depth to the mine workings,
degree of fooding, and thickness of the coal seam. Downhole geophysical methods can defne ver-
tical variations in physical properties. In particular, crosshole and downhole seismic tests induce
mechanical waves within the ground mass to provide information on the dynamic elastic properties
including the shear (S) wave velocity required for seismic site amplifcation studies of ground shaking
and for soil liquefaction evaluations. The application of surface and borehole geophysical methods
suitable for siting and engineering evaluations of refuse disposal sites is presented in Section 6.4.4.
Test pits are small excavations dug to a depth of 10 to 15 feet (i.e., to the typical reach of an excavator
or to refusal in rock). Compared to other exploration methods, test pits are more efcient because
they can provide information about a relatively large area inexpensively, and they expose a large
amount of soil for detailed examination by the feld geologist or engineer. Because the side walls of
a test pit can collapse rapidly, feld personnel should not climb into a hole deeper than about four
feet without assessment of soil stability and use of shoring, as appropriate. The locations for test pits
are typically selected in the feld as the site investigation program progresses. Test pits are generally
used to supplement data between borings or to explore areas where only near-surface conditions are
important, such as potential source areas for borrow material. The use of test pits for geologic map-
ping and material sampling is discussed in Section 6.4.3.3.
< PREVIOUS VIEW
6-52
Chapter 6
MAY 2009
Disturbed samples can be used to determine soil type, gradation, classifcation, water content, con-
sistency, relative density, and stratifcation. The samples are considered to be disturbed because
the sampling process modifes their natural structure. Disturbed samples are typically obtained
using track- or truck-mounted augers and other rotary-drilling techniques. The most common
disturbed sampling method is the Standard Penetration Test (SPT), which is performed using a
split-barrel sampler during the drilling of geotechnical borings. Geotechnical borings allow: (1)
testing as drilling progresses and recovery of samples, (2) measurement of groundwater levels
and collection of groundwater samples, and (3) installation of instrumentation for monitoring the
groundwater level or the deformation of the soil and rock at any depth. In planning an investi-
gation, boring locations should be selected so as to optimize the amount of useful data from the
drilling program. The basis for choosing boring locations and procedures for drilling borings is
discussed in Section 6.4.3.1.
Other in-situ test and geophysical methods can be used to supplement soil boring data. For instance,
the cone penetrometer test (CPT), also referred to as the cone penetration test, provides informa-
tion on subsurface soils without sampling. Stratigraphy and strength characteristics of soils can be
determined as the cone penetrometer is advanced. In-situ methods are most efective when they
are used in combination with conventional sampling to reduce the cost and the time required for
feld work. Data from these tests can be correlated with sampling and testing data obtained by
conventional means.
Undisturbed samples are obtained for laboratory testing when determinations of the in-place
strength, compressibility (setlement), natural moisture content, unit weight, or hydraulic conduc-
tivity are needed. They also allow observation of discontinuities, fractures and fssures associated
with subsurface formations. Although the sampling equipment is designed to minimize distur-
bance and these sample types are designated as undisturbed, in reality they are disturbed to
some degree. The degree of disturbance depends on the type of subsurface materials, type and
condition of the sampling equipment used, the skill of the drillers, and the storage and transporta-
tion methods used.
6.4.3.1 Program Planning
The number and depth of borings and locations of in-situ tests required for a particular subsurface
exploration program will depend on the size of the disposal facility site, the nature and uniformity of
the site geology, the magnitude of loads to be applied to the natural materials, the groundwater con-
ditions, the presence of past or active underground mining in the vicinity of the embankment, and
the complexity of the facility design. For example, explorations for new impounding embankments
normally require signifcantly more borings than those for non-impounding embankments.
The appropriate depth and spacing of borings is difcult to generalize because they depend upon
site conditions and disposal facility plans. An exploration program for a specifc site should provide
sufcient information for identifying, delineating and correlating geologic and soil conditions for
designing a safe and environmentally acceptable disposal facility. Ofen the fnal location of borings
and in-situ test sites must be determined in the feld, or additions must be made to the boring pro-
gram based upon evaluation of the initial data obtained. Excavation of test pits should be considered
as a means for supplementation of the boring program. Test pits facilitate examination of shallow
subsurface conditions and the recovery of bulk samples.
The spacing and number of borings beneath a dam depends on the complexity of the geology. Some
of the more important factors to consider are the character and continuity of the beds, elevation of the
strata, presence or absence of joints or faults, and proximity to previous underground mining. The
depth, thickness, sequence, extent, and continuity of the various strata should be determined.
< PREVIOUS VIEW
6-53
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
USDA (1978) guidance on exploratory borings for fresh-water dams includes the following:
Centerline of Dam Minimum of one boring on each abutment and at the outlet
structure transect, plus one boring at any abandoned stream channel, plus additional
borings for correlation of strata. Boring depths should not be less than the height of
the dam unless unweathered rock is encountered and is not underlain by compress-
ible strata or mine workings.
Outlet Conduit In addition to the boring at the transect with the centerline of the
dam, borings should be located at the vertical riser intake, downstream toe of the
dam, outlet of the conduit, and additional locations as needed to defne the rock sur-
face. Boring depths should not be less than the height of the backfll over the conduit
or 12 feet, whichever is greater, unless unweathered rock is encountered. At the riser
intake, the depth should not be less than the planned height of the riser above natu-
ral ground or 12 feet, whichever is greater.
Emergency Spillway Geologic cross sections based upon three or more borings
should be developed at the control section, intake section, and outlet section with
additional borings at cross sections as needed for correlation and location of strata
contacts and identifcation of excavation materials. Borings should extend to a depth
not less than 2 feet below the botom of the proposed spillway.
Foundation Drain Carefully-logged borings at the centerline of the dam and toe
may provide sufcient information, although additional borings or test pits may be
necessary where subsurface conditions are highly variable.
Tables 6.15 and 6.16 present guidelines for typical subsurface exploration and in-situ testing pro-
grams for new disposal facilities both with and without impoundments. The exploration program
to be used for a particular disposal facility and the associated in-situ testing must be developed by
a qualifed geotechnical engineer familiar with the requirements of the proposed facility. If inad-
equate data are obtained from the initial boring program, additional borings and/or test pits should
be advanced to supplement the original boring data.
An in-situ testing program typically involves SPTs obtained during boring advancement and uncon-
fned compression tests on recovered split-barrel samples performed by feld personnel using pocket
penetrometers or feld torvane equipment. Where sof sediments or fne coal refuse are critical to
embankment stability, CPTs provide an efective method for characterizing the consistency and
strength of the material. In-situ testing also generally includes hydraulic conductivity testing per-
formed in soil and rock to assess foundation conditions in valley botom and abutment areas. Instal-
lation of piezometers in completed borings can facilitate feld hydraulic conductivity testing as well
as site groundwater characterization.
Table 6.17 presents guidelines for a typical subsurface exploration and in-situ testing program for an
existing disposal facility. Similar to a new disposal facility, the in-situ testing program for an existing
disposal facility and the interpretation of subsurface data must be performed by a qualifed geotech-
nical engineer familiar with the facility design parameters. For an existing disposal facility, borings
are normally required when it is suspected that the embankment factors of safety are low or when
expansion of the embankment is planned, as discussed in Section 6.3.4, and some of the information
indicated in Table 6.17 may already be available from previous plans. The subsurface exploration and
in-situ testing program is greatly infuenced by site conditions, the disposal facility hazard potential,
the type of disposed refuse and its current condition, the stage of facility development, and the extent
of records of previous construction and placement of materials. Ofen, the in-situ testing program
will be designed to obtain the same information relative to the underlying foundation materials that
would be required for a new disposal facility. Extensive data relative to the materials and quality of
construction of the existing embankment must generally be obtained.
< PREVIOUS VIEW
6-54
Chapter 6
MAY 2009
TABLE 6.15 GUIDELINE SUBSURFACE EXPLORATION PROGRAM
FOR A NEW IMPOUNDING EMBANKMENT
Location Guideline Exploration Program
Abutments
One boring (minimum) should be drilled in each abutment to the maximum extension of the
embankment, to the depth of the valley bottom, or to a depth where the boring will over-
lap geologically with adjacent borings. Multiple borings may be required in each abutment
depending on the length of the dam and complexity of the geology.
Other locations
Additional borings, particularly for large embankments, beneath the structural portion of
embankment with spacing and depth to provide suffcient overlap between adjacent borings
to correlate data and to develop a geologic profle along the embankment.
Valley bottom or lowest
portion of embankments
not located in valley
bottom
A least one and generally multiple borings should be planned beneath the critical structural
portion of embankment. The depth of at least one boring should be approximately equal
to the planned height of the embankment unless frm bedrock is encountered at shallower
depth. Even if rock is encountered, deeper borings may be needed to suffciently reveal
ground water seepage/fow, to evaluate underlying deep mining, or for other special require-
ments. Test pits should be provided for the purpose of observing and documenting the
continuity of shallow soil stratigraphy and for obtaining bulk samples.
Upstream and
downstream
of crest
At least one boring upstream and one downstream from valley bottom borings for correlating
data and developing a geologic profle across embankment. The downstream boring should
be near the toe of the embankment where slope stability is expected to be most critical. Bor-
ings should penetrate to at least softest layer to be incorporated in stability analyses and pref-
erably into competent rock. Suffcient test pits to observe and document continuity of shallow
soil stratigraphy and obtain bulk samples should be provided.
Decant/spillway
structures
Borings and/or test pits along probable axes of structures to be founded on natural soil or
soft rock should be planned and should be drilled to a depth at least equal to the width of
the structures and preferably into very stiff soils or frm bedrock.
Embankment in vicinity
of past mining
Unless available mine maps and other information confrm that underground mining is dis-
tant enough to preclude potential impacts to the embankment, borings should be drilled at
locations and to depths necessary for assessment of the accuracy of the mine mapping
and for determination of the potential amount of subsidence, the probability of additional
subsidence, and the potential for mine breakthrough. Typically at least one boring (and to
assess some subsidence and breakthrough situations, several borings are required) should
be drilled to mine elevation to obtain a full profle of the overlying rock and to defne the
groundwater level in the mine. This is particularly important if the mine is located so that it
could adversely affect embankment stability, or if the mine is still in use and water infow
from the impoundment could imperil the miners or the mining operation. Appropriate safety
provisions should be taken, including wet drilling, with respect to the potential for encounter-
ing a potentially explosive gas mixture within the mine.
Impoundment area
Suffcient borings and/or test pits to observe and document continuity of shallow soil strati-
graphy and obtain bulk samples in order to determine the potential for ground water impacts.
Borrow areas
Suffcient borings and/or test pits to characterize materials, estimate available volume, and
obtain bulk samples.
General
Additional borings, as determined by a qualifed engineer or geologist, to meet special
requirements of planned disposal facility or to gain knowledge of special geologic factors
affecting design.
6.4.3.2 Subsurface Exploration and In-situ Test Methods and Applicability
This section provides information on various subsurface exploration and in-situ test methods that are
currently used for site characterization, sampling and determination of site-specifc soil and rock prop-
erties for the design of coal refuse disposal facilities. Conventional subsurface exploration and test-
ing programs typically include test pits, rotary drilling, SPTs, and disturbed and undisturbed sample
recovery. In-situ testing methods described in this section include SPT, CPT, piezocone penetrometer
test (CPTu), seismic piezocone penetrometer test (SCPTu), and vane shear test (VST). Additionally,
borehole testing to measure hydraulic conductivity is typically conducted. Standardized test proce-
< PREVIOUS VIEW
6-55
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
Location/Condition Guideline Exploration Program
Embankment toe
Suffcient borings and/or test pits should be provided near the planned location of toe to
explore conditions where foundation material may have a lower strength than embankment
material. Borings must extend past the depth where stability is a consideration and preferably
should extend to competent bedrock. Where the depth to bedrock is less than 10 feet, test
pits may be substituted for these borings.
Abutments
Suffcient borings and/or test pits should be provided to observe and document subsurface
conditions where stability is a consideration and for planned excavations for diversion ditches
or access roads.
Potential for ground-
water impacts
If seepage through refuse may create an undesirable environmental condition, borings should
be drilled suffciently deep to identify the groundwater level. These borings will provide data
for determining need for a drainage collection system and/or an impervious liner between the
coal refuse and the underlying foundation. Often one or two additional borings and several
test pits will need to be advanced at the general disposal facility site to determine groundwa-
ter conditions throughout the area that will be covered by refuse.
Embankment in vicinity
of past mining
Unless available mine maps and other information confrm that underground mining is suf-
fciently distant to preclude potential impacts to the embankment, borings should be drilled to
evaluate the potential effects of subsidence, the stability of the structural portion of embank-
ment, the potential for breakthrough, or if leachates from facility could adversely affect ground-
water quality in the mine.
Borrow areas
Suffcient borings and/or test pits should be provided to characterize materials, estimate
available volume, and obtain bulk samples.
General
Additional borings, as determined by a qualifed engineer or geologist, should be provided to
meet special requirements of the planned disposal facility or to gain knowledge of geologic
factors that could affect design.
dures for these in-situ testing methods have been developed by ASTM, and these are identifed in the
sections that follow. ASTM D 420, Standard Guide to Site Characterization for Engineering Design
and Construction Purposes, summarizes the various methods for site characterization. For in-situ test
methods that do not have standardized ASTM procedures, references for additional details are cited.
Conventional subsurface exploration methods typically involve the retrieval of soil samples and rock
core. Soil samples may be either disturbed (but representative) or undisturbed. Disturbed samples
are those obtained using equipment that destroys the macro structure of the soil, but does not alter its
mineralogical composition. Specimens from these samples can be used for determining the general
lithology of soil deposits; for identifying soil components and general classifcation purposes; and
for determining grain size, Aterberg limits, and compaction characteristics of soils. Undisturbed
samples are obtained in cohesive or fne-grained soil strata for use in laboratory testing to determine
engineering properties. Undisturbed samples of granular soils can be obtained, but specialized and
costly procedures are required such as freezing or resin impregnation and block or core type sam-
pling. The term undisturbed refers to the relative degree of disturbance to the in-situ properties
of the soil. Undisturbed samples are obtained with specialized equipment designed to minimize the
disturbance to the in-situ structure and moisture content of the soils. Specimens obtained by undis-
turbed sampling methods are used for determining the strength, soil layering, hydraulic conductiv-
ity, density, consolidation, dynamic properties, and other engineering properties. Common methods
used to obtain disturbed and undisturbed soil samples are presented in Table 6.18.
6.4.3.3 Test Pits
To evaluate materials near the ground surface and to supplement the information gained from the bor-
ings, test pits or test trenches are frequently employed. Test pits are usually excavated by a backhoe or
bulldozer and can range from a few cubic feet to a few cubic yards in volume. In addition to providing
TABLE 6.16 GUIDELINE SUBSURFACE EXPLORATION PROGRAM FOR
A NEW EMBANKMENT WITHOUT AN IMPOUNDMENT
< PREVIOUS VIEW
6-56
Chapter 6
MAY 2009
Location/Condition Guideline Exploration Program
Embankment
centerline
Typically, three or more borings should be drilled in a line perpendicular to embankment
axis. Borings should extend to a depth below the phreatic surface in the embankment
unless a water surface is not encountered for a depth signifcantly below that which
could reasonably be expected to affect the stability of the embankment. Piezometers
should be installed in selected borings to monitor the level of water surface within the
embankment and foundation. Special efforts should be made to evaluate materials
and water levels encountered as a function of depth within these borings to deter-
mine if horizontal impervious zones are present that could affect seepage through the
embankment.
Downstream
embankment
face
A line of borings parallel to the axis of the embankment should be drilled to a depth below
the phreatic surface. Normally these borings should be drilled in the downstream face
of embankment where the phreatic surface level is most critical to stability. This location
should be determined based on a profle developed from data from the frst line of bor-
ings perpendicular to the embankment axis.
No information on
foundation preparation or
foundation stability is a
concern
If no information is available relative to procedures originally used to prepare the embank-
ment foundation or if the designer believes that stability along the existing foundation
may be critical, at least one and preferably multiple borings should be drilled in the
embankment axis and downstream face and should extend through the embankment
and into competent rock.
Downstream valley
bottom
At least one boring should be drilled at the highest section of embankment where stabil-
ity is likely to be critical. Multiple borings may be required in this area if expansion of the
disposal facility to a greater elevation is planned.
Facility expansion
Additional borings should be drilled at locations that will appropriately allow for analyses
associated with enlarging the embankment, as determined by designer.
Settled fne
refuse
If the embankment impounds settled fne refuse slurry and an expansion of the
embankment over slurry is contemplated, borings should be extended into the slurry
in order to obtain samples for laboratory testing. The exploration program may also
entail the use of cone penetration testing and geophysical surveys. If the embank-
ment was built using the upstream method, the extent to which fne refuse underlies
the embankment should be determined by drilling borings and/or use of geophysical
methods.
Embankment in vicinity
of past mining
If past mining has been completed beneath any portion of the existing disposal facility,
borings should be drilled at locations and depths necessary to determine the potential
subsidence, the probability of additional subsidence, and the potential for mine break-
through. Generally, at least one boring should be drilled to below the mine elevation
to obtain a full profle of overlying rock and to defne the groundwater level at the mine
elevation. This is particularly important if the mine is located such that it could adversely
affect the stability of the existing embankment or if the embankment impounds water and
the mine is still in use (i.e., where water infow from the impoundment could be unsafe to
miners or mining operations). Appropriate safety provisions should be taken to prevent
possible gas release through any boring drilled to an abandoned mine.
Decant/spillway
structures
If changes in the existing disposal facility include plans for decant or spillway structures
founded on soil, soft rock or refuse materials, borings should be drilled and/or test pits
advanced along the probable structure axis to provide data for a suffcient depth to prop-
erly design foundations or bedding requirements.
General
The required boring program and any modifcations should be determined by a qualifed
geotechnical engineer. Flexibility will be required on the part of the engineer because the
boring program will frequently need to be modifed as the feld investigation proceeds in
order to resolve issues arising from data obtained from earlier borings.
TABLE 6.17 GUIDELINE SUBSURFACE EXPLORATION PROGRAM
FOR AN EXISTING EMBANKMENT
< PREVIOUS VIEW
6-57
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
access to larger samples than possible from borings, test pits permit direct visual examination of the
soil in place. In-situ density and feld shear strength tests also may be conducted in test pits at various
depths. Test trenches permit observation of lateral variations of soil conditions over the trench length.
This may be particularly useful in residual soils produced by in-place weathering of rock where several
degrees of weathering and initial rock quality may be present and in colluvial soils where large varia-
tions in gradation may be observed. If compressible cohesive soils are present, test pits can provide
access for undisturbed sampling of soil blocks, as described in ASTM D 7015, Standard Practices for
Obtaining Undisturbed Block (Cubical and Cylindrical) Samples of Soils.
6.4.3.4 Boring Methods and the Standard Penetration Test (SPT)
A variety of drilling methods can be used for subsurface geotechnical exploration, including dis-
placement, wash, and auger borings, and percussion drilling. Augering in soil and coring in rock are
preferred because these methods permit recovery of representative samples for classifcation and
testing. Depending on the terrain, either truck- or skid-mounted drilling rigs are used. Truck rigs are
generally more powerful and drill faster, but skid rigs are more maneuverable in rough or heavily
wooded terrain and are less difcult to mobilize.
For drilling in soil, augers ranging from 6 to as much as 18 inches in diameter are used. Soil samples
can be obtained from: (1) the auger cutings, (2) a split-barrel drive sampler (disturbed sample), or (3)
a thin-walled tube (undisturbed sample). Usually, the boring is advanced by augering a set distance
(2 to 10 feet) or until there is a change in soil layer, and either a split-barrel or thin-walled tube sample
is then taken.
Where undisturbed samples are not required, split-barrel samples are obtained in accordance with
ASTM D 1586, Standard Test Method for Penetration Test and Split-Barrel Sampling of Soils. The
SPT is accomplished by placing a hollow, thick-walled-tube sampler at the botom of the boring
TABLE 6.18 COMMON SAMPLING METHODS
Sampler Sample Type Appropriate Soil Types Method of Penetration
Split-barrel
(Split-spoon)
Disturbed Sands, silts, clays Hammer driven
Thin-walled
Shelby tube
Undisturbed
Clays, silts, fne-grained soils, clayey
sands
Mechanically pushed
Continuous
push
Partially
Undisturbed
Sands, silts, and clays Hydraulic push with plastic lining
Piston Undisturbed Silts and clays Hydraulic push
Pitcher Undisturbed
Stiff to hard clay, silt, sand, partially
weathered rock, and frozen or resin
impregnated granular soil
Rotation and hydraulic pressure
Dennison Undisturbed
Stiff to hard clay, silt, sand and partially
weathered rock
Rotation and hydraulic pressure
Continuous
auger
Disturbed Cohesive soils Drilling with hollow-stem augers
Bulk Disturbed Gravels, sands, silts, clays Hand tools, bucket augering
Block Undisturbed
Cohesive soils and frozen or resin-
impregnated granular soil
Hand tools
(ADAPTED FROM MAYNE ET AL., 2002)
< PREVIOUS VIEW
6-58
Chapter 6
MAY 2009
and driving it 18 inches into the underlying soil by blows from a 140-pound hammer dropping 30
inches. The number of blows required to drive the sampler each 6-inch interval is recorded. The frst
6-inch interval is regarded as a seating value, and the blows for the second and third increments
are summed to give the SPT N-value or blow count resistance of the soil. If the sampler cannot be
driven 18 inches, the number of blows for each 6-inch increment and for each partial increment is
recorded on the boring log. For partial increments, the depth of penetration is recorded in addition
to the number of blows. The SPT can be performed in a wide variety of soil types, as well as weak
rocks, but it is not particularly useful for the characterization of gravel deposits or sof clays. The
SPT provides a semi-quantitative measure of the stifness or density of the soil in place. When the
sampler is removed from the boring, a representative soil sample is recovered for classifcation and
for laboratory tests that are applicable to disturbed soil samples, including moisture content, grain
size analysis and Aterberg limits. The advantages and limitations of the SPT are summarized in
Table 6.19.
A properly conducted boring program entails close supervision by an experienced engineer or geolo-
gist. This supervision includes: (1) careful and detailed classifcation of the materials recovered from
the boring, (2) preparation of a detailed log for each boring, noting the classifcation of the material
and its condition, and (3) other signifcant observations such as water levels in the boring.
The boring log is the basic record for geotechnical exploration and provides a detailed record of the
work performed and the subsurface conditions at the boring location and can be recorded on paper
or on electronic data loggers. If recorded on paper, feld boring logs should be writen or printed
legibly and should be as clean as is practical considering site conditions and weather. All appropri-
ate portions of the boring logs should be completed in the feld as the work is being performed.
A wide variety of drilling log forms are in use, but the specifc form(s) to be used for a given type of
boring will depend upon local practice. A boring log should provide a description of exploration pro-
cedures and subsurface conditions encountered during drilling, sampling and coring. The following
information should be provided on a boring log:
Survey data including boring location in reference to site coordinates, surface eleva-
tion, and bench mark location and datum, if available.
An accurate record of any change in the planned boring locations.
Identifcation of the soil and bedrock encountered, including density, consistency,
color, moisture, structure, and geologic origin.
The depths of the various generalized soil and rock strata encountered.
Sampler type, depth, penetration, and recovery.
Sampling resistance in terms of hydraulic pressure or blows per depth of sampler
penetration; size and type of hammer; height of drop.
Soil sampling interval and recovery.
Rock core run numbers, including depths and lengths, core recovery, and rock qual-
ity designation (RQD).
Drilling method used to advance and stabilize the hole.
Comparative resistance to drilling.
Observed loss of drilling fuid.
Water level observations.
The date and time that the boring was started, completed, and when water level
measurements were made.
Closure of borings.
< PREVIOUS VIEW
6-59
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
TABLE 6.19 ADVANTAGES AND LIMITATIONS OF THE STANDARD PENETRATION TEST
Advantages Limitations
Obtain both a sample and a number Disturbed sample (index tests only)
Simple, rugged and suitable in many soils Analysis required if varying numerical results are obtained
Can be used in weak rock Not applicable for soft clays and silts
Available throughout the U.S. High variability and uncertainty
(ADAPTED FROM MAYNE ET AL., 2002)
6.4.3.5 Undisturbed Soil Sampling
Undisturbed samples are usually obtained when the structure of the soil (e.g., strength and com-
pressibility) is important to its behavior. Relatively undisturbed samples are commonly obtained by
pushing a thin-walled tube into the soil at the botom of the boring and removing the soil sample
from the boring in the protective tube. The sampler typically has an approximate outside diameter of
3 inches and inside diameter of 2 inches. Thin-walled samplers vary in outside diameter between
2 and 3 inches and typically come in lengths from 28 to 36 inches. Larger diameter tubes are used
where higher quality samples are desired and sampling disturbance must be kept to a minimum. The
procedure for thin-walled tube sampling is described in ASTM D 1587, Standard Practice for Thin-
Walled Tube Sampling of Soils for Geotechnical Purposes.
Undisturbed sampling is generally practical only for fne-grained, cohesive soils that contain few rock
particles. Undisturbed tube sampling of coarse gravels and coarse refuse with large particles is not
practical because of the resistance to pushing the tube and potential damage to the tube. Special types
of tubes and procedures are sometimes employed to obtain suitable undisturbed samples of very sof
or sensitive soils, such as very wet fne refuse. These are described in Mayne et al. (2002). If these types
of samples are desired, the ability of the potential drillers to obtain them should be verifed.
The thin-walled tubes used for undisturbed sampling are manufactured using carbon steel, galva-
nized carbon steel, stainless steel, and brass. Carbon steel tubes are ofen used, but are unsuitable if
the samples are to be stored in the tubes for more than a few days because of rusting. In stif soils,
galvanized carbon steel tubes are preferred because carbon steel is stronger, less expensive, and the
galvanizing provides additional resistance to corrosion. Thin-walled tubes are manufactured with
a beveled front edge to reduce pushing resistance and sample disturbance. Thin-walled tubes can
be pushed with a fxed head or piston head. ASTM D 4220, Standard Practices for Preserving and
Transporting Soil Samples, provides guidance for feld preparation, transport and storage of undis-
turbed samples prior to laboratory testing.
6.4.3.6 Rock Coring
Where borings must extend into weathered and unweathered rock, rock drilling and sampling are
required. For disposal sites, defning the top of rock by drilling can be difcult, especially when large
boulders are present and where the top of rock profle is irregular. In all cases, care must be taken in
determining the top of rock because improper identifcation may lead to a miscalculated thickness of
rock overburden above a mine or inaccurate determination of material quantities.
Destructive rock drilling is a relatively quick and inexpensive means for advancing a boring when
an intact rock sample is not required. Destructive drilling can be used to determine the elevation
of the top of rock or the elevation of the top of a mine void. Types of destructive drilling include
air-track drilling, down-the-hole percussive drilling, rotary tricone (roller bit) drilling, rotary drag
bit drilling and, in very sof rocks, augering with carbide-tipped bits. When destructive drilling is
< PREVIOUS VIEW
6-60
Chapter 6
MAY 2009
employed, caution should be exercised in determining the top of sof rock because drilling pro-
ceeds rapidly, and weathered and sof rock can be easily penetrated, resulting in an inaccurate
top-of-rock elevation.
When formations are encountered that are too hard to be sampled by soil sampling methods (typi-
cally more than 50 blows per inch with a 2-inch-diameter, split-barrel sampler), core drilling proce-
dures should be employed, as described in ASTM D 2113, Standard Practice for Rock Core Drilling
and Sampling of Rock for Site Investigation. Seismic refraction or other geophysical methods can
be used to assist in determining the top-of-rock elevation. Seismic-refraction data can also provide
information between borings. The depth of rock coring will vary depending on site conditions, but as
a minimum coring should extend to a depth sufcient to account for the presence of pervious or sof
strata that could afect the stability of the embankment.
Core barrels may be single-, double-, or triple-tube types. A double-tube core barrel is commonly
used because the inner and outer core tubes beter isolate the rock core from the drilling fuid
stream and the inner tube isolates the core from the rotating outer tube. In triple-tube core bar-
rels, the inner tube may be longitudinally split to allow observation and removal of the core with
reduced disturbance.
Rock coring can be accomplished with either conventional or wireline equipment. With conventional
drilling equipment, the entire string of rods and core barrel are raised to the surface afer each core
run for rock core retrieval. Wireline drilling equipment allows the inner tube to be uncoupled from
the outer tube and raised rapidly to the surface by means of a wire-line hoist. The main advantage
of wireline drilling is the increased drilling production resulting from the rapid removal of the core
from the hole. Wireline coring also provides improved quality of recovered core, particularly in sof
rock, because this method minimizes rough handling of the core barrel during retrieval of the barrel
from the borehole and when the core barrel is opened. Wireline drilling can be used on any rock
coring project, but typically is used where boreholes are more than about 75 feet deep and rapid
removal of the core from the hole has a greater efect on cost.
Although NX (2.154-inch-diameter) core is the size most frequently used for engineering explora-
tions, both larger and smaller sizes are sometimes used. Larger core sizes will usually produce greater
recovery and less fracturing during drilling.
The length of each core run should be limited to a maximum of 10 feet. Core run lengths should be
reduced to 5 feet or less just below the rock surface and in highly fractured or weathered rock zones.
Shorter core runs generally reduce the degree of damage to the core and improve core recovery in
poor quality rock.
The core bit provides the grinding action at the botom of the core barrel assembly that cuts the core
from the rock mass. Diamond, carbide-tipped and sawtooth core bits are the most commonly used.
Core bits are generally selected by the driller and are ofen approved by the geotechnical engineer.
Bit selection should be based on drill bit performance for the expected formations and the proposed
drilling fuid.
Clear water is most ofen used as the drilling fuid in rock coring because it is readily available, does
not react with most rock types and does not require special disposal procedures. If collapsing holes or
zones where there is loss of drill water are encountered, a drilling mud may be required for stabiliz-
ing the borehole. Drilling mud should be used with care because it will clog open joints and fractures
and can adversely afect hydraulic conductivity measurements and piezometer installations. A setling
basin should be used to remove drill cutings and to allow recirculation of the drilling fuid. Unless
contaminated with oil or other substances, drilling fuids can be discharged onto the ground surface.
< PREVIOUS VIEW
6-61
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
Rock core should be carefully removed from the core barrel, placed in a rock core box appropriately
sized for the diameter of core drilled, and visually classifed. The rock core recovery and RQD should
be calculated and recorded. ASTM D 6032, Standard Test Method for Determining Rock Quality
Designation (RQD) of Rock Core, should be followed in determining the RQD, which is a normal-
ized measure of the degree of rock fracturing. The rock core should be preserved and transported
following the guidance in ASTM D 5079, Standard Practices for Preserving and Transporting Rock
Core Samples. Additional guidance for visual classifcation, core handling and labeling, and other
feld practices is provided in Mayne et al. (2002). The application of RQD in geotechnical design is
presented in Section 6.6.
6.4.3.7 Cone Penetrometer Test (CPT) and Piezocone Penetrometer Test (CPTu or PCPT)
An alternative or supplement to the SPT is the Cone Penetrometer Test (CPT), also referred to as Cone
Penetration Test, an in-situ test that is fast, economical, and provides continuous profling of soil
strata and soil properties. The test is described in ASTM D 5778, Standard Test Method for Perform-
ing Electronic Friction Cone and Piezocone Penetration Testing of Soils, and consists of pushing a
cylindrical steel probe into the ground at a constant rate of 2 centimeters per second and measuring
the resistance to penetration. The standard cone penetrometer has a conical tip with an apex angle
of 60 degrees, a 10-cm
2
projected area for the cone, and a 150-cm
2
surface area for the friction sleeve.
The ASTM standard also permits a larger diameter unit that has a 15-cm
2
tip and 200-cm
2
sleeve. The
measured point or tip resistance is q
c
and the measured side or sleeve resistance is f
s
. An illustration
of a typical cone penetrometer is provided in Figure 6.12.
The CPT can be used in very sof clays to dense sands, but it does not work well in gravels or rocky
terrain. The advantages and limitations of using the device are summarized in Table 6.20. Because
the CPT provides more accurate and reliable parameters for analysis, it is an excellent complement
to traditional soil borings with SPT measurements. The CPT is not practical for coarse refuse where
larger rock fragments can impede the penetrometer, but the method has been used to characterize the
consistency and to estimate the engineering properties of setled fne refuse in impoundments.
A piezocone penetrometer test (CPTu or PCPT) is performed by advancing a cone penetrometer with
transducers for measuring pore-water pressures. In clean sands, the measured pore pressures are
nearly hydrostatic because the high hydraulic conductivity of the sand permits immediate dissipa-
tion of excess pore-water pressures mobilized by advancement of the cone. In clays, the advancement
of the cone may result in the development of elevated pore-water pressures. If the advancement of
the penetrometer is halted, the decay of pore-water pressures can be monitored with time and used
to calculate an in-situ rate of consolidation and soil hydraulic conductivity. Details related to test
methods, cone types and calibration, data reduction, and cone maintenance are provided in ASTM
D 5778 and Lunne et al. (1997).
Piezocone penetrometer testing can be a viable technique for determination of gradational vari-
ability, strength, hydraulic conductivity and consolidation properties of fne coal refuse at existing
refuse disposal sites. However, the measured cone resistance q
c
must be corrected for pore-water
pressures acting on unequal areas of the cone tip. This correction is most important for sof to stif
clays and silts and for very deep soundings where the hydrostatic pressures are high. Usually in
sands, the correction is minimal because q
c
is much greater that any mobilized pore pressures.
Because soil samples are not obtained with the CPT, indirect assessment of soil behavior is typically
inferred from an examination of the test data. The data can be processed for use in empirical chart
classifcation systems, or the raw readings can be interpreted by eye to determine soil strata changes.
For example, clean sands are generally indicated by a total tip resistance q
T
greater than 50 tsf, while
for sof to stif clays and silts, q
T
is less than 20 tsf. The total tip resistance q
T
is a function of the pore
< PREVIOUS VIEW
6-62
Chapter 6
MAY 2009
pressure behind the cone tip q
c
and some factors related to cone geometry. This value is automatically
calculated and ploted during the test.
Generally, pore-water pressures associated with penetration in loose sands are approximately equal
to hydrostatic pressures, in contrast to penetration in dense sands where the pore-water pressure
is typically less than hydrostatic. In sof to stif intact clays, pore-water pressures associated with
advancement of the penetrometer are generally several times the hydrostatic pressure. Notably, neg-
ative pore-water pressures are observed in fssured overconsolidated materials. The sleeve friction,
FIGURE 6.12 PIEZOCONE PENETROMETER
(ADAPTED FROM FHWA, 1992)
FRICTION SLEEVE
PIEZO-ELEMENT IN
CONE TIP
PIEZO-ELEMENT
BEHIND CONE TIP
PIEZO-ELEMENT
BEHIND FRICTION
SLEEVE
ELECTRONIC HOUSING
60
CONE TIP
TABLE 6.20 ADVANTAGES AND LIMITATIONS OF THE CONE PENETROMETER TEST
Advantages Limitations
Fast and continuous profling No soil samples are obtained
Economical and productive Unsuitable for gravel or boulder deposits
(1)
Results not operator-dependent Requires skilled operator to run
Strong theoretical basis for interpretation Electronic drift, noise, and calibration
Particularly suitable for soft soils High capital investment
Note: 1. Except where special rigs are provided and/or additional drilling support is available.
(MAYNE ET AL., 2002)
FIGURE 6.12 PIEZOCONE PENETROMETER
< PREVIOUS VIEW
6-63
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
ofen expressed in terms of a friction ratio (FR = f
s
/q
T
), is also a general indicator of soil type. In sands,
FR usually falls in the range of 0.5 to 1.5 percent; in clays FR normally falls between 3 and 10 percent.
A notable exception is that in sensitive and quick clays, a low FR is observed. An approximate esti-
mate of clay sensitivity is 10/FR (Robertson and Campanella, 1983).
6.4.3.8 Field Vane Shear Test (FVST)
The feld vane shear test (FVST) is used to evaluate the in-situ undrained shear strength of sof to stif
clays and silts, mine tailings and organic muck. The test is conducted in accordance with ASTM D
2573, Standard Test Method for Field Vane Shear Test in Cohesive Soil, by inserting a four-bladed
vane (Figure 6.13) into cohesive soil at the botom of a boring and rotating the device about a vertical
axis. The torque required to turn the device is measured. A variety of vane sizes, shapes, and confgu-
rations are available depending upon the consistency and strength characteristics of the soil. Vanes
can have a blade diameter D ranging from 1.5 to 4.0 inches, a vane height H ranging between 1.0 and
2.5 D, and a blade thickness ranging from 0.006 to 0.125 inches. The end of the vane is usually rectan-
gular or tapered at 45 degrees.
ASTM D 2573 provides relationships for converting the measured peak torque to a value of peak
undrained vane shear strength S
uv
based on the geometry of the vane. For a rectangular vane with
H/D = 2:
S
uv
= 6T
max
/ (7D
3
) (6-1)
where:
T
max
=
maximum measured torque corrected for apparatus and rod friction
(length times force)
Relationships for other vane geometries are presented in ASTM D 2573.
Afer the test to determine S
uv
is completed, the undrained steady-state (residual) shear strength S
ur
can be determined by quickly rotating the vane another 5 full revolutions to fully remold the soil and
then repeating the shear test. The ratio of peak to remolded undrained strengths is the sensitivity S
t
.
Table 6.21 provides a summary of the advantages and limitations of the FVST. Additional guidelines
related to application of the FVST are presented in Mayne et al. (2002).
ASTM D 2573 recommends a loading rate of no faster than 0.1 degree per second (15 minutes for 90
degrees of rotation). At this rotation rate, the time required to reach undrained peak strength typically
ranges from 2 to 5 minutes, but in very sof clays the time to failure may be as much as 10 to 15 minutes.
Chandler (1988) and Morris and Williams (2000) investigated the applied loading rate. Chandler
applies a theoretical method by Blight (1968), and indicates that for typical vanes and a time to failure
of 1 minute, the test will be undrained if the coefcient of consolidation (c
v
) is less than 0.035 cm
2
/sec
(3.3 f
2
/day). Blight defned time to failure as the time from the beginning of vane rotation. Morris
and Williams (2000) proposed a revision to Blights theoretical method, which accounts for the pore
pressure increase due to vane insertion as well as vane rotation and defnes time to failure as the time
from vane insertion. Morris and Williams indicate that, for a vane diameter of 63 mm (2.5 in), a time
to failure of 2 minutes will result in an undrained test for materials with c
v
as high as 1450 m
2
/year (42
f
2
/day). This value of c
v
should encompass coal refuse and natural materials with a plasticity index
(plastic limit minus liquid limit) of 10 or higher.
To minimize drainage in fne coal refuse during the FVST, the rotational loading should be applied as
soon as possible afer vane insertion, and the loading rate should be increased signifcantly from the
< PREVIOUS VIEW
6-64
Chapter 6
MAY 2009
ASTM D 2573 recommendation in order to achieve failure within about 1 minute. For a sof material,
if only the peak strength is being measured, a loading rate of 2 to 10 degrees per second is reasonable,
but if the undrained, steady-state (residual) strength is being measured as well as the undrained peak
strength, then the following procedure is recommended:
1. Initially apply the torque at a rate of about 10 degrees per second.
2. Afer the peak strength has been reached, increase the rate of rotation to at least 60
degrees per second (6 seconds per revolution or faster) for at least 5 complete revolu-
tions to remold the material.
3. Avoiding a rest period, slow the rate of rotation to about 10 degrees per second to
measure the steady-state strength.
The rotation and torque should be measured and recorded as the test is conducted, which can be
accomplished with a gear box and stylus recording system or other type of data acquisition system.
The rod and apparatus friction corrections (per ASTM D 2573) should be performed for the rates of
rotation actually used in steps 1 and 3.
FIGURE 6.13 FIELD VANE SHEAR TEST SETUP
ADAPTER
TORQUE WRENCH
DRILL ROD
BEARING GUIDE
COUPLING
CASING
VANE
COLLAR THRUST
BEARING
THRUST BEARING
GUIDE
VANE TO DRILL
ROD ADAPTER
H
D
45
FIGURE 6.13 FIELD VANE SHEAR TEST SETUP
FIGURE 6.13 FIELD VANE SHEAR TEST SETUP
< PREVIOUS VIEW
6-65
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
Studies by several researchers have demonstrated the importance of correcting the measured vane
strength for use in stability analyses involving embankments on sof ground, bearing capacity analy-
ses, and for analyses associated with excavations in sof clays. The correction to obtain the mobilized
shear strength is given by:
S
u (mobilized)
=
R
S
uv
(6-2)
where:
R
= empirical correction factor related to plasticity index (PI) based on
back calculation from failure case history records of full-scale projects
(dimensionless)
Bjerrum (1972) recommended values of
R
to correct the measured peak feld vane strength (with
a time to failure of a few minutes or less) to a value of S
u (mobilized)
(during a full-scale slope failure
corresponding to a time to failure of several weeks to several months) that may be appropriate for
stability failures.
Chandler (1988) combined Bjerrums case history data and other data sets to develop a more specifc
strain-rate correction factor:
R
= 1.05 b (PI)
0.5
(6-3)
where b is a dimensionless rate factor that depends on the time to failure (t
f
) in minutes (for a full-
scale slope stability failure):
b = 0.015 + 0.0075 log t
f
(10 min < b < 10,000 min) (6-4)
Values of
R
as a function of PI and t
f
are presented in Figure 6.14. Slope stability failures should gen-
erally be considered to have t
f
values of 10,000 minutes (7 days) because of the construction methods
involved.
These strain rate correction factors are for peak undrained strengths and are based on case histories
for natural clays, not fne coal refuse. However, until more research is available, the same correction
factors should be applied to fne coal refuse and to remolded undrained (steady-state) strength as
well as to undrained peak strength.
The FVST is applicable only to sof to medium stif clay-like materials that are relatively impermeable
such that they remain undrained during the test. If drainage occurs, the measured torque and result-
ing calculated strength will exceed the actual value. No published guidance is available on limitation
TABLE 6.21 ADVANTAGES AND LIMITATIONS OF THE FIELD VANE SHEAR TEST
Advantages Limitations
Assessment of undrained strength (S
uv
) Limited application to soft to stiff clays
Simple test and equipment Slow and time-consuming
Measurement of in-situ clay sensitivity (S
t
) Raw S
uv
needs correction (empirical)
Long history of use in practice Can be affected by sand lenses and seams
(MAYNE ET AL., 2002)
< PREVIOUS VIEW
6-66
Chapter 6
MAY 2009
relative to PI; this Manual recommends that the FVST should not be used for materials with a PI of 7
or less because these materials are likely to drain during the test. FVST should be used with caution
in materials with higher PI that contain thin layers of sand-like material because the sand-like layers
may allow drainage. For material with PI between 7 and 10, FVST should only be used if supporting
data are provided to demonstrate that the test was undrained. Material samples should be recovered
from each test zone for geotechnical index testing (moisture content, grain size distribution, and
Aterberg Limits at a minimum). Also, CPT and piezocone measurements performed adjacent to an
FVST can be used to measure the rate of pore-pressure dissipation, so that it can be confrmed that
the zone in which the FVST is run is relatively impermeable. The piezocone data may be used to esti-
mate the value of c
v
in order to confrm that the rotation rate is sufciently rapid that the test can be
considered undrained.
The FVST is not intended for stif clay-like materials because these materials will normally exceed the
torque capacity of the FVST device. The FVST cannot be used for testing sand-like material or coarse
refuse because (1) these materials will drain during the test and (2) the shear strength of these materi-
als will exceed the capacity of the FVST device.
6.4.3.9 Directional (Longhole) Drilling
Directional or longhole drilling refers to: (1) in-mine drilling operations used to identify geological
and mining conditions in advance of mining and (2) surface drilling through an outcrop to determine
cover and coal barrier thickness. Development in the 1990s of systems with instrumentation to mea-
sure drill bit location, high-thrust drilling equipment, powerful downhole motors and high-strength
drill tubing has allowed the implementation of this technique. In combination with hydraulic frac-
PLASTICITY INDEX, PI (%)
V
A
N
E
C
O
R
R
E
C
T
I
O
N
F
A
C
T
O
R
,
R
FIGURE 6.14 VANE CORRECTION FACTOR
0 20 40 60 80 100 120
0.5
0.6
0.7
0.8
0.9
1.0
t
f =
10
t
f =
1
0
0
t
f =
1
0
0
0
t
f =
1
0
,0
0
0
CORRECTION FACTOR FOR
EMBANKMENTS UNDER NORMAL
RATES OF CONSTRUCTION
tf = TIME TO FAILURE (MINUTES)
(MAYNE ET AL., 2002)
FIGURE 6.14 VANE CORRECTION FACTOR
FIGURE 6.14 VANE CORRECTION FACTOR
< PREVIOUS VIEW
6-67
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
turing techniques used to increase connectivity between boreholes, directional drilling has been
employed to reduce in-situ methane gas contents in low-hydraulic-conductivity coal and to fracture
a massive sandstone roof in advance of longwall mining (Brunner and Schwoebel, 1999). Depending
on conditions, drilling rates of 300 feet per shif and drilling accuracies of approximately 1 degree in
azimuth and 0.5 inches in pitch can be achieved. The longest reported in-mine horizontal borehole
exceeds 5,000 feet in length (Brunner and Schwoebel, 1999). Directional drilling has also been used to
locate old abandoned workings, drain accumulations of mine water, and degasify gob areas (Kravits
and Schwoebel, 1994). While conventional coring should not be atempted in a long, directionally-
drilled borehole, spot cores can be taken at selected locations (Kravits and Schwoebel, 1994). As such,
the technique may have applicability for accurately locating and determining the thickness of hori-
zontal in-mine barriers and can be a means of validating geophysical methods for locating and sizing
barriers in mines.
6.4.3.10 Field Hydraulic Conductivity Tests
The hydraulic conductivity of soil or rock is ofen measured in place during subsurface exploration
to determine if seepage through foundation materials will be an important design consideration. The
hydraulic conductivity of soils near the ground surface can be measured in hand-dug pits or cased
holes. At greater depths the hydraulic conductivity can be determined in the borings used for sam-
pling, provided that the borehole is cased and the hydraulic conductivity test will not afect sampling
of an important soil layer immediately below the test level. Table 6.22 summarizes various meth-
ods for measurement of hydraulic conductivity in the feld. Additional guidance in selecting feld
test methods is provided in ASTM D 4043, Standard Guide for Selection of Aquifer-Test Method in
Determining of Hydraulic Properties by Well Techniques.
The most commonly performed feld hydraulic conductivity test involves a sudden change (increase
or decrease) of water level in a borehole and measurement of the response in terms of water level
versus time. The test procedure and methods of analysis are presented in ASTM D 4044, Stan-
dard Test Method (Field Procedure) for Instantaneous Change in Head (Slug) Tests for Determin-
ing Hydraulic Properties of Aquifers. Hydraulic conductivity values determined by this procedure
ofen fail to correspond well with values predicted from laboratory testing due to: (1) characteristic
diferences (e.g., gradation and density) between the soils tested in the feld and in the laboratory, (2)
clogging of the boring face by soil particles suspended in the water, (3) varying directional hydraulic
conductivity of layered soil that cannot be duplicated in the laboratory with disturbed soil samples,
or (4) failure to conduct the feld test in saturated soils, resulting in measurement of the rate of satura-
tion rather than hydraulic conductivity. Because feld hydraulic conductivity tests inherently account
for the efects of geologic variations, they are generally more representative of in-situ conditions than
laboratory tests. However, the evaluation and interpretation of the test data require knowledge of the
test conditions and of the possible efects of these test conditions on the results.
The feld hydraulic conductivity test for rock is similar to that for soil, as described in ASTM D 4630,
Standard Test Method for Determining Transmissivity and Storage Coefcient of Low-Permeability
Rocks by In Situ Measurements Using the Constant Head Injection Test. The test is performed in a
rock boring using water pumped under pressure from the ground surface. Two types of tests can be
performed: a single-packer test or a double-packer test.
In the single-packer test, a pipe is inserted into a boring with a packer at the lower end of the pipe.
The packer is expanded mechanically or pneumatically from the ground surface to seal the annulus
between the walls of the boring and the pipe, and water is pumped down the pipe into the boring
below the packer. Since the depth and diameter of the hole below the packer, the applied water pres-
sure, and the rate of fow of water through the system are known, the average hydraulic conductivity
of the rock below the packer can be calculated.
< PREVIOUS VIEW
6-68
Chapter 6
MAY 2009
TABLE 6.22 FIELD METHODS FOR MEASUREMENT OF HYDRAULIC CONDUCTIVITY
Test Method Applicable Soils Reference
Various feld methods Soil and rock aquifers ASTM D 4044
Pumping tests Drawdown in soils ASTM D 4044
Slug tests Soils at depth ASTM D 4044
Constant head injection Low-hydraulic-conductivity rocks ASTM D 4630
Pressure pulse technique Low-hydraulic-conductivity rocks ASTM D 4630
(ADAPTED FROM MAYNE ET AL., 2002)
In the double-packer test, a selected zone within the boring is tested by placing one packer at the
botom and another packer at the top of the test zone. Water is then pumped through the pipe into the
annular space between the packers. The hydraulic conductivity is computed by the same procedure
as for the single-packer test.
During hydraulic pressure testing of rock, the water pressure applied to the test zone must not exceed
the pressure caused by the weight of overburden above the test zone. Excess pressures may force
water into joints, bedding planes or fractures and cause additional fracturing of the rock by lifing the
overburden. This jacking of the rock can seriously increase the amount of leakage that will occur
later and may also decrease the stability of the rock mass and its ability to resist the loads applied by
an embankment or other surface loading.
Hydraulic pressure testing to measure rock hydraulic conductivity is an essential part of subsurface
exploration where an impoundment is planned and where groundwater leakage could create unsafe
uplif pressures or piping of the embankment or foundation soils. In addition to posing a threat to
the safety of the embankment and natural slopes, excessive leakage can increase stream and ground-
water pollution down gradient from the refuse disposal facility. The hydraulic conductivity values
obtained from hydraulic pressure testing allow the prediction of quantities and locations of leakage
from the impoundment. If rock zones where excessive leakage could occur are observed, grouting
of the rock formations may be required. Houlsby (1990) and Weaver and Bruce (2007) discuss proce-
dures and materials for grouting rock formations to reduce water fow.
6.4.3.11 Groundwater-Level Measurements
For new embankments, determination of the groundwater level in the planned construction area is
important to construction requirements, particularly where excavations are planned. An understand-
ing of the groundwater regime is essential in determining the direction and rate of possible seepage
from an impoundment and may aid in estimating the overall hydraulic conductivity of the founda-
tion materials.
Using piezometers to measure the phreatic surface level within an existing embankment is ofen
the most important of the feld tests used for evaluation of an existing coal refuse disposal facility,
whether there is an impoundment or not. ASTM D 4750, Standard Test Method for Determining
Subsurface Liquid Levels in a Borehole or Monitoring Well (Observation Well), describes proce-
dures that should be followed in measuring groundwater levels. Piezometers or standpipes should
generally be installed in exploration boreholes. Common techniques for installing piezometers and
standpipes are presented in Chapter 13 along with monitoring procedures. The accuracy of data
obtained from piezometers is directly related to the care taken in their installation. Therefore piezom-
eter installation should always be under the supervision of a qualifed engineer or geologist.
< PREVIOUS VIEW
6-69
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
6.4.3.12 Water Flow and Quality Tests
In addition to phreatic surface level measurement with piezometers, valuable information for an
existing impounding refuse disposal facility can be gained from: (1) monitoring fows from nearby
springs or seep areas with weirs and (2) evaluating water quality aspects of these fows. As an exam-
ple, measuring the volume of fow from a spring below an impoundment during both wet and dry
periods, and as the impoundment level changes, will provide an indication of the rate of seepage
from the impoundment and the efect on the overall groundwater system. Likewise, simple feld
measurements of temperature and acidity of seepage will allow comparison with similar measure-
ments for the water in the impoundment and/or the infow groundwater. Flow measurements and
related water quality data are generally less important in the design of a new disposal facility, but
the data are useful for future evaluation of the efect of the impoundment on the local and regional
groundwater and surface water quality.
The type of weir to be used for fow monitoring and related construction requirements are deter-
mined by the magnitude of fow to be measured and the type of material in which the weir will be
placed. In the case of very small seeps, the fow volume can be estimated simply by observation. If the
fow initially passes over a natural weir, such as a rock outcrop or through an existing pipe, esti-
mates can be made without installing special instrumentation. Construction of weirs and methods for
accurately measuring fows from weirs and pipes are discussed in detail in Chapter 13.
Field testing of water can be performed using portable equipment to obtain indicator parameters
such as pH, specifc conductance and temperature, as well as some other mining-related parameters.
Where measurements of additional constituents and characteristics of seepage water (including sul-
fate, chloride, iron, manganese, acidity, alkalinity, and dissolved and suspended solids) are desirable,
additional water sampling and laboratory testing can be performed.
6.4.3.13 Backflling of Boreholes
Boreholes at coal refuse disposal facility sites should not be lef open, particularly if they are located
beneath embankments or impoundments, or if they can potentially provide a pathway for fuid fow
that is detrimental to site safety. Open boreholes can be backflled with drill cutings, cement grout,
bentonite, and other materials depending on the objectives. Where there is no concern related to
fuid migration or the impact of seepage on ground conditions, backflling with cutings or other
materials may be acceptable. As a practical mater, boreholes in cohesionless soils may collapse
when not supported, and it may not be possible (or necessary) to backfll such boreholes. Grout-
ing of an open borehole is generally performed for the purpose of constructing a barrier that will
prevent the vertical migration of fuids between geologic units. At active mine sites, the purpose
may be to maintain the barrier between the mine workings and other strata. Materials employed
for backflling boreholes include cement, bentonite slurries, dry bentonite, and fast-seting cement
grouts. Placement can be accomplished by tremie, pumping, and surface pouring. Site-specifc
considerations for a grouting program include: (1) whether to grout, (2) where to grout, and (3) the
method of deployment.
ASTM D 5299, Decommissioning of Ground Water Wells, Vadose Zone Monitoring Devices, Bore-
holes, and Other Devices for Environmental Activities, presents guidance on methods and materi-
als for closing of boreholes. While this standard is primarily oriented to environmental activities, it
can be used to decommission boreholes where no contamination is observed. Atributes of common
borehole plugging materials are discussed in the standard.
Grouting of boreholes that penetrate mines requires special provisions for supporting the borehole
plug above the mine void (and potentially in the mine foor, if the boring is advanced through the
mine). Frequently, sacrifcial casing is lef in the mine void to support the plug, although a grout
basket has been used to allow sealing the borehole and to permit grouting.
< PREVIOUS VIEW
6-70
Chapter 6
MAY 2009
6.4.4 Geophysical Methods
Applied geophysics is a rapidly evolving feld, and the applicability of geophysical techniques to
coal refuse disposal facilities will continue to advance with respect to the aspects of data gather-
ing, processing, interpretation and presentation of the geophysical data. Two basic deployments of
geophysics are available: (1) surface surveys and (2) measurements from boreholes. Airborne geo-
physical techniques are not discussed herein, as their application in terms of identifying features of
interest with respect to coal refuse disposal facilities is still in the experimental stage. Nevertheless
it is worth noting that in some cases airborne electromagnetic (EM) surveys have been used to map
fooded, abandoned coal workings (Love et al., 2005), and aeromagnetic surveys have been used
for many years to map abandoned well casings, which can be a signifcant hazard to coal mining
(Frischknecht et al., 1985).
Numerous sources of information related to geophysics are available in the general literature. A good
source of information is the Environmental and Engineering Geophysical Society (EEGS) in Denver,
Colorado, which annually holds the Symposium on the Application of Geophysics to Engineering
and Environmental Problems (SAGEEP). The SAGEEP proceedings are an excellent source of up-to-
date information on the application of engineering geophysics to the types of problems that could be
encountered at a coal refuse disposal facility. Other comprehensive compilations of geophysical tech-
niques for subsurface exploration for engineering applications include Ward (1990), USACE (1995a),
Sabatini et al. (2002), Wightman et al. (2003), and Sirles (2006). The Federal Highway Administration
(FHWA) presents summaries of geophysical techniques at their web site. This material is substan-
tially based on the USACE (1995a) work.
MSHA (2008) is a summary report of mine void detection demonstration projects that were per-
formed to evaluate the use of geophysical techniques for detection of underground mine workings.
These projects include actual feld demonstrations of void detection at mine sites using seismic meth-
ods, electrical resistivity, electromagnetics, and radar.
The following ASTM standards provide guidance for conducting geophysical exploration:
D 6429, Standard Guide for Selecting Surface Geophysical Methods
D 6430, Standard Guide for Using the Gravity Method for Subsurface Investigation
D 6431, Standard Guide for Using the Direct Current Resistivity Method for Sub-
surface Investigation
D 6432, Standard Guide for Using the Surface Ground Penetrating Radar Method
for Subsurface Investigation
D 5753, Standard Guide for Planning and Conducting Borehole Geophysical Logging
D 6639, Standard Guide for Using the Frequency Domain Electromagnetic Method
for Subsurface Investigations
D 7128, Standard Guide for Using the Seismic-Refection Method for Shallow Sub-
surface Investigation
D 6820, Standard Guide for Use of the Time Domain Electromagnetic Method for
Subsurface Investigation
While ASTM guides provide useful background information on geophysical techniques, they may be
dated in terms of defning procedures for data acquisition, processing, interpretation and presenta-
tion. For example, ASTM D 6431, Standard Guide for Using the Direct Current Resistivity Method
for Subsurface Investigation, discusses the technique in terms of acquisition with a four-electrode
system and processing of one-dimensional data sets with computer programs developed in the 1970s.
Modern resistivity surveys are commonly conducted with multi-electrode arrays, and the data are
routinely processed and interpreted in terms of 2D profles or 3D blocks.
< PREVIOUS VIEW
6-71
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
6.4.4.1 Surfcial Geophysical Techniques
In general terms, surface geophysical testing can be used to create a general image of subsurface
conditions that can be checked by intrusive means such as borings or test pits. Data from geophysi-
cal testing should always be correlated with information from direct methods of exploration. Ofen,
a combination of geophysical and direct exploration methods provides the best approach for inter-
preting subsurface conditions. When conducted at the outset of a subsurface investigation program,
geophysical exploration can prove to be cost-efective through reduction of the number of borings
needed for characterization of a site.
Conventional applications of surface geophysics include: (1) establishing the stratifcation of subsur-
face materials, (2) mapping the top of bedrock, depth to groundwater, and extent and quantity of soil
deposits, and (3) determining the rippability of hard soil and rock. Over the past several years some
improvements in traditional geophysical techniques have enhanced capabilities for the detection of
abandoned mine workings and the presence of karst-related voids.
As summarized in Table 6.23, surface geophysical testing ofers some advantages and limitations that
should be understood before a technique is selected for a specifc application.
Table 6.24 presents an overview of surfcial geophysical methods and techniques in relation to the
physical parameters measured and exploration objectives for coal refuse disposal facilities. The fol-
lowing text describes the most commonly used geophysical techniques listed in the table.
6.4.4.1.1 Seismic Refraction
The seismic refraction technique consists of measuring the frst arrival of P- and/or S-waves at vary-
ing distances from a seismic source. The most common application of this technique is the determi-
nation of the depth to bedrock, which requires that the upper layer velocity (soil/weathered or sof
rock) is less than that of the lower layer (competent rock). If a high-velocity surface layer is present,
the technique is not efective. For this reason, the technique is generally not applicable for detection
of mine voids.
TABLE 6.23 ADVANTAGES AND LIMITATIONS OF SURFACE GEOPHYSICAL TESTING
Advantages Limitations
1. Many geophysical tests are non-invasive and thus offer
signifcant benefts in cases where conventional drill-
ing, testing, and sampling are diffcult (e.g., deposits of
gravel, talus deposits or access constraints).
2. Geophysical testing generally covers a relatively large
area, thus providing the opportunity to characterize large
areas with relatively limited testing. It is particularly well
suited to projects having large areal extent (e.g., new
refuse disposal facility).
3. Some types of geophysical measurement can assess
the characteristics of soil and rock at very small strains
(0.001%), thus providing information on truly elastic
properties.
4. Most geophysical methods are relatively inexpensive
when considering cost relative to the relatively large
areas over which data can be obtained.
5. A properly performed geophysical survey can reduce the
number of borings required for site characterization.
1. Geophysical testing, when applied to locating
changes in soil and/or rock properties, will be
effective only if a target of interest has a physi-
cal contrast with the surrounding ground.
2. Results are generally interpreted qualitatively.
Useful results can only be obtained by an ex-
perienced engineer or geologist who is familiar
with the particular testing method.
3. Specialized equipment is required, as com-
pared to more conventional subsurface explo-
ration methods.
4. Results from surface geophysical testing
should be validated using direct methods of
exploration such as borings.
(ADAPTED FROM SABATINI ET AL., 2002)
< PREVIOUS VIEW
6-72
Chapter 6
MAY 2009
Seismic waves are usually created using a sledge hammer for depths up to about 50 feet and with
explosives for depths up to about 100 feet. Other sources such as vibrators are sometimes used. Ini-
tially, the seismic waves travel solely through the soil to arrive at geophones (vibration transducers)
located away from the source. The seismic waves also propagate through the overburden and refract
along the bedrock surface. While the waves are traveling along this surface, they continually refract
seismic waves back to the ground surface that are also detected by the geophones and recorded
with a seismograph. The result is the generation of travel-time curves, as shown in Figure 6.15. The
method can ofen be a low-cost (compared to boreholes) method of bedrock mapping and overbur-
den estimation. When measurements are obtained with a high degree of redundancy, the result is a
reliable acoustic image of the subsurface in terms of layers and variations of seismic velocity within
the individual layers.
Seismic refraction data can also be useful for determining the rippability of rock materials using heavy
construction equipment. Companies such as Caterpillar have prepared graphs comparing rippability
versus P-wave velocity for various equipment types, an example of which is shown in Figure 6.16.
The design of a seismic refraction survey involves locating the profles where data are desired and
determining the length of the array of geophones (the geophone spread) and the geophone spacing.
TABLE 6.24 SURFACE GEOPHYSICAL METHODS IN RELATION TO
TYPICAL INVESTIGATION OBJECTIVES
Geophysical Method
Dependent Physical
Property
Applications (see key below)
1 2 3 4 5 6 7 8 9 10
Seismic refraction Elastic moduli; density P P P S X X X M M S
Seismic refection Elastic moduli; density S M S S X P X X P P
Resistivity Resistivity P X X P S P X P M S
Spontaneous potential (SP) Potential differences X X X X P X X X X X
Electromagnetics (EM) Conductivity; inductance M X X P S S P S P S
Ground penetrating radar Permittivity; conductivity M X X X M S M X S S
Gravity Density X X X X X P X X X X
Magnetics Magnetic susceptibility X X X X X M P X X X
Key to Techniques and
Applications
Technique Applicability Applications
P primary technique 1 depth to bedrock
6 location of mine workings or
other subsurface voids
S secondary technique 2 rippability of rock and hard soil 7 abandoned well detection
M may be used but probably not
the best approach
3 elastic properties of coal refuse,
soil and rock
8 variations of coal refuse
composition
X not applicable 4 hydrogeological investigations
9 location of faults/fractures/
geologic structures
5 location of seepage pathways in
a dam
10 determination of soil/bedrock
stratigraphy
(ADAPTED FROM SIRLES, 2006)
< PREVIOUS VIEW
6-73
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
The length of the spread depends on the required depth of penetration. As a rule of thumb, the length
of the spread needs to be at least four times the depth of interest. Seismic noise originates from
ambient vibrations that can be caused by sources such as trafc or wind. Obtaining multiple record-
ings from the same location and summing (stacking) the results or covering the geophones with sand
bags can reduce the efect of this type of interference. Improved results with the seismic refraction
technique can also be obtained when multiple refractions from the same refractive interface gener-
ated by multiple shots are received at a given geophone. This multiplicity of data is obtained by
using a single geophone spread with multiple shot points. Commercial engineering seismographs
designed for the seismic refraction technique usually allow for simultaneous recording from spreads
with either 12 or 24 geophones, although some commercial equipment will allow for recording with
96 or more geophones.
There are several steps in the processing and interpretation of seismic refraction data. Firstly, it is
necessary to pick the frst arrival time for P-wave analysis and to tabulate this information. If the
purpose is to obtain the S-wave velocity, then it is necessary to pick the onset of the S-wave arrival,
which is more difcult than picking the P-wave arrival, because the S-waves lie within the wave train
of the P-wave. Analytical procedures for processing the data commonly use either the generalized
reciprocal method (GRM) or the delay-time method for the inversion and interpretation of refraction
data. Both methods are suitable for resolving multilayer profles with structural complexities (dips
up to about 20 degrees). The acquisition of redundant data optimizes the accuracy of both interpre-
tive techniques. The end result is a cross section of the ground with velocities assigned to each layer,
as depicted in Figure 6.15.
FIGURE 6.15 TYPICAL SEISMIC REFRACTION DATA WITH INTERPRETATION
0 200 400 600 800 1000 1200 1400
200
150
DISTANCE (FT)
100
50
0
T
I
M
E
(
M
S
E
C
)
TIME-DISTANCE PLOT
UNSATURATED SEDIMENTS
BEDROCK
DISTANCE (FT)
0 200 400 600 800 1000 1200 1400
D
E
P
T
H
(
F
T
)
0
50
100
150
200
250
INTERPRETED CROSS SECTION
BORING BORING
SATURATED SEDIMENTS
(HAENI, 1988)
FIGURE 6.15 TYPICAL SEISMIC REFRACTION DATA WITH INTERPRETATION
FIGURE 6.15 TYPICAL SEISMIC REFRACTION DATA WITH INTERPRETATION
< PREVIOUS VIEW
6-74
Chapter 6
MAY 2009
When S-wave arrivals are picked from a seismic refraction record, it is possible to use the S-wave
velocity to calculate the elastic properties of the identifed layers. In practice, it can be difcult to
identify S-waves, even when a horizontally-polarized source and horizontal geophones are used.
Although there are numerous published examples (Johnson and Clark, 1992; Ellefsen et al., 2005) of
the successful calculation of S-wave velocity from refraction data, in practice the most reliable meth-
ods for measuring elastic properties of the subsurface are from crosshole or downhole surveys, as
discussed in Section 6.4.4.2.
6.4.4.1.2 Seismic Refection
Application of the seismic refection technique involves measuring the travel time required for a
seismic wave generated at or near the surface (P-wave or S-wave, depending on the survey setup)
to return to surface or near-surface geophones afer refection from acoustic interfaces between sub-
surface layers (Figure 6.17). Seismic refection is the most powerful of all geophysical techniques
for mapping subsurface layering and is by far the most commonly applied method for oil and gas
exploration. It is also the most sophisticated of all geophysical methods and requires highly spe-
cialized equipment and processing sofware for its successful application. For this reason, the tech-
TOPSOIL
CLAY
IGNEOUS
GRANITE
BASALT
SEDIMENTARY
SHALE
SANDSTONE
SILTSTONE
CONGLOMERATE
BRECCIA
LIMESTONE
METAMORPHICS
SCHIST
SLATE
OTHER
COAL
IRON ORE
0 5 10 15
SEISMIC VELOCITY (FT/SEC x 1000) RIPPABLE
UNRIPPABLE
FIGURE 6.16 CORRELATION OF RIPPABILITY WITH P-WAVE
VELOCITY FOR CATERPILLAR D9
(CATERPILLAR, 2008)
FIGURE 6.16 CORRELATION OF RIPPABILITY WITH P-WAVE VELOCITY
FOR CATERPILLAR D9
FIGURE 6.16 CORRELATION OF RIPPABILITY WITH P-WAVE VELOCITY
FOR CATERPILLAR D9
< PREVIOUS VIEW
6-75
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
nique requires highly experienced practitioners, and no atempt has been made herein to describe the
details of the data acquisition, processing and interpretation. Seismic refection is not commonly used
in environmental and engineering projects because of its relatively high cost. Contrary to most other
geophysical methods, shallow seismic refection studies are more expensive than the deep surveys
conducted for oil and gas exploration because of the need for closely spaced measurements. Never-
theless, the method ofers the potential for defning subsurface structure beter than other methods.
In cases where it is important to know the location of faults or other lithologic breaks or when the
target is a deep abandoned mine working, the high-resolution seismic refection technique may be
the only practical means to obtain the desired data.
Seismic refection has been applied to mapping the continuity of coal seams in advance of longwall
mining, particularly in Europe where mines are commonly at depths greater than 1000 feet and it is
difcult and expensive to characterize a coal seam with borings (DAppolonia, 1982). The method has
also been successfully applied to the mapping of mine voids (Clark et al., 1994; Johnson et al., 2002),
but the experience base is limited and few practitioners are equipped to properly conduct this type
of survey. An example of a high-resolution seismic refection survey performed over shallow mine
workings is presented in Figure 6.18.
FIGURE 6.17 SEISMIC REFLECTION PRINCIPLE AND SCHEMATIC
OF REFLECTION DATA RECORD
GEOPHONES GROUND SURFACE
LAYER 1
LAYER 2
ENERGY SOURCE
C
L
C
L
X X
t = 0 t = 0
t
t
F
I R
S
T
A
R
R
I V
A
L
S
F
I R
S
T
A
R
R
I V
A
L
S
S
E
C
O
N
D
A
R
R
I V
A
L
S S
E
C
O
N
D
A
R
R
I V
A
L
S
(JOHNSON AND CLARK, 1992)
FIGURE 6.17 SEISMIC REFLECTION PRINCIPLE
AND SCHEMATIC
OF REFLECTION DATA RECORD
FIGURE 6.17 SEISMIC REFLECTION PRINCIPLE AND SCHEMATIC
OF REFLECTION DATA RECORD
< PREVIOUS VIEW
6-76
Chapter 6
MAY 2009
FIGURE 6.18 SEISMIC REFLECTION SURVEY PROFILE OVER
ABANDONED COAL MINE WORKINGS
10 50 100 150 200 240
0
50
100
150
200
BORING PENETRATING
MINE WORKINGS
POSSIBLE MINE
WORKING
CHANNEL FILL
CUTTING COAL
SEAM
INTACT COAL SEAM
DISTANCE (FT)
D
E
P
T
H
(
F
T
)
NOTE: DATA WERE OBTAINED USING
VIBRATORY S-WAVE SOURCE. (JOHNSON ET AL., 2002)
FIGURE 6.18 SEISMIC REFLECTION SURVEY PROFILE OVER ABANDONED
COAL MINE WORKINGS
6.4.4.1.3 Electrical Resistivity
The purpose of electrical resistivity surveying is to determine the subsurface resistivity distribu-
tion by making measurements at the ground surface. From these surface measurements, the true
resistivity of the subsurface can be estimated, and variations or anomalies in the observed resistiv-
ity may indicate limits of surface deposits such as coal refuse, the bedrock surface, or fooded mine
workings. Resistivity is typically described in units of ohm-meters or ohm-feet. Ground resistivity
is afected by various physical parameters such as the mineral and fuid content, porosity, and the
degree of saturation.
The measurement of electrical resistivity is normally performed using four electrodes, two that
induce current into the ground and two that measure potential diference (voltage). Figure 6.19
provides some examples of electrode confgurations commonly used for electrical resistivity mea-
surements. Electrical resistivity surveys have been performed for many decades as part of hydrogeo-
logical, mining and geotechnical investigations, but the use of this technique has recently increased
due to improvements in both data acquisition and data processing technologies. Multi-electrode
systems have greatly improved the efciency of data acquisition, as measurements can now be made
automatically without moving the current insertion and voltage measurement points. Also, the DC
resistivity method had been limited by the need to perform complex calculations to model subsur-
face electrical properties. With the availability of high-speed personal computers and improved 2D
and 3D processing sofware (Gan; 2004, 2005), the technique has seen increased interest from the
mining industry, including as a means for detection of subsurface openings.
FIGURE 6.18 SEISMIC REFLECTION SURVEY PROFILE OVER ABANDONED
COAL MINE WORKINGS
< PREVIOUS VIEW
6-77
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
FIGURE 6.19 COMMONLY USED ELECTRODE CONFIGURATIONS
FOR GROUND ELECTRICAL MEASUREMENTS
V I
a na a
DIPOLE-DIPOLE CONFIGURATION
V
WENNER CONFIGURATION
a a a
M N A B A B M N
V
SCHLUMBERGER CONFIGURATON
a na na
V
A M N B
POLE-DIPOLE CONFIGURATION
I
M N
a
A B
na
I
I
NOTE: A AND B ARE CURRENT (I) INSERTION POINTS;
M AND N ARE VOLTAGE (V) MEASUREMENT POINTS. (ADAPTED FROM GAN, 2004)
FIGURE 6.19 COMMONLY USED ELECTRODE
CONFIGURATIONS
FOR GROUND ELECTRICAL MEASUREMENTS
At coal refuse disposal sites, electrical resistivity surveys have proven useful for estimating the
amount of coal refuse disposed at existing facilities by enabling location of the base of the refuse.
Furthermore, the method also has the potential to diferentiate zones with varying physical charac-
teristics related to coal content within the coal refuse. An example of the application of electrical resis-
tivity to generate profles across a fne coal refuse deposit is presented in Figure 6.20. In the fgure, the
depth to the base of the coal refuse, which is in contact with bedrock, is clearly visible (yellow line),
and variations in resistivity within the fne coal refuse can be atributed to physical diferences such
as the level of coal content.
D
E
P
T
H
(
F
T
)
DISTANCE (FT)
FIGURE 6.20 RESISTIVITY PROFILE ACROSS FINE COAL REFUSE IMPOUNDMENT
0 100 200 300 400 500 600
10
20
30
40
50
60
70
80
RESISTIVITY (OHM-M)
5 25 150 1200 4500
FIGURE 6.20 RESISTIVITY PROFILE ACROSS FINE COAL REFUSE IMPOUNDMENT
FIGURE 6.19 COMMONLY USED ELECTRODE CONFIGURATIONS
FOR GROUND ELECTRICAL MEASUREMENTS
FIGURE 6.20 RESISTIVITY PROFILE ACROSS FINE COAL REFUSE IMPOUNDMENT
< PREVIOUS VIEW
6-78
Chapter 6
MAY 2009
Another application of the electrical resistivity method at coal refuse sites is the detection of aban-
doned mine workings, as shown in Figure 6.21. The ability to detect voids is enhanced when the void
has a physical contrast with the surrounding rock. If the void is empty (no water), it will be dif-
cult to detect with electrical measurements. Air does not transmit an electrical current, and, unless the
coal has an unusually low resistivity, it may be difcult to distinguish a void. The resistivity contrast
between fooded mine voids and typical coal will approach two orders of magnitude (Johnson, 2003),
thus allowing for detection of mine workings as resistivity lows. Project experience with electrical resis-
tivity demonstrates that commercially available technology can be efective, especially for the detection
of fooded mine workings at depths up to about 100 feet (Figure 6.21). DAppolonia (2006) conducted
a demonstration project for MSHA to illustrate the application of the method at the perimeter of an
impoundment where abandoned workings in the 40- to 60-foot-depth range contained limited water.
For deeper workings, the method has the potential to be efective, but theoretical models and practical
experience indicate that the target size/depth ratio needs to be favorable and that the length of the resis-
tivity profle required for acquiring deep images is ofen limited by surface interference. Therefore, the
method is usually most efective for mine subsidence applications.
6.4.4.1.4 Spontaneous Potential (SP)
The spontaneous or self-potential (SP) method consists of measuring naturally occurring electrical
potentials (voltage diferences) in the subsurface. One of the sources of these electrical potentials is
the movement of water through a porous medium, which produces electro-fltration or streaming
potentials. These potentials can be used for the evaluation of seepage (Figure 6.22). As water fows
through a capillary system, it collects and transports positive ions from surrounding materials. The
positive ions accumulate at the exit point of the capillary, leaving a net positive charge. The untrans-
ported negative ions accumulate at the entry point of the capillary, leaving a net negative charge. If
the streaming potentials developed by this process are of sufcient magnitude to be measured, the
entry point and the exit point of zones of concentrated seepage can be determined due to the negative
and positive (respectively) SP anomalies.
SP is measured with a pair of non-polarizing electrodes and a high-impedance voltmeter. One of
the electrodes is placed in the ground at a convenient location and remains in place throughout the
survey. This is referred to as the base electrode. It is connected to a multi-meter via an insulated,
single-conductor wire mounted on a reel. This wire may be hundreds of meters long. The second
FIGURE 6.21 ELECTRICAL SURVEY OF FLOODED MINE WORKINGS
DISTANCE (FT)
E
L
E
V
A
T
I
O
N
(
F
T
)
ZONE WHERE PROFILE
CROSSES EDGE OF WORKINGS
ZONE WHERE PROFILE
CROSSES WORKINGS
(JOHNSON, 2003) RESISTIVITY (OHM-M)
0 500 1000 1500
0 100 200 300 400 500 600
1360
1400
1440
1480
BASE OF LOWER
KITTANNING COAL
FIGURE 6.21 ELECTRICAL SURVEY OF FLOODED MINE WORKINGS
FIGURE 6.21 ELECTRICAL SURVEY OF FLOODED MINE WORKINGS
< PREVIOUS VIEW
6-79
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
electrode, or measuring electrode, the reel, and the multi-meter are then moved from point to point
in a planned grid. At each point of the grid, the electrical potential between the base and the measur-
ing electrode is recorded.
If the potential diferences are ploted in profle or contoured to identify zones with negative poten-
tial, areas of seepage can be identifed. Additional information and examples of SP surveys used to
locate seeps in embankment dams and tailings impoundments are provided by Butler et al. (1989),
Butler et al. (1990), Brub (2004), Song et al. (2005), and Mainali (2006). Recent advances in the SP
technique allow for predictive modeling of the SP anomalies associated with seepage, enhancing the
interpretability of SP results (UBC-GIF, 2005; Brub, 2004).
6.4.4.1.5 Electromagnetics (EM)
Similar to electrical resistivity measurements, electromagnetic methods (EM) allow mapping of
the distribution of subsurface electrical properties, except that the EM methods are designed for
measurement of variations in conductivity, not resistivity. Resistivity and conductivity are difer-
ent parameters related to the same physical property and are simply the inverse of one another.
As noted above, the unit most commonly used to measure ground resistivity is the ohm-meter
(ohm-m). The term mho refects its inverse relationship to the ohm, but was discontinued in
favor of the term siemen (symbol S) in the late 1970s. The corresponding unit of conductivity is
the inverse of an ohm-meter, referred to as a mho (or siemen) per meter, or S/m. The most common
unit of conductivity is the S/cm, which is 0.0001 S/m. With this conversion, 10,000 S/cm = 1 S/m
= 1 ohm-m.
FIGURE 6.22 SCHEMATIC ILLUSTRATION OF USE OF THE SP METHOD
TO IDENTIFY AN AREA OF EMBANKMENT SEEPAGE
MULTIMETER
REFERENCE
ELECTRODE
+
_
0
P
O
T
E
N
T
I
A
L
DISTANCE
IDENTIFIED
SEEPAGE AREA
NON-POLARIZABLE
ELECTRODE LOCATIONS
SEGMENT OF WATER-
IMPOUNDING EMBANKMENT
(ADAPTED FROM BUTLER ET AL., 1989)
FIGURE 6.22 ILLUSTRATION OF USE OF SP METHOD TO IDENTIFY AN AREA OF SEEPAGE
FIGURE 6.22 ILLUSTRATION OF USE OF SP METHOD TO IDENTIFY AN AREA OF SEEPAGE
< PREVIOUS VIEW
6-80
Chapter 6
MAY 2009
An advantage of all EM systems as compared to the resistivity method is that it is not necessary to
insert electrodes in the ground and thus the surveying is more rapid. Disadvantages of EM methods
are: (1) they are generally not as good as the DC resistivity method in resolving variations of electrical
properties with depth and (2) they are more subject to cultural interference from electrical lines and
metallic objects. For these reasons, EM methods are most commonly used to rapidly measure lateral
variations of soil electrical properties, as well as to delineate the distribution of metal objects.
Electromagnetic (EM) techniques can be grouped into active methods, where an active EM signal is
induced in the ground by human activity, and passive systems, where measurements are made of
natural variations of the earths EM feld. Active systems are further grouped into frequency domain
and time domain. Passive systems include very low frequency (VLF), and magnetotelluric methods.
McNeill (1990) provides a discussion of various EM techniques.
EM methods have potential application for characterization of coal refuse sites (e.g., mapping
abandoned workings) as long as the workings are fooded. For example, time-domain EM (TDEM)
measurements have been used to map fooded workings, as shown in Figure 6.23. Where this
technique has been atempted over workings that are not fooded, the method was less success-
ful (MSHA, 2008). EM techniques that measure bulk ground conductivity are commonly used
at operating or abandoned refuse disposal facilities to determine the migration of contaminated
groundwater or to delineate the extent of waste deposition. EM techniques can also be used to
characterize variations in the physical properties of existing coal refuse deposits related to coal
content. Results of this type of survey with a commonly-used conductivity meter (Geonics EM-31)
over an existing coal refuse deposit are shown in Figure 6.24. The plan location of the resistivity
profle in Figure 6.20 is shown in Figure 6.24. In this example, the EM survey provides mapping
of the near-surface horizontal variations of the coal refuse, while the resistivity profle in Figure
6.20 shows vertical variations.
6.4.1.1.6 Ground Penetrating Radar (GPR)
Ground penetrating radar (GPR) has evolved over the past two decades into one of the most com-
monly applied techniques for imaging the shallow subsurface. The method ofers the highest resolu-
tion of geophysical techniques commercially available today. In many cases, the time required for
the acquisition of GPR profles is minimal, and subsurface profles can normally be generated in real
time, making this tool very cost-efective. GPR works best in non-conductive soils, such as dry sand
or sand saturated with fresh water.
The typical result of a GPR survey is a profle that presents radar wave amplitude as a function of
distance along the line and two-way travel time. To determine the depth to a refector, it is necessary
to know the average propagation velocity from the ground surface. The velocity of a radar pulse in
an earth material is dependent on the relative dielectric constant (
r
) of the material according to the
following relationship:
V = c / (
r
)
0.5
(6-5)
where:
V = velocity in propagating material (m/sec)
c = speed of light (m/sec)
r
= relative dielectric constant (dimensionless)
This velocity can sometimes be estimated from the characteristics of the subsurface lithology. Table
6.25 presents typical velocities in terms of two-way travel time (nanoseconds/meter) for various earth
materials along with their approximate relative dielectric constants.
< PREVIOUS VIEW
6-81
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
Until the advent of commercial systems with separate transmiting and receiving antennas, depth
estimation based on subsurface material properties or from observations from refectors of a known
depth was the only means to interpret a GPR profle. Modern systems with the ability to record refec-
tions at varying distances from the transmiting antenna allow for the calculation of the subsurface
velocity profle by means of a normal moveout (NMO) correction based on hyperbolic refections
from subsurface features.
Depth of penetration depends on the selection of an appropriate antenna frequency. An antenna
frequency of one gigahertz would be suitable for mapping rebar in concrete, but would only have at
most a few feet of penetration in typical soil. Most GPR surveys in soil use antennas with frequencies
FIGURE 6.23 DELINEATION OF FLOODED MINE WORKINGS WITH TDEM METHOD
0 100 200
1300
1350
1500
1400
1450
300 400 500 600
E
L
E
V
A
T
I
O
N
(
F
T
)
DISTANCE (FT)
BASE OF LOWER
KITTANNING COAL
ZONE WHERE PROFILE CROSSES
EDGE OF WORKINGS
ZONE WHERE PROFILE
CROSSES WORKINGS
FIGURE 6.23 DELINEATION OF FLOODED MINE WORKINGS WITH TDEM METHOD
FIGURE 6.24 RESISTIVITY OF COAL REFUSE FROM EM-31 MEASUREMENTS
PIT
POND
R
O
A
D
R
E
S
I
S
T
I
V
I
T
Y
P
R
O
F
I
L
E
PILE
-700 -600 -500 -400 -300 -200 -100 0 +100
EASTING (FT)
N
O
R
T
H
I
N
G
(
F
T
)
0
-100
-200
-300
-400
-500
GROUND
RESISTIVITY
(OHM-M)
0
10
20
30
100
1000
2500
FIGURE 6.24 RESISTIVITY OF COAL REFUSE FROM EM-31 MEASUREMENTS
FIGURE 6.23 DELINEATION OF FLOODED MINE WORKINGS WITH TDEM METHOD
FIGURE 6.24 RESISTIVITY OF COAL REFUSE FROM EM-31 MEASUREMENTS
< PREVIOUS VIEW
6-82
Chapter 6
MAY 2009
between about 100 and 400 MHz, with the greatest penetration achieved with the 100-MHz antenna,
but with a substantial loss of resolution as compared to the 400-MHz antenna.
Another factor afecting the depth of penetration of the GPR signal is atenuation. Atenuation is
caused by spreading and scatering losses, as well as electrical losses. Scatering and electrical losses
are due primarily to the conductivity of the subsurface materials, which in soils relates mainly to clay
and moisture content. In dry sand, penetration can reach as much as 50 to 70 feet. In wet, saturated
clay penetration may be as litle as 3 to 7 feet.
The main limitation of the GPR technique is depth of penetration under conditions commonly
encountered in areas with coal workings. The soils commonly encountered in coal mining areas are
clays weathered from the claystones associated with the sedimentary sequences that include the coal,
and these soils can severely restrict the efective penetration of the radar waves. Thus, use of GPR is
generally limited to the identifcation of near-surface features such as buried waste, pipes, etc. Nev-
ertheless, where abandoned mine workings are shallow, GPR can sometimes be used to detect these
workings. An example of a GPR record with identifed mine workings is shown in Figure 6.25.
6.4.4.1.7 Gravity
At mine sites the gravity method can detect shallow abandoned mine workings by measurement
of minute changes in the earths gravity feld resulting from the lack of near-surface mass associ-
ated with mine openings. The measurement of the gravity feld for this application is referred to as
microgravimetry and requires the use of specialized gravimeters with a sensitivity of one microgal
(approximately one billionth of the earths gravity feld). An air-flled mine void would in theory be
detectable with commercial equipment at a depth of about 30 feet. In practice, it is time-consuming
to acquire the data, and accurate elevation control is needed, and it is desirable to have a topographic
survey crew accompany the geophysicist to measure the precise elevation of the instrument at each
reading point. For a target as shallow as 30 feet, the surface width of the gravity anomaly is about
100 feet. Thus, a survey requires a signifcant amount of accessible space, which is ofen not avail-
TABLE 6.25 PHYSICAL PROPERTIES AND TYPICAL GROUND PENETRATING RADAR
VELOCITIES FOR COMMON EARTH MATERIALS
Material
Approximate
Conductivity
(mS/m)
Approximate Relative
Dielectric Constant
r
(dimensionless)
Two-Way
Travel Time
(sec 10
-9
/m)
Air 0 1 6.6
Fresh Water 10
-1
30 81 59
Fresh-Water Ice 10
-1
10 4 13
Permafrost 10
-2
10 4 11 13 15
Limestone 10
-6
1 6 8 22
Granite 10
-6
1 5.6 8 18.7
Dry Sand 10
-4
1 4 6 13 16
Saturated Sand (fresh water) 10
-1
10
2
30 32 36
Saturated Silt (fresh water) 10 l0
2
10 21
Saturated Clay (fresh water) 10
2
10
4
8 - 25 18.6 23
Average Dirt 10
-1
10
2
16 20 30
(BENSON ET AL., 1984)
< PREVIOUS VIEW
6-83
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
able. Furthermore, it is ofen difcult to correct the gravity data for variations caused by surrounding
topography, instrumental drif, and elevation. In particular, micro-topographic changes can signif-
cantly afect gravity readings. Unless the target is in a fat, open area and the depth does not exceed
about 40 to 50 feet, the gravity method will probably not be practical. Nevertheless, if the mine work-
ings are expected to be very shallow and air-flled, the gravity method is one of the few geophysical
methods that can provide conclusive evidence of the presence of a mine void. A theoretical gravity
response over an air-flled mine void is presented in Figure 6.26.
6.4.4.1.8 Magnetics
The primary application of the magnetic method in a coal refuse environment is the detection of metal,
including abandoned metal well casings. Measurements are made by an instrument called a magnetome-
ter, and the unit of magnetic intensity is the nanotesla (nT), sometimes referred to as a gamma. Diferences
in the normal value of the earths magnetic feld correspond to magnetic anomalies that can be measured
with a magnetometer. Surveys for well casings can be conducted from the ground or from the air, as previ-
ously noted. As shown in Figure 6.27, well casings produce very strong anomalies, detectable even when
the magnetometer is located at an elevation of 250 feet above the well casing. The potential for detecting
abandoned mines is minimal, unless mine openings are associated with metal, as might be the case if old
mine rails are present. Coal has a low magnetic susceptibility when compared to most other rocks. A void
in a coal seam, therefore, will not produce a signifcant disturbance to the earths natural magnetic feld,
but a sensitive magnetometer can detect old mine rails at a depth of several tens of feet.
6.4.4.2 Borehole Geophysical Techniques
Borehole logging includes numerous geophysical techniques involving the lowering of sensing
devices into a borehole and continuously recording physical parameters associated with the sur-
112
126
T
I
M
E
(
N
S
)
D
E
P
T
H
(
F
T
)
0
14
28
42
56
70
84
98
4
8
12
16
20
0 50 150 100 200
DISTANCE (FT)
FIGURE 6.25 GPR RECORD OF SHALLOW COAL MINE WORKINGS
KNOWN MINE TUNNEL
SUSPECTED MINE TUNNELS
AIR WAVES REFLECTED
FROM SURFACE OBJECTS
LEGEND
(JOHNSON ET AL., 2002)
FIGURE 6.25 GPR RECORD OF SHALLOW COAL MINE WORKINGS
FIGURE 6.25 GPR RECORD OF SHALLOW COAL MINE WORKINGS
< PREVIOUS VIEW
6-84
Chapter 6
MAY 2009
FIGURE 6.26 THEORETICAL RESPONSE OF GRAVITY GEOPHYSICAL METHOD
OVER 20-FOOT-DIAMETER, AIR-FILLED TUNNEL
LINE DISTANCE (FT)
G
R
A
V
I
T
Y
(
M
I
C
R
O
G
A
L
S
)
0
-10
-20
-30
-40
-50
-60
0 -100 -80 -60 -40 -20 20 40 60 80 100
30-FT DEPTH
90-FT DEPTH
60-FT DEPTH
APPROXIMATE LIMIT OF DETECTION WITH GRAVITY METHOD
NOTE: RESPONSE WOULD BE APPROXIMATELY
ONE-THIRD OF THE ABOVE FOR A WATER-
FILLED TUNNEL.
(JOHNSON ET AL., 2002)
FIGURE 6.26 THEORETICAL RESPONSE OF GRAVITY GEOPHYSICAL METHOD
OVER 20-FOOT-DIAMETER, AIR-FILLED TUNNEL
rounding rock, soil, pore fuids, or other physical parameters. The FHWA lists 23 borehole logging
techniques on their web site.
A general grouping of the most commonly applied borehole techniques is provided in ASTM D 5753,
Standard Guide for Planning and Conducting Borehole Geophysical Logging. This general group-
ing of borehole geophysical techniques is shown in Table 6.26, where a division is made in terms of:
(1) acoustic logs intended to determine ground variations related to seismic velocity, (2) electrical and
induction logs that identify lithologic and groundwater variations on the basis of resistivity/conductiv-
ity, (3) nuclear logs that relate to variations in natural or induced radioactivity, and (4) miscellaneous
techniques that defne the physical characteristics of boreholes and/or voids penetrated by boreholes.
For coal refuse facilities, most of the commonly used borehole geophysical methods are suit-
able for general site characterization in combination with conventional drilling and sampling.
Techniques with particular relevance to coal refuse disposal facilities are those that provide the
S-wave velocity as a function of depth (crosshole and downhole seismic methods), because this
information is useful for the evaluation of the potential for seismically-induced liquefaction.
Borehole logging and techniques that have potential for characterization of abandoned mine
workings (video, laser imaging, sonic imaging) are also directly applicable to coal refuse disposal
facility evaluations.
FIGURE 6.26 THEORETICAL RESPONSE OF GRAVITY GEOPHYSICAL METHOD
OVER 20-FOOT-DIAMETER, AIR-FILLED TUNNEL
< PREVIOUS VIEW
6-85
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
A borehole technique with the potential for imaging abandoned mine workings is borehole GPR.
Mine voids within approximately 10 f can be detected from borehole GPR. Research has shown that
crosshole GPR tomography can identify tunnels, but this is not a commonly applied technique and
the difculties of this technique in mapping coal mine voids is well described by the Colorado School
of Mines in an experimental mine detection study for MSHA (CSM, 2007).
Information on commonly applied general borehole geophysical techniques can be found in USEPA
(1993), USACE (1995a), Keys (1997), Sabatini et al. (2002), Wightman et al. (2003), and Sirles (2006).
The following discussion focuses on techniques for measuring S-wave velocity that are primarily
used for determining the seismic properties of coal refuse and soils for liquefaction analyses, but
these techniques may also be useful for characterizing mine voids.
6.4.4.2.1 Crosshole and Downhole/Uphole Seismic Surveys
Crosshole and/or downhole/uphole seismic testing in boreholes is conducted for determining soil
and rock properties (P- and S-wave velocities). The information obtained from these tests can be used
to compute shear modulus, Youngs modulus, and Poissons ratio for use in static/dynamic analyses.
The basic relationships are as follows:
E = V
p
2
[(1 + ) (1 2 ) / (1 )] (6-6)
G = V
s
2
(6-7)
where:
V
p
= compressional (P-wave) velocity (length/time)
V
s
= shear (S-wave) velocity (length/time)
E = Youngs modulus (force/length
2
)
G = shear modulus (force/length
2
)
= mass density of soil (mass/length
3
)
= Poissons ratio of soil (dimensionless)
If both V
p
and V
s
are known, the Poissons ratio of the soil can be determined from the following rela-
tionship between E and G:
G = E / 2 (1 +
) (6-8)
With borehole seismic surveys, one or more boreholes are drilled into the soil to the desired depth
of exploration. Wave sources and/or receivers (borehole geophones normally oriented to record both
horizontal and vertical components of wave motion) are then lowered into the boreholes. There are
three basic approaches to borehole seismic surveys:
Crosshole Survey In a crosshole survey, the energy source is located in one bore-
hole and detectors are placed in another borehole at the same depth as the energy
source. The energy source is usually a mechanical pulse instrument composed of a
stationary part and a hammer. The pulse instrument is held against the side of the
borehole by a pneumatic or hydraulic bladder. Travel times between the source and
receivers are measured, allowing determination of wave velocities.
Uphole Survey Geophones are laid out on the ground surface in an array around
the borehole. The energy source is set of within the borehole at successively decreas-
< PREVIOUS VIEW
6-86
Chapter 6
MAY 2009
FIGURE 6.27 MAGNETIC INTENSITY RECORDED AT VARIOUS HEIGHTS ABOVE WELL CASING
FIGURE 6.27 MAGNETIC INTENSITY RECORDED AT VARIOUS
HEIGHTS ABOVE WELL CASING
T
O
T
A
L
F
I
E
L
D
(
N
A
N
O
T
E
S
L
A
S
)
250 FT
200 FT
150 FT
100 FT
HEIGHT OF AIRPLANE
ABOVE GROUND
S N
0 5,000
DISTANCE (FT)
(FRISCHNECHT AND RAAB, 1984)
5,000
1
0
0
N
T
1
0
0
N
T
ing depths starting at the botom of the hole. The travel times from the source to the
surface are analyzed to determine the wave velocity versus depth.
Downhole Survey In a downhole survey, the energy source is located on the sur-
face and a detector (geophone) is placed in a borehole. The travel time is measured
with the geophone placed at progressively increasing depth, and a wave-velocity
profle is generated.
Crosshole seismic surveys involve measurement of the travel time of seismic energy transmited
between two or preferably three boreholes to derive information relative to the elastic properties
of the intervening materials. The travel times of the seismic waves are derived from the identifed
frst-arrivals of the P- and S-waves on the seismic trace for each shot-receiver position and are used
with the known distance (s) between the shot/receiver boreholes to calculate the apparent veloci-
ties (P- and S-wave) for each depth interval. The borings are usually cased and grouted to the sur-
rounding soil/rock. PVC casing is normally used for the tests, so that the casing is not a seismic
pathway. A typical feld setup for a crosshole seismic survey is shown in Figure 6.28.
Crosshole geophysical testing is described in ASTM D 4428, Standard Test Methods for Crosshole
Seismic Testing. Crosshole measurements are generally preferred to downhole measurements
because they provide higher resolution and greater accuracy. However, the distances between the
energy source and the detector must be measured precisely. An inclinometer survey is generally per-
formed in crosshole test boreholes to correct the data for deviation of the boreholes from vertical. To
calculate P- and S-wave velocity, the wave arrivals must be processed with a computer program that
accounts for situations where the waves may be refracted between the boreholes according to Snells
Law. The data are then used to develop vertical profles of the various elastic moduli.
FIGURE 6.27 MAGNETIC INTENSITY RECORDED AT VARIOUS HEIGHTS ABOVE WELL CASING
< PREVIOUS VIEW
6-87
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
TABLE 6.26 APPLICABILITY OF COMMON BOREHOLE GEOPHYSICAL METHODS
Borehole Geophysical
Method
Dependent Physical
Property
Application (see key below)
1 2 3 4 5 6 7 8 9 10
Acoustic Logs
In-hole acoustic velocity Elastic moduli, density A-2 A-2 X X A-2 X A-2 X X X
Crosshole S-wave velocity Elastic moduli, density A-4 X X X M X X X X M
Downhole S-wave velocity Elastic moduli, density A-4 X X X M X X X X M
Electric and Induction
Spontaneous potential Potential difference A-2 X X X A-2 X X X X X
Single-point resistance Resistance A-2 X X X A-2 X A-2 X X X
Multi-electrode resistivity Resistivity A-2 X A-2 X A-2 X A-2 X X X
Induction Conductivity A-4 X A-4 X M X X X A-4 X
Nuclear
Gamma Gamma radiation A-6 A-6 X X X X X X X X
Gamma-gamma Density A-6 X A-6 X A-6 A-6 X X X X
Neutron Hydrogen content A-6 X A-6 X A-6 X X X X X
Fluid logs
Borehole fuid
characteristics
Resistivity/
conductivity
X A-5 X A-5 A-5 X X X X X
Fluid fow Velocity X A-5 X X X X X X A-1 X
Temperature Temperature X A-5 X A-5 A-5 X X X A-1 X
Miscellaneous
Borehole deviation Inclination X X X X X X X X A-6 X
Video Visual characteristics M X X M A-6 X X A-6 4 A-6
Laser imaging Physical dimensions X X X X X X X X X A-3
Sonic imaging Physical dimensions X X X X X X X X X A-2
Caliper Physical dimensions X X X X X X X A-3 A-3 M
Applications Key
1. Lithology and correlation
2. Hydraulic Conductivity
3. Porosity
4. Fluid properties
5. Depth to groundwater
6. Bulk density
7. Rock structure
8. Borehole parameters
9. Elastic properties of coal refuse, soil and rock
10. Characterization of mine workings or other subsurface voids
Technique Applicability and Required Hole Conditions
A-1 Applicable (cased, fuid-flled hole)
A-2 Applicable (uncased, fuid-flled hole)
A-3 Applicable (uncased, dry hole)
A-4 Applicable (open or fuid-flled hole,
non-conductive casing)
A-5 Applicable (screened or uncased, fuid-flled hole)
A-6 Applicable (any hole condition, but fuid must be clear for video)
M May be applicable, but probably not the best approach
X Not applicable
(ADAPTED FROM USACE, 1995)
< PREVIOUS VIEW
6-88
Chapter 6
MAY 2009
The uphole and downhole techniques are more economical alternatives to the crosshole technique
because only one borehole is required. Downhole measurements are not as accurate as crosshole
measurements, especially if the layers of interest are thin. However, if critical thin layers are not pres-
ent, downhole and uphole measurements may be the preferred means for determining the variation
of the P- and S-wave velocities with depth. Downhole surveys are generally preferred to uphole sur-
veys, because it is usually more practical to induce a strong seismic signal at the surface than it is in
a borehole. Figure 6.29 depicts the deployment for a downhole seismic survey. In terms of analysis,
the downhole or uphole methods difer from the crosshole method in that it is necessary to calculate
incremental velocities on the basis of diferences in travel time between geophones at varying depths
rather than from direct pathways.
Seismic tomography employing surveys from boreholes can be used as a tool for detecting aban-
doned mines. A vertical seismic profle (VSP) can be developed by deploying geophone sensors in a
borehole and a seismic source at multiple surface locations. An emerging technology whereby a drill
bit is used as a downhole source and seismic waves are recorded at the surface is referred to as reverse
vertical seismic profle (RVSP). Although these techniques are generally available and are used in the
oil and gas industry, they are rarely applied to investigations at coal refuse facilities because of their
relatively high cost. Nevertheless, there is some experience in the application of these techniques as
documented by the Colorado School of Mines (2007) and Grito (2003).
FIGURE 6.28 FIELD SETUP FOR CROSSHOLE SEISMIC SURVEY
SOURCE
HOLE
RECEIVER
NO. 1
RECEIVER
NO. 2
LAYER 1
LAYER 2
LAYER 3
PVC CASING
GROUTED INTO
BOREHOLE
DOWNHOLE
SEISMIC
SOURCE
BODY WAVES
GEOPHONE OR
HYDROPHONE
TRIGGER SWITCH
FOR IMPACT TIME
DIGITAL RECORDING
SYSTEM
CLAMPING DEVICE
NOTE: THE SEISMIC WAVES GENERATED MAY BE
P-, SV-, OR SH BODY WAVES DEPENDING
UPON THE TYPE OF SOURCE EMPLOYED. (SIRLES, 2006)
FIGURE 6.28 FIELD SETUP FOR CROSSHOLE SEISMIC SURVEY
FIGURE 6.28 FIELD SETUP FOR CROSSHOLE SEISMIC SURVEY
< PREVIOUS VIEW
6-89
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
A specialized application of a cross borehole technique to identify mine voids is seam wave seis-
mic. Because coal typically has a relatively low seismic velocity compared to the rock formations
that confne coal seams, seismic waves can become trapped in a coal seam and propagate over
long distances with relatively litle atenuation if the coal is continuous. Conversely, obstructions
to a coal seam such as abandoned mine workings will prevent the propagation of a seam wave.
DAppolonia (1982) describes the use of this technique, but since the publication of this report,
seam wave technology has only rarely been used due to the expense and difculty in interpreting
the results. Additional discussion of this technology is provided by Marshall Miller & Associates
(MM&A, 2006) and Pennsylvania State University (2006).
6.4.4.2.2 Video and Laser/Sonar Imaging
Imaging of portions of abandoned mine workings can be critical to an understanding of the orienta-
tion and condition of these mine workings with respect to an existing or planned coal refuse facility.
Video imaging with a borehole camera is a mature technology that also allows for the identifcation
of fractures and collapse zones from a borehole. A disadvantage of using a borehole video camera
in an abandoned mine is that it is difcult to image very far into the mine because of lack of light.
Another disadvantage is that only visual data are obtained, and it is ofen difcult to determine dis-
tances because of a lack of scale.
If a mine is not fooded, an alternative means to image abandoned workings is a laser, range-based
geometric scanner inserted through a dry borehole. Once deployed into a mine void space, a pan and
SEISMIC RECORDER
TRIGGERING
GEOPHONE
PNEUMATIC
PACKER
TRIAXIAL
GEOPHONE
PNEUMATIC PUMP
DYNAMIC
LOAD
STATIC LOAD
FIGURE 6.29 FIELD SETUP FOR DOWNHOLE SEISMIC SURVEY
FIGURE 6.29 FIELD SETUP FOR DOWNHOLE SEISMIC SURVEY
FIGURE 6.29 FIELD SETUP FOR DOWNHOLE SEISMIC SURVEY
< PREVIOUS VIEW
6-90
Chapter 6
MAY 2009
tilt sequence is initialized, producing a scan of the void. The collected data set is then converted in the
feld into a 3D point cloud model of the void. The point cloud model is then converted into a 3D mesh
model of the underground space. These data can subsequently be processed to produce plan views,
sectional views, and volume estimations. Figure 6.30 is a 2D image of a mine entry obtained from a
borehole laser scanner, and Figure 6.31 is a 3D laser image of mine workings.
If a mine is submerged, it is still possible to image mine openings with a submersible, sonar, range-
based scanner inserted into a borehole. Data collected can be oriented using an on-board compass.
Through correlation of several scans at varying elevations, it has proven practical to prepare 3D
models of the fooded space.
6.5 MATERIAL PROPERTY DETERMINATION THROUGH TESTING
Accurate and reliable laboratory soil, rock and materials testing requires selection of the appropriate
tests and care in sample preparation and performing the tests. Laboratory test results must be care-
fully interpreted, based upon the: (1) sampling and testing procedures, (2) types of soil and rock at
the site, (3) geologic history of the site, and (4) types of coal refuse present and their possible use in
refuse disposal facility construction. Suggested references for soil, rock and materials testing proce-
dures include the most current ASTM standards, Mayne et al. (2002), Bardet (1997), and Head (1980,
1982, 1986)
Careful work and atention to detail in laboratory testing are important if accurate and representative
results are to be obtained. This is true for all soils, but it is especially important for soils whose struc-
ture or fabric, and consequently their tested engineering characteristics, can be afected by distur-
bance. When undisturbed sample tests are to be conducted on fne-grained soils, sample disturbance
must be minimized during sampling, transport, storage and testing. Similarly, some rock types (e.g.,
mudstones and claystones) can degrade following stress relief and exposure to the air following drill-
ing. Such materials should be carefully stored and transported so that their in-situ moisture condition
is preserved to the extent possible.
FIGURE 6.30 IMAGE OF MINE ENTRY FROM BOREHOLE LASER SCANNER
LENGTH (FT)
W
I
D
T
H
(
F
T
)
-120 -100 -80 -60 -40 -20 0 20 40
40
20
0
-20
-40
(WORKHORSE TECHNOLOGIES, 2006)
FIGURE 6.30 IMAGE OF MINE ENTRY FROM BOREHOLE LASER SCANNER
FIGURE 6.30 IMAGE OF MINE ENTRY FROM BOREHOLE LASER SCANNER
< PREVIOUS VIEW
6-91
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
Laboratory tests that provide essential data for the analysis and design of an earthen dam or coal
refuse embankment are summarized in Tables 6.27 and 6.28. Table 6.27 lists typical soil laboratory
tests for site characterization, and Table 6.28 summarizes typical laboratory soil tests applicable to
the types of soil, rock and refuse materials used in the construction of embankments and other earth/
refuse structures. In most investigations, all of the classifcation or index property tests listed in the
tables should be performed on representative samples. The need for other tests depends upon the
purpose and subject of the investigation. The use of test data in the analysis and design of earthen
dams and coal refuse impoundments is discussed in Section 6.6.
Standard testing procedures used in soil and rock mechanics are generally applicable to coal refuse,
although modifcations may be appropriate because of the characteristics of coal refuse. Tests for
ash content, pyrite content and leachate water quality, as indicated in Table 6.28, are parameters not
included in a typical embankment testing program. However, these parameters may be important in
any portion of a disposal facility to be constructed from coal refuse. The ash content is an indication of
the amount of coal remaining in the refuse and can be directly correlated with measurements of spe-
cifc gravity and density. In cases where signifcant amounts of coal may remain in the refuse, there is
a possibility of spontaneous combustion. Knowledge of pyrite content and leachate water quality can
facilitate placement procedures that will minimize the potential for environmental impacts.
6.5.1 Selection of Samples for Testing
Samples used for laboratory testing include: (1) bulk and disturbed samples, (2) undisturbed sam-
ples, and (3) reconstituted samples. Reconstituted samples are samples created to have characteristics
similar to in-situ properties and to meet specifed test criteria (e.g., maximum particle dimension
cannot be greater than some proportion of the minimum dimension of the test apparatus). The fol-
lowing text describes the three basic types of samples and their possible use for laboratory testing.
6.5.1.1 Bulk and Disturbed Samples
Representative bulk and disturbed samples of soils and refuse materials (for existing refuse disposal
facilities) are collected from refuse delivered from the preparation plant, test pit excavations and
disturbed sampling (e.g., split-barrel samples) for use in conducting laboratory index (e.g., classif-
cation, moisture content, Aterberg limits) and property characterization (e.g., compaction tests and
FIGURE 6.31 LASER IMAGE OF MINE WORKINGS
(WORKHORSE TECHNOLOGIES, 2006)
FIGURE 6.31 LASER IMAGE OF MINE WORKINGS
FIGURE 6.31 LASER IMAGE OF MINE WORKINGS
< PREVIOUS VIEW
6-92
Chapter 6
MAY 2009
TABLE 6.27 TYPICAL SOIL AND ROCK LABORATORY TESTS FOR COAL
REFUSE DISPOSAL SITE CHARACTERIZATION
Test Category Test Description
ASTM
Designation
Visual
Identifcation
Standard Practice for Description and Identifcation of Soils (Visual-Manual
Procedure)
D 2488
Index
Properties
Standard Test Method for Laboratory Determination of Water (Moisture)
Content of Soil and Rock by Mass
D 2216
Standard Test Methods for Specifc Gravity of Soil Solids by Water
Pycnometer
D 854
Standard Test Method for Particle-Size Analysis of Soils D 422
Standard Practice for Classifcation of Soils for Engineering Purposes
(Unifed Soil Classifcation System)
D 2487
Standard Test Methods for Amount of Material in Soils Finer than the No.
200 (75-m) Sieve
D 1140
Standard Test Methods for Liquid Limit, Plastic Limit, and Plasticity Index
of Soils
D 4318
Corrosivity
Standard Test Method for pH of Soils D 4972
Standard Test Method for Measuring pH of Soil for Use in Corrosion
Testing
G 51
Standard Test Method for Sulfate Ion in Water D 516
Standard Test Method for Field Measurement of Soil Resistivity Using the
Wenner Four-Electrode Method
G 57
Standard Test Methods for Chloride Ion in Water D 512
Organic
Content
Standard Test Methods for Moisture, Ash, and Organic Matter of Peat and
Other Organic Soils
D 2974
Compaction
Test
Standard Test Methods for Laboratory Compaction Characteristics of Soil
Using Standard Effort (12,400 ft-lbf/ft
3
) (600 kN-m/m
3
)
D 698
Standard Test Methods for Laboratory Compaction Characteristics of Soil
Using Modifed Effort (56,000 ft-lbf/ft
3
) (2,700 kN-m/m
3
)
D 1557
Standard Test Methods for Maximum Index Density and Unit Weight of
Soils Using a Vibratory Table
D 4253
Standard Test Methods for Minimum Index Density and Unit Weight of
Soils and Calculation of Relative Density
D 4254
Hydraulic
Conductivity
Standard Test Method for Permeability of Granular Soils (Constant Head) D 2434
Standard Test Methods for Measurement of Hydraulic Conductivity of
Saturated Porous Materials Using a Flexible Wall Permeameter
D 5084
Consolidation
Properties
Standard Test Methods for One-Dimensional Consolidation Properties of
Soils Using Incremental Loading
D 2435
Standard Test Method for One-Dimensional Consolidation Properties of
Saturated Cohesive Soils Using Controlled-Strain Loading
D 4186
< PREVIOUS VIEW
6-93
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
TABLE 6.27 TYPICAL SOIL AND ROCK LABORATORY TESTS FOR COAL
REFUSE DISPOSAL SITE CHARACTERIZATION
(Continued)
Test Category Test Description
ASTM
Designation
Static Strength
Properties
Standard Test Method for Unconfned Compressive Strength of Cohesive
Soil
D 2166
Standard Test Method for Unconsolidated-Undrained Triaxial Compression
Test on Cohesive Soils
D 2850
Standard Test Method for Consolidated-Undrained Triaxial Compression
Test for Cohesive Soils
D 4767
Standard Test Method for Direct Shear Test of Soils under Consolidated-
Drained Conditions
D 3080
Standard Test Method for Laboratory Miniature Vane Shear Test for
Saturated Fine-Grained Clayey Soil
D 4648
Cyclic/Dynamic
Strength Properties
Standard Test Method for the Determination of the Modulus and Damping
Properties of Soils Using the Cyclic Triaxial Apparatus
D 3999
Standard Test Methods for Modulus and Damping of Soils by the
Resonant-Column Method
D 4015
Standard Test Method for Load-Controlled Cyclic Triaxial Strength
of Soil
D 5311
Rock Properties
Standard Test Method for Determination of the Point Load Strength Index
of Rock
D 5731
Standard Test Method for Compressive Strength and Elastic Moduli
of Intact Rock Core Specimens under Varying States of Stress and
Temperatures
D 7012
Standard Test Method for Splitting Tensile Strength of Intact Rock Core
Specimens
D 3967
Standard Test Method for the Slake Durability of Shales and Similar Weak
Rocks
D 4644
(ADAPTED FROM SABATINI ET AL., 2002)
strength and compressibility testing of reconstituted samples) testing. Table 6.18 provides a sum-
mary of common sampling methods for obtaining bulk and disturbed soil samples.
Field personnel directing feld sampling activities need to be aware that the quantity of material
needed depends on the laboratory tests to be performed, the relative amount of coarse (> 3 inches)
particles present, and the size limitations of the test equipment. ASTM D 420, Standard Guide to Site
Characterization for Engineering Design and Construction Purposes, provides general guidelines
for minimum sample weights. These guidelines are presented in Table 6.29. More specifc guidance
on minimum sample weight is provided in the instructions for individual test procedures.
For moisture-sensitive, fne-grained soils, samples should be retained in sealed containers, and bulk
samples should be labeled, indicating information such as test pit number, depth below the ground
surface, and date sampled.
< PREVIOUS VIEW
6-94
Chapter 6
MAY 2009
TABLE 6.28 TYPICAL LABORATORY SOIL TESTS FOR VARIOUS MATERIALS
(1)
Test
ASTM
Test
Method
Type of Material
(2)
Use in Design
F
i
n
e
-
G
r
a
i
n
e
d
S
o
i
l
C
o
a
r
s
e
-
G
r
a
i
n
e
d
S
o
i
l
R
o
c
k
C
o
a
r
s
e
R
e
f
u
s
e
F
i
n
e
R
e
f
u
s
e
C
o
m
b
i
n
e
d
R
e
f
u
s
e
Classifcation or Index
Property Tests
Evaluation of feasible
confguration
Correlation of materials
Selection of samples for
other tests
Selection of borrow
areas
Specifcation of
construction procedures
Determination of
flter and drainage
requirements
Moisture Content
D 2216 a a a a a
Unit Weight
c c c
Specifc Gravity
D 854 b, c b, c b b b b
Atterberg Limits
D 4318 b, c b, c b, c
Particle-Size
Analysis
D 422,
D 2217
b, c b, c b b b, c b
Soil Classifcation D 2487 a a
Compaction Tests
Evaluation of sample
preparation for other
tests
Specifcation of
placement requirements
Standard Proctor
D 698 b b b b
Modifed Proctor
D 1557 b b b b
Relative Density
D 4253,
D 4254
b b b
Hydraulic Conductivity
D 2434,
D 5084
c, d d c, d c, d
Seepage analyses
Determination of pore
pressure for stability
Consolidation
D 2435,
D 4186
c, d c, d Settlement analyses
Shear Strength
Stability analyses
Structure foundation
design
Direct Shear D 3080 c, d d d d c, d c, d
Triaxial compression
D 2850,
D 4767
c, d d d c, d c, d
Unconfned
compression
D 2166 c, d
Vane Shear D 4648 c c
Direct Simple Shear D 6528
< PREVIOUS VIEW
6-95
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
TABLE 6.28 TYPICAL LABORATORY SOIL TESTS FOR VARIOUS MATERIALS
(1)
(Continued)
Test
ASTM
Test
Method
Type of Material
(2)
Use in Design
F
i
n
e
-
G
r
a
i
n
e
d
S
o
i
l
C
o
a
r
s
e
-
G
r
a
i
n
e
d
S
o
i
l
R
o
c
k
C
o
a
r
s
e
R
e
f
u
s
e
F
i
n
e
R
e
f
u
s
e
C
o
m
b
i
n
e
d
R
e
f
u
s
e
Rock Property and
Behavior Tests
Design of rock slopes
Stability of
underground mine
roofs, pillars and
barriers
Evaluation of rock
degradation
Point Load
D 5731 c
Unconfned
Compression
D 7012 c
Slake Durability
D 4644 c
Indirect Tensile
Strength
D 3967 c
Miscellaneous Tests
Evaluation of burning
potential
Corrosion analyses
Ash Content D 2415 b b b
Pyrite Content
D 4239,
D 2492
b b b b
Leachate Water
Quality
D 1068,
D 858,
D 516
b
(3)
b
(3)
b
(3)
Note: 1. The testing program for all signifcant coal refuse embankments should be established by a qualifed,
experienced geotechnical engineer. The types and numbers of tests needed will vary depending of the
purpose of the testing program and the condition being evaluated. Use of these guidelines should not be
substituted for evaluation of specifc site conditions by a qualifed engineer. Additional discussion is pro-
vided in Section 6.5.
2. a tests normally conducted on all samples
b tests normally conducted on representative disturbed samples
c tests normally conducted on representative undisturbed samples
d tests should be conducted on specially-prepared samples to simulate as-constructed behavior
3. Discussion related to conducting water quality tests as part of the geotechnical investigation is provided
in Section 6.4.4.
6.5.1.2 Undisturbed Samples
Undisturbed samples are obtained from cohesive soil strata for laboratory testing to determine
properties such as strength, stratifcation, hydraulic conductivity, density, consolidation, dynamic
behavior, and other engineering characteristics. Specialized procedures are required for obtaining
undisturbed samples of granular soils, thus their application at coal refuse disposal sites is limited
to locations where void ratio, density and strength tests are needed for seismic design. Undisturbed
samples are obtained with specialized equipment designed to minimize the disturbance to the in-
situ structure and moisture content of the soils. Table 6.18 provides a summary of common methods
for obtaining undisturbed soil samples. The importance of sample preservation during undisturbed
sample recovery and transport is described in Section 6.4.3.5.
< PREVIOUS VIEW
6-96
Chapter 6
MAY 2009
6.5.1.3 Reconstituted Samples
Occasionally due to lack of adequate sample volume, difculties encountered in the feld in retrieving
undisturbed samples, or dimensional requirements imposed by specifc test methods, samples must
be created or reconstituted in the laboratory for testing to establish engineering properties needed for
design. The need to use reconstituted samples is more common for granular soils because undisturbed
sampling of sands and gravels is difcult and costly and because the particle sizes in the coarser frac-
tion of a sample may exceed particle-size limits in some tests. For example, the relative density of a
saturated sand or of fne coal refuse can be estimated by in-situ testing, but an acceptably undisturbed
sample for cyclic triaxial testing in the laboratory is difcult and costly to obtain. As a result, samples
may need to be prepared in the laboratory to reasonably recreate the in-situ relative density or void
ratio of the soil or sand-like refuse material. For predominantly coarse-grained soils, samples can be
reconstituted by compaction in a mold. For sand-like refuse material, samples can be reconstituted
by molding moist material or setling from a slurry. For clay-like fne refuse, the depositional history
cannot be readily recreated in the laboratory, so reconstituted samples should not be used.
Reconstituted samples may also be necessary when the coarse particle-size fraction of a sample (usu-
ally a bulk sample) exceeds a dimensional limitation associated with the desired test. For example,
ASTM D 3080, Standard Test Method for Direct Shear Test of Soils under Consolidated Drained
Conditions, prescribes that the maximum particle size not exceed 0.1 times the tested sample diam-
eter (for circular samples) or sample width (for square samples). With this criterion, the maximum
sample particle size cannot exceed 0.2 inches (corresponding approximately to a No. 4 sieve) for a
2-inch-diameter sample and 0.4 inches (corresponding approximately to a -inch sieve) for a 4-inch-
square sample. For this example, if the maximum particle size exceeded 0.4 inches, the particle-size
distribution for the sample used for direct shear testing would need to be adjusted to accommodate
the maximum-particle-size criterion. ASTM test methods identify such gradational limitations and
describe sample preparation techniques and test result evaluation methods to account for the removal
of over-size particle fractions.
6.5.2 Classifcation and Index Property Tests
To catalog soils and coal refuse materials that will form the foundation, embankment cross sec-
tion, and impoundment of a coal refuse disposal facility, samples from the feld testing program
should be examined and accurately classifed. The system of classifcation used by most geotechni-
cal engineers and government agencies is the Unifed Soil Classifcation System (USCS) as described
in ASTM D 2487, Standard Practice for Classifcation of Soils for Engineering Purposes (Unifed
Classifcation System). This classifcation system for engineering purposes is based on laboratory
determination of particle-size distribution and Aterberg limits. ASTM D 2488, Standard Practice
for Description and Identifcation of Soils (Visual-Manual Procedure), provides a companion pro-
cedure for preliminary classifcation of soils based on visual and manual techniques available to
feld and laboratory personnel.
TABLE 6.29 GUIDELINES FOR MINIMUM SAMPLE WEIGHTS
Test and Soil Characteristics Minimum Sample Weight
Visual classifcation
2 ounces to 1 pound
Soil constants and particle-size analysis of non-gravelly soil
1 to 5 pounds
Soil compaction tests and sieve analysis of gravelly soils
40 to 80 pounds
Aggregate properties
100 to 400 pounds
(ADAPTED FROM ASTM, 2008b,c)
< PREVIOUS VIEW
6-97
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
As shown in Table 6.30, the USCS divides soils into two main classes: coarse-grained and fne-grained
soils. Highly organic soils form an additional division. Coarse-grained soils are soils composed of
predominantly gravel- and/or sand-sized particles (greater than 50 percent retained on a No. 200
sieve) that can be separated into eight groups based primarily on the coarseness, gradation, and per-
centage of fnes and secondly on the plasticity of fnes. Fine-grained soils are soils composed of pre-
dominantly silt- and/or clay-sized particles (greater than 50 percent passing a No. 200 sieve) that can
be separated into six groups, based primarily on plasticity and secondly on coarseness, gradation,
and percentage of coarse fractions, if present. Generally, the system is arranged so that any sample
can be classifed by visual observation and simple tests that ofen can be conducted in the feld by the
engineer or geologist supervising an exploration program. However, laboratory tests on representa-
tive samples are needed to confrm the visual classifcation of soil properties or to classify borderline
cases. Table 6.30 provides numerical criteria used for classifcation based upon laboratory test results.
Figure 6.32 presents the plasticity chart used to classify fne-grained soils based on the liquid limit
(LL) and plasticity index (PI) of tested samples.
Table 6.8, as adapted from Sherard et al. (1963), presents an approximate correlation between the
USCS classifcation and the engineering and design properties of soils. Although the table is not a
substitute for detailed laboratory tests, it can be used to help determine which tests should be con-
ducted and to preliminarily evaluate available embankment materials.
There are no categories in the USCS for coal refuse materials, and current practice is to classify and
test them in the same manner as other soil materials. Each of the following discussions of classifca-
tion tests concludes with information on coal refuse properties as compared to properties of other
soils. The basis for the discussion includes published data, as cited, and project experience, although
it should be recognized that substantial variation can occur due to site-specifc conditions and mining
and coal preparation practices.
6.5.2.1 Moisture Content
Moisture content tests are typically conducted on disturbed and undisturbed samples obtained
at a site to: (1) beter characterize in-situ conditions and evaluate other tests, (2) evaluate borrow
material suitability through comparison of natural moisture content and the moisture content
required for proper compaction, and (3) provide information for calculating the void ratio of
saturated samples.
Void ratio is defned as the ratio of void space to the volume of the solid particles:
e = V
v
/ V
s
(6-9)
where:
V
v
= volume of voids (length
3
)
V
s
= volume of solids (length
3
)
Properly obtained samples of fne-grained soils sealed in plastic, wax or airtight jars at the time of
sampling can be accurately tested for moisture content at a later time in the laboratory. Testing of
coarse-grained soils may not be accurate if water is lost by drainage during sampling. The feld engi-
neer should note whether moisture content measurements for coarse-grained soil samples may have
been afected by the sampling.
As described in ASTM D 2216, Standard Test Method for Laboratory Determination of Water (Mois-
ture) Content of Soil and Rock by Mass, the procedure for measuring moisture content is to weigh
< PREVIOUS VIEW
6-98
Chapter 6
MAY 2009
TABLE 6.30 SOIL CLASSIFICATION CHART (LABORATORY METHOD)
Criteria for Assigning Group Symbols and Group Names
Using Laboratory Tests
(1)
Soil Classifcation
Group
Symbol
Group Name
(2)
GRAVELS
50% of coarse
fraction retained
on No. 4 Sieve
CLEAN GRAVELS C
u
4 and 1 C
c
3
(5)
GW Well-graded Gravel
< 5% fnes C
u
< 4 and/or 1 > C
c
> 3
(5)
GP Poorly-graded Gravel
(6)
GRAVELS WITH
FINES
Fines classify as ML or MH GM Silty Gravel
(6,7,8)
> 12% fnes
(3)
Fines classify as CL or CH GC Clayey Gravel
(6,7,8)
SANDS
50% of coarse
fraction retained
on No. 4 Sieve
CLEAN SANDS Cu 6 and 1 Cc 3
(5)
SW Well-graded Sand
(9)
< 5% fnes
(4)
Cu < 6 and 1 > Cc > 3
(5)
SP Poorly-graded Sand
i
SANDS WITH
FINES
Fines classify as ML or MH SM Silty Sand
(7,8,9)
> 12% fnes
(4)
Fines classify as CL or CH SC Clayey Sand
(7,8,9)
SILTS AND
CLAYS
LL < 50
Inorganic
PI > 7 and plots on or above A
line
(10)
CL Lean Clay
(11,12, 13)
PI < 4 or plots below A line
(10)
ML Silt
(11,12,13)
Organic
LL after oven drying < 0.75 LL
before oven drying
OL Organic Clay
(11,12,13,14)
OL Organic Silt
(11,12,13,15)
SILTS AND
CLAYS
LL 50
Inorganic
PI plots on or above A line CH Fat Clay
(11,12,13)
PI plots below A line MH Elastic Silt
(11,12,13)
Organic
LL after oven drying < 0.75 LL
before oven drying
OH Organic Clay
(11,12,13,16)
OH Organic Silt
(11,12,13,17)
Highly fbrous
organic soils
Primarily organic matter, dark in color, with
organic odor
PT
Peat and
Muskeg
Note: 1. Based on the material passing the 3-in (75-mm) sieve.
2. If feld sample contained cobbles or boulders, or both, add with cobbles or with boulders to group name.
3. C
u
= D
60
/D
10
= uniformity coeffcient (UC);
C
c
= (D
30
)
2
/ (D
60
x D
10
) = coeffcient of curvature
4. If soil contains 15% sand, add with sand to group name.
5. Gravels with 5 to 12% fnes require dual symbols:
GW-GM well-graded gravel with silt
GW-GC well-graded gravel with clay
GP-GM poorly-graded gravel with silt
GP-GC poorly-graded gravel with clay
< PREVIOUS VIEW
6-99
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
a wet sample, dry it in a constant-temperature oven at 105 C until the weight is constant (approxi-
mately 24 hours for small samples of fne-grained soils), and then weigh the dry sample. In soil
mechanics practice, the moisture content is defned as the ratio of the weight of water (wet weight
minus dry weight) to the dry weight. This is sometimes referred to as the dry-weight-basis moisture
content. In other disciplines, moisture content may be defned on a wet-weight basis, i.e., moisture
content is defned as the ratio of the weight of water to the wet weight of soil.
FIGURE 6.32 PLASTICITY CHART
LIQUID LIMIT, LL
P
L
A
S
T
I
C
I
T
Y
I
N
D
E
X
,
P
I
(ASTM, 2008)
10
20
0
4
7
MH or OH
"
A
"
L
I
N
E
"
U
"
L
I
N
E
ML or OL CL-ML
0 10 16 20 30 40 50 60 70 80 90 100 110
30
40
50
60
FOR CLASSIFICATION OF FINE-GRAINED SOILS AND THE
FINE-GRAINED FRACTION OF COARSE-GRAINED SOILS,
THE FOLLOWING RELATIONSHIPS APPLY:
"A" LINE - HORIZONTAL AT PI=4 TO LL-25.5,
THEN PI=0.73 (LL-20)
"U" LINE - VERTICAL AT LL = 16 to PI=7,
THEN PI=0.9 (LL-8)
C
H
o
r
M
H
C
L
o
r
O
L
FIGURE 6.32 PLASTICITY CHART
TABLE 6.30 SOIL CLASSIFICATION CHART (LABORATORY METHOD)
(Continued)
Note 6. If fnes classify as CL-ML, use dual symbol GC-GM or SC-SM.
7. If fnes are organic, add with organic fnes to group name.
8. If soil contains 15% gravel, add with gravel to group name.
9. Sands with 5 to 12% fnes require dual symbols:
SW-SM well-graded sand with silt
SW-SC well-graded sand with clay
SP-SM poorly-graded sand with silt
SP-SC poorly-graded sand with clay
10. If Atterberg limits plot in the orange area in Figure 6.32, soil is a CL-ML (silty clay).
11. If soil contains 15 to 29% plus No. 200 sieve, add with sand or with gravel, whichever is predominant.
12. If soil contains 30% plus No. 200 sieve, predominantly sand, add sand to group name.
13. If soil contains 30% plus No. 200 sieve, predominantly gravel, add gravelly to group name.
14. PI 4 and plots on or above A line.
15. PI < 4 or plots below A line.
16. PI plots on or above A line.
17. PI plots below A line.
(ADAPTED FROM ASTM, 2008b,c)
FIGURE 6.32 PLASTICITY CHART
< PREVIOUS VIEW
6-100
Chapter 6
MAY 2009
If a soil sample contains a signifcant amount of organic material, this method of measuring moisture
content is not always satisfactory, because heating the sample to 105 C may drive of some of the
organic material in addition to the water. Alternatively, ASTM D 2216 permits oven drying at 60 C
for the moisture content of organic soils and organic materials. Most coal refuse is not signifcantly
afected by oven drying at 105 C, although this may need to be considered when working with lower
grade coals such as lignite.
6.5.2.2 Specifc Gravity
The specifc gravity of a soil is the ratio of the weight of a given volume of soil solid particles to the
weight of an equal volume of distilled water at 4 C. As used in geotechnical engineering, the term
specifc gravity refers to the average specifc gravity of the individual soil particles in a sample rather
than bulk specifc gravity. Specifc gravity is determined in the laboratory in accordance with ASTM
D 854, Standard Test Methods for Specifc Gravity of Soil Solids by Water Pycnometer. The test is
performed by weighing a calibrated botle containing soil particles suspended in distilled water and
comparing this to the weight of the same botle containing an equal volume of distilled water only.
Specifc gravity is used to determine relationships between soil weight and soil volume. Specifc
gravity is used for: (1) computing the void ratio of a soil, (2) hydrometer analyses, and (3) predicting
the unit weight of a soil. Occasionally, the specifc gravity may be useful in soil mineral classifcations
(e.g., iron minerals have a higher specifc gravity than silica).
Soils typically have a specifc gravity ranging from 2.4 to 2.8. For many design purposes specifc
gravity can be estimated without testing. For coal refuse facilities, specifc gravity is an important
design parameter because coal refuse ofen contains a signifcant amount of materials with specifc
gravity in the range of 1.3 to 1.6. As a result, coarse coal refuse can have a specifc gravity ranging
from as low as 1.5 to as high as 2.8. The most common range is between 1.9 and 2.4. Similarly, spe-
cifc gravity measured for fne coal refuse typically ranges from 1.4 to 2.3. Published data on spe-
cifc gravity and unit weight of coarse and fne coal refuse from sites in the northern Appalachian
region illustrating some of the variability in these parameters is presented in Tables 6.31 and 6.32,
respectively, as compiled by Hegazy et al. (2004). The database for these summaries from Hegazy
et al. (2004) was developed from geotechnical investigations of existing coal refuse disposal sites in
western Pennsylvania and England. In-situ samples were collected from the sites using both dis-
turbed methods (bucket samples from test pits and fne coal refuse deltas and split-barrel samples
from boreholes) and undisturbed sampling methods (Shelby- and Dennison-tube samples).
The coefcient of variability (COV) is the ratio of the standard deviation of a set of data divided by
the mean. Typically, values of COV below 10 percent are thought to be low, between 10 and 30 per-
cent moderate, and above 30 percent high. Values of total unit weight
T
, dry unit weight
D
, and
specifc gravity G
s
for coarse and fne coal refuse are provided in Tables 6.31 and 6.32. For coarse coal
refuse, Table 6.31 indicates low variability for
T
and
D
and moderate variability for G
s
. For fne coal
TABLE 6.31 IN-PLACE UNIT WEIGHT AND SPECIFIC GRAVITY OF COARSE COAL REFUSE
Property Dimension Average
Standard
Deviation
Coeffcient of
Variation
T
lb/ft
3
124 5.8 0.048
D
lb/ft
3
115 5.5 0.047
G
s
Dimensionless 2.02 0.31 0.154
(ADAPTED FROM HEGAZY ET AL., 2004)
< PREVIOUS VIEW
6-101
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
refuse, Table 6.32 indicates low variability for
T
and moderate variability for
D
and G
s
. Figure 6.33
shows the efect of carbon content on the specifc gravity of coal refuse materials.
Designers should recognize that values of specifc gravity and unit weight for coal refuse are lower
than for commonly encountered soils. The lower specifc gravity of coal refuse results in lower den-
sities, higher moisture contents at a given void ratio, and the potential for reduced stability with
respect to seepage forces. These characteristics are discussed further in Section 6.6.4.
6.5.2.3 Atterberg Limits
The Aterberg Limits defne the boundaries between four states of consistency (hardness or sofness)
of fne-grained soils. In order of decreasing moisture content, these states are: liquid, plastic, semi-
solid and solid. The boundaries or limits between these states are:
Liquid limit (LL) boundary between the liquid and plastic states
Plastic limit (PL) boundary between the plastic and semi-solid states
Shrinkage limit (SL) boundary between the semi-solid and solid states
The plasticity index (PI) is defned as LL minus PL. The liquid and plastic limits and plasticity index
are determined in accordance with ASTM D 4318, Standard Test Methods for Liquid Limit, Plastic
Limit, and Plasticity Index of Soils. The shrinkage limit is determined in accordance with D 4943,
Standard Test Method for Shrinkage Factors of Soils by the Wax Method.
TABLE 6.32 UNIT WEIGHT AND SPECIFIC GRAVITY FOR FINE COAL REFUSE
Property Dimension Average
Standard
Deviation
Coeffcient of Variation
T
lb/ft
3 86 7.7 0.096
D
lb/ft
3 62 9.1 0.162
G
s
Dimensionless
1.52 0.25 0.165
(ADAPTED FROM HEGAZY ET AL., 2004)
The liquid limit is the moisture content at which the soil sample fows and closes a standard width
groove when the sample is jarred in a specifed way. For practical purposes, it is the moisture con-
tent at which the soil has essentially no shear strength (the soil becomes liquid). The test for the LL is
conducted using a standard liquid-limit device and grooving tool. The plastic limit is the moisture
content at which the soil begins to crumble when rolled by hand into -inch-diameter threads. The
shrinkage limit (SL) is the moisture content sufcient to fll the pores of the soil at the minimum
volume it will atain by drying. Other useful parameters from these tests are the plasticity index (PI)
and the liquidity index (LI). The plasticity index is an indicator of soil plasticity, and the LI, which
equals (w PL)/PI, is an indicator of stress history and soil sensitivity. The liquidity index is approxi-
mately 1 for normally-consolidated soils and zero for over-consolidated soils. An LI greater than 1
indicates high sensitivity.
Many properties of fne-grained clays and silts can be correlated to the Aterberg limits, and the plas-
ticity chart shown in Figure 6.32 can be useful in interpreting the correlation. Procedures for using the
plasticity chart for various soil types are discussed by Terzaghi et al. (1996).
Caution and considerable judgment should be used when applying the plasticity chart to coal refuse,
because the chart is based on the behavior of natural, fne-grained soils. Tests conducted on only the
< PREVIOUS VIEW
6-102
Chapter 6
MAY 2009
fnes portion of coarse refuse have obtained LLs in the range of 25 to 35 percent and PIs typically
below 12 percent. Table 6.4 cites published data characterizing the properties of fne coal refuse. Tests
conducted on fne refuse samples from impoundments show LLs in the range of 20 to 40 percent and
PIs generally below 15 percent, although higher PIs have been reported. Factors afecting plastic-
ity are discussed in Section 6.2.3.2. A summary of statistical properties related to Aterberg limits for
fne coal refuse from northern Appalachian sites is provided in Table 6.33. Other data suggest that
the plasticity of fne refuse from an impoundment increases with distance from the slurry discharge
point, as the content of clay-size refuse particles increases.
6.5.2.4 Particle-size Distribution
The particle-size distribution of a coarse-grained soil is an important factor in its engineering behav-
ior. The particle-size distribution of a coarse-grained soil can be used to classify the soil, to estimate
hydraulic conductivity and to provide a qualitative indication of soil strength and deformation char-
acteristics. For fne-grained soils, plasticity and moisture content are beter indices of soil behav-
ior than particle-size distribution. For both fne and coarse material, the particle-size distribution is
needed for checking flter criteria requirements as part of prevention of piping of fnes into coarser
flter and drainage zones. Figure 6.34 shows typical gradation curves for several types of materials
including coarse and fne coal refuse. Of particular note is the correlation between types of soil (clay,
silt, sand and gravel) and sieve sizes and particle diameters.
For comparative purposes between soil types and for certain design applications, the uniformity of
a soil can be expressed by the uniformity coefcient, which is the ratio of D
60
to D
10
, where D
60
is the
particle diameter at which 60 percent of the soil weight is fner and D
10
is the particle diameter at
which 10 percent of the soil weight is fner. Sand and gravel soils having uniformity coefcients less
than 6 and 4, respectively, are considered to be uniform. For example, the uniformity coefcients of
1.0 1.4 1.8 2.2 2.6 3.0
0
20
40
60
80
100
SPECIFIC GRAVITY, G
S
F
I
X
E
D
C
A
R
B
O
N
P
E
R
C
E
N
T
(
D
R
Y
B
A
S
I
S
)
FIGURE 6.33 EFFECT OF CARBON CONTENT ON SPECIFIC
GRAVITY OF COAL REFUSE MATERIAL
(HEGAZY ET AL., 2004)
FIGURE 6.33 EFFECT OF CARBON CONTENT ON SPECIFIC GRAVITY
OF COAL REFUSE MATERIAL
FIGURE 6.33 EFFECT OF CARBON CONTENT ON SPECIFIC GRAVITY
OF COAL REFUSE MATERIAL
< PREVIOUS VIEW
6-103
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
TABLE 6.33 STATISTICAL PROPERTIES FOR ATTERBERG LIMITS AND MOISTURE
CONTENTS OF FINE COAL REFUSE AT NORTHERN APPALACHIAN SITES
Property
Average
(%)
Standard
Deviation
Coeffcient of
Variation
LL 31.2 5.2 0.17
PL 20.1 3.4 0.17
PI 11.2 3.1 0.28
w 33.0 11.5 0.35
(HEGAZY ET AL., 2004)
the two sandy soils shown in Figure 6.34 are about 6.1 and 1.4. These soils are termed well-graded
sand and uniform sand, respectively.
As described in ASTM D 422, Standard Test Method for Particle-Size Analysis of Soils, the particle-
size distribution of the portion of a soil sample coarser than a No. 200 sieve (0.074-mm-square open-
ings) is generally determined by sieve analysis. This procedure consists of shaking a dry soil sample
or washing a wet soil sample through a stack of wire screens of decreasing opening size. The diam-
eter of an individual soil particle is defned as the minimum side dimension of a square hole through
which the soil particle will pass.
The particle-size distribution of the portion of a sample fner than a No. 200 sieve can be measured
with a hydrometer test as described in ASTM D 422. In this test, a sample of soil is vigorously mixed
with water and a defocculating agent to form a suspension. The suspension is then allowed to sit
FIGURE 6.36 PARTICLE-SIZE DISTRIBUTION
100 8 6 4 3 2 10 8 6 4 2 1 3
10
20
0
30
40
50
60
70
80
90
100
8 6 4 3 2 0.1 8 6 4 3 2 0.01 8 6 4 3 2 0.001
P
E
R
C
E
N
T
F
I
N
E
R
B
Y
W
E
I
G
H
T
4
"
3
/
8
"
1
/
2
"
3
/
4
"
1
"
3
"
2
1
/
2
"
2
"
1
1
/
2
"
2
0
0
1
0
0
5
0
7
0
4
0
3
0
1
6
81
0
4
CLEAR SQUARE OPENINGS U.S. STANDARD SIEVE NUMBERS HYDROMETER ANALYSIS
PARTICLE DIAMETER (mm)
COBBLES
GRAVEL SAND
SILT OR CLAY (PLASTIC OR NON-PLASTIC)
FINE MEDIUM COARSE FINE COARSE
FINE COAL REFUSE
COARSE COAL REFUSE
WELL-GRADED
GRAVEL
POORLY-GRADED
GRAVEL
WELL-GRADED
SAND
UNIFORM SAND
FIGURE 6.34 PARTICLE-SIZE DISTRIBUTION
FIGURE 6.34 PARTICLE SIZE DISTRIBUTION
FIGURE 6.34 PARTICLE SIZE DISTRIBUTION
< PREVIOUS VIEW
6-104
Chapter 6
MAY 2009
undisturbed in a 1000-ml glass cylinder to permit setlement of the suspended soil particles. Periodi-
cally, the change in the specifc gravity of the suspension is measured with a calibrated hydrometer,
allowing the approximate distribution of particle sizes to be calculated. The specifc gravity of the
suspension changes with time because the larger particles setle faster than smaller particles of the
same specifc gravity. The test result is only approximate because the calculation is based upon the
assumption that the particles are spherical, of equal specifc gravity, and do not interfere with each
other during setlement. In actual tests, the particles are not spherical, there are variations in specifc
gravity, and considerable interference between particles likely occurs during setlement. Because the
distributions determined by the hydrometer test are primarily used for comparative purposes, the
accuracy of the test is generally not of major concern.
The efect of diferences in specifc gravity of various particles in the hydrometer test is greater for
coal refuse than for other soils, because coal refuse consists of coal with specifc gravity as low as 1.3
and rock fragments with specifc gravity as high as 2.8. For most analyses of coal refuse, it is appro-
priate to consider the average specifc gravity of the entire sample. However, it should be understood
that the variation in specifc gravity of coal refuse particles adds greater than normal inaccuracy to
the portion of the gradation curve determined by the hydrometer method.
Coarse coal refuse delivered to disposal facilities typically has the grain-size distribution of a well-
graded silty sand and gravel. Coal preparation usually limits the upper size to about 5 inches,
although this size has more characteristically been less than 3 inches. Uniformity coefcients for
coarse refuse range from about 20 to several hundred. Because many coal refuse particles are
extremely friable and highly susceptible to both chemical and mechanical deterioration, the par-
ticle-size distribution changes during preparation, transportation and placement at the disposal
facility. As delivered to the point of disposal, coarse refuse typically contains between 5 and 30
percent of particles fner than the No. 200 sieve (0.075 mm). The clay-size fraction (fner than 0.002
mm) is usually very small and ofen less than 2 percent. Sampling programs should be designed
to evaluate the potential for particle degradation through collection of both as-delivered samples
and afer-placement samples.
Particle-size degradation occurs at the surface of a coarse refuse embankment or disposal facility, due
to both chemical and mechanical deterioration. This behavior is described in Andrews et al. (1980)
where the environmental efects of slaking of surface mine spoils in the eastern and central U.S. were
evaluated. The study showed that degradation of surface mine spoil occurred over periods of years
depending on the rock type(s) found in the spoil and their depth of burial. Field examination showed
that degradation was more predominant with fner-grained, rock-type spoils (e.g., shales, mudstones
and claystones) and was most prevalent in the upper 5 to 10 feet depending on the amount of com-
paction (if any) during spoil placement. For embankments constructed using coarse refuse, particle-
size degradation is also afected by hauling equipment and mechanical compaction that occurs as the
material is placed.
Fine coal refuse delivered to disposal facilities from preparation plants is typically a slightly clayey,
sandy silt. Generally, 50 to 80 percent of the material will pass the No. 200 sieve, most of which is
silt-size. Fine refuse segregates afer being deposited, with the larger and heavier particles setling
out of suspension near the discharge point. Farther from the discharge point, the setled refuse is
predominantly fner-grained, and samples containing nearly 50 percent clay-size particles have been
reported. A summary of statistical properties related to particle-size distribution for fne coal refuse
samples from sites in northern Appalachia is provided in Table 6.34.
6.5.2.5 Chemical Characterization
Some soil and rock materials encountered at refuse disposal sites and the refuse being placed can have
adverse chemical characteristics that may lead to undesirable environmental conditions or deteriora-
< PREVIOUS VIEW
6-105
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
tion of construction materials. Therefore, the chemical characteristics of these materials (e.g., corro-
sivity, organic content, ash content, pyrite content, and leachate water quality) should be determined.
When deleterious conditions are encountered, appropriate amendment or containment/protection
requirements should be implemented. Table 6.35 lists laboratory test methods that can be used for
chemical characterization of soil, rock and refuse materials.
6.5.3 Compaction and Density Tests
Any soil placed as part of a structural fll, including coal refuse in embankments, is normally com-
pacted to increase density and shear strength and to decrease compressibility and hydraulic conduc-
tivity. This makes relatively steep slopes possible, reduces seepage from impoundments, and reduces
the potential for spontaneous combustion by reducing the fow of air and water into the embank-
ment. In the feld, compaction is accomplished with various types of rollers, including sheepsfoot,
static and vibrating steel drum, and rubber-tired. The type of roller that is most appropriate depends
upon the material being compacted, as discussed in Section 11.4.3.
6.5.3.1 Fine-grained Soils and Coal Refuse
Standard laboratory test procedures to control and evaluate the degree of compaction that can be
achieved in the feld during placement have been established. The test most commonly used for
fne-grained soils and coal refuse utilizes impact compaction. The test procedure entails placing soil
or refuse in several layers in a standard mold and compacting each layer by dropping a hammer of
specifed weight a specifed distance for a specifed number of times per layer.
The two standardized impact compaction tests are the standard Proctor and the modifed Proc-
tor tests. The standard Proctor compaction test was developed in the 1930s and was designed to
approximate the compactive energy applied by feld compaction equipment then available. As feld
equipment became larger and more efcient, the modifed proctor compaction test was developed
to approximate the greater compactive energy of the newer equipment. The test procedures for the
standard and modifed Proctor compaction tests are described in ASTM D 698, Standard Test Meth-
ods for Laboratory Compaction Characteristics of Soil Using Standard Efort (12,400 f-lbf/f
3
(600
kN-mm
3
)) and ASTM D 1557, Standard Test Methods for Laboratory Compaction Characteristics
of Soil Using Modifed Efort (56,000 f-lbf/f
3
(2,700 kN-mm
3
)), respectively.
These test methods are suitable for soils and coal refuse that have 30 percent or less retained on the
-inch sieve and that have more than 15 percent by dry weight passing a No. 200 sieve. If more than 30
percent is retained on a -inch sieve for either test, the unit weight and moisture content should be cor-
rected in accordance with ASTM D 4718, Standard Practice for Correction of Unit Weight and Water
TABLE 6.34 PARTICLE-SIZE DISTRIBUTION FOR FINE COAL REFUSE
Particle Size or
Percent Passing
Dimension Average
Standard
Deviation
Coeffcient of
Variation
D
10
mm
0.010 0.015 1.50
D
30
mm
0.037 0.055 1.49
D
50
mm
0.127 0.128 1.01
D
60
mm
0.196 0.209 1.07
Passing No. 200
(0.075-mm) sieve
% 57.7 25.0 0.43
(HEGAZY ET AL., 2004)
< PREVIOUS VIEW
6-106
Chapter 6
MAY 2009
Content for Soils Containing Oversize Particles. Alternatively, a 12-inch-diameter compaction mold
using standard Proctor compactive efort can be employed using the U.S. Army Corps of Engineers
procedure, Compaction Test for Earth-Rock Mixtures, described in Section 11.5.1 (USACE, 1986). If
less than 15 percent by dry weight passes the No. 200 sieve, the density of the soil may not be afected
by changes in moisture, and the guidelines presented in Section 6.5.3.2 for coarse-grained soils may be
applicable. Coarse coal refuse may contain less than 15 percent by dry weight passing a No. 200 sieve,
but it generally does respond to changes in moisture content when compacted. Accordingly, in practice,
the standard Proctor test is typically used to evaluate the compaction and density of coarse refuse.
Normally, a series of compaction tests is performed on several soil samples prepared at varying mois-
ture contents. From the test results ploted as shown in Figure 6.35, the maximum density atainable
and the moisture content at which the maximum density is atained (the optimum moisture content),
can be determined. The greater compactive energy of the modifed Proctor compaction test produces
higher maximum densities at lower optimum moisture contents than the standard Proctor compac-
tion test. The 100-percent-saturation (or zero-air-voids) curve to the right of the compaction curves in
Figure 6.35 can be determined using the relationship:
d
=
w
G
s
/ (1 + w G
s
) (6-10)
where:
d
= dry density of soil (force/length
3
)
w
= unit weight of water (force/length
3
)
G
s
= specifc gravity of solids (dimensionless)
w = moisture content expressed as a decimal value (dimensionless)
TABLE 6.35 TEST METHODS FOR CHEMICAL CHARACTERIZATION
OF SOIL, ROCK AND REFUSE MATERIALS
Item Description ASTM Test Method
Corrosivity
Standard Test Method for pH of Soils
Standard Test Method for pH of Soil for Use in Corrosion Testing
Standard Test Method for Sulfate Ion in Brackish Water
Standard Test Method for Field Measurement of Soil Resistivity Using
the Wenner Four-Electrode Method
Standard Test Methods for Chloride Ion in Water
D 4972
G 51
D 4130
G 57
D 512
Organic Content
Standard Test Methods for Moisture, Ash, and Organic Matter of Peat
and Other Organic Soils
D 2974
Ash Content Standard Test Method for Ash in Coal Tar and Pitch D 2415
Pyrite Content Standard Test Method for Forms of Sulfur in Coal D 2492
Leachate Water
Quality
(1)
Standard Test Method for Leaching Solid Material in a Column
Apparatus
Standard Test Method for Shake Extraction of Mining Waste by the
Synthetic Precipitation Leaching Procedure
D 4874
D 6234
Note: 1. State regulatory agencies may require specifc test procedures. Other standard test method references
include EPA Method 1312, Synthetic Precipitation Leaching Procedure (SPLP) and EPA Method 1320,
Multiple Extraction Procedure (MEP).
< PREVIOUS VIEW
6-107
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
Equation 6-10 provides a check on the compaction test results to confrm that no compaction test result
plots to the right of the 100-percent-saturation curve and that the shape of the dry density-moisture
relationship wet of optimum moisture content generally parallels the 100-percent-saturation curve.
Additionally, Equation 6-10 demonstrates the need to know the specifc gravity of a material when
determining compaction and density.
Specifcations for construction of compacted flls usually require that the density atained in the feld
be equal or greater than a certain percentage of the maximum density atained in the laboratory
compaction tests (for structural embankment zones, normally 95 percent of the maximum density at
the optimum moisture content from the standard Proctor test). To help achieve good density control,
specifcations also usually require the fll to be placed at a moisture content near the optimum mois-
ture content (normally in the range from 2 percent below optimum to 3 percent above optimum).
DAppolonia (1973) reported a measured in-place dry unit weight for uncompacted coarse refuse of
80 to 110 pounds per cubic foot (pcf), with a median value of 94 pcf based upon data from 200 tests.
For compaction testing performed at mine sites in northern Appalachia, Hegazy et al. (2004) reported
a median dry density for compacted coarse refuse of 115 pcf. While these results demonstrate dry
densities encountered in some specifc situations, there can be considerable variation depending
upon geographic location and mining and coal processing activities.
Because of the degradation of coarse coal refuse due to equipment trafc during placement and weath-
ering afer placement, the percentage of fnes in aged coarse coal refuse will likely be greater than in
fresh or recently placed coarse coal refuse. Saxena et al. (1984) report that particle breakdown due to a
combination of weathering and compaction produces beter graded materials with higher density and
strength and lower compressibility and hydraulic conductivity. Hegazy et al. (2004) presented the results
of sieve analyses performed for fresh coarse coal refuse samples and repeated afer compaction to deter-
mine the efect of compaction on the fnes content (i.e., the percent passing the No. 200 sieve). For the
tests ploted in Figure 6.36, the average increase of fnes due to compaction was approximately 4 percent.
WATER CONTENT (%)
D
R
Y
D
E
N
S
I
T
Y
(
P
C
F
)
FIGURE 6.35 TYPICAL MOISTURE-DENSITY RELATIONSHIPS
FOR DYNAMIC COMPACTION
1
0
0
P
E
R
C
E
N
T
S
A
T
U
R
A
T
I
O
N
MAXIMUM DRY DENSITY
(MODIFIED PROCTOR)
MAXIMUM DRY DENSITY
(STANDARD PROCTOR)
OPTIMUM WATER CONTENT
(MODIFIED PROCTOR)
OPTIMUM WATER CONTENT
(STANDARD PROCTOR)
FIGURE 6.35 TYPICAL MOISTURE-DENSITY RELATIONSHIPS FOR DYNAMIC COMPACTION
FIGURE 6.35 TYPICAL MOISTURE-DENSITY RELATIONSHIPS FOR DYNAMIC COMPACTION
< PREVIOUS VIEW
6-108
Chapter 6
MAY 2009
In applications sensitive to fnes content, breakdown due to weathering and compaction can be evalu-
ated through slake durability testing of fresh and weathered compacted samples (Section 6.5.9.4).
Specifcations for clay core materials designed to restrict seepage through an embankment are
normally based upon Proctor test results but these specifcations ofen require that the material to
be placed slightly wet of optimum. This results in lower strengths, but eliminates the potential for
britle soil behavior. The resulting core should be sufciently fexible to allow for small amounts
of diferential movement without the development of cracks that would allow the passage of large
volumes of water and cause piping. Sherard et al. (1963), Sherard (1973), and Lo (1990) report a
number of dam failures caused by cracking, and they discuss the related importance of compaction
specifcations and control.
Slurry-placed fne coal refuse typically has a very low in-situ density because of low specifc gravity
and moisture contents near the liquid limit. Dry densities near 50 pcf have been reported. Coarser or
dryer portions of the fne refuse may have dry densities of 70 pcf or higher.
6.5.3.2 Coarse-Grained Soils and Coal Refuse
Relative density is the dry density of a soil in relation to the minimum and maximum dry densities
that can be achieved by specifc laboratory procedures. The relative density test is applicable to free-
draining, cohesionless soils with low fnes content (i.e., less than 15 percent non-plastic fnes passing
the No. 200 sieve) that do not have a well-defned moisture-density relationship. The maximum dry
density is determined in accordance with ASTM 4253, Standard Test Methods for Maximum Index
Density and Unit Weight of Soils Using a Vibratory Table, and the minimum dry density and rela-
FIGURE 6.36 PARTICLE-SIZE DISTRIBUTION OF COARSE
COAL REFUSE AFTER COMPACTION
0 20 30 40 50
0
10
20
30
40
50
PERCENT PASSING NO. 200 SIEVE BEFORE COMPACTION
P
E
R
C
E
N
T
P
A
S
S
I
N
G
N
O
.
2
0
0
S
I
E
V
E
A
F
T
E
R
C
O
M
P
A
C
T
I
O
N
(HEGAZY ET AL., 2004)
10
AVERAGE 4 PERCENT
INCREASE IN SAMPLE
PASSING NO. 200 SIEVE
AFTER COMPACTION
FIGURE 6.36 PARTICLE-SIZE DISTRIBUTION OF COARSE COAL REFUSE AFTER COMPACTION
FIGURE 6.36 PARTICLE-SIZE DISTRIBUTION OF COARSE COAL REFUSE AFTER COMPACTION
< PREVIOUS VIEW
6-109
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
tive density are determined in accordance with ASTM 4254, Standard Test Methods for Minimum
Index Density and Unit Weight of Soils and Calculation of Relative Density.
For coal refuse embankments, the relative density test is normally applicable to materials used for
granular drainage and flter zones that require compaction. Relative density tests are also conducted
for evaluating the capability of a saturated, coarse-grained soil to resist seismic loadings. Seismic
issues are discussed in Chapter 7.
The minimum density (zero percent relative density) in accordance with ASTM 4254 is obtained
by placing dried soil as loosely as possible in a mold of known volume. The preferred method for
placing soil in the mold requires using a pouring device that limits the height of free fall to no more
than inch. The weight of the known volume of soil is then measured and used in the determina-
tion of the minimum test dry density. The minimum density is the weight of soil divided by the
volume of the mold.
The maximum dry density (100 percent relative density) in accordance with ASTM 4253 can be
obtained using either dry soil (method A) or wet soil (method B). The soil is densifed using either an
electromagnetic, vertically vibrating table or an eccentric or cam-driven, vertically vibrating table. If
method A is used, dry soil is placed in a mold using a scoop or funnel, a surcharge base plate is placed
atop the level soil surface, the flled mold with surcharge weight is atached to the vibrating table, and
the table is vibrated for 8 to 12 minutes depending on the frequency of vibration.
If method B is used, the mold is atached to the vibrating table, the table is turned on, and wet soil
is placed in the mold over a 5- to 6-minute period during which care is taken to avoid excessive
vibration that causes the soil to boil excessively. The table is then turned of, a surcharge is placed
atop the soil, and the table is vibrated for 8 to 12 minutes as for Method A. Afer the table is turned
of (both methods), the surcharge is removed and the height of the sample in the mold is measured.
The material in the mold is then weighed. If the sample is wet, it is oven dried and weighed again
afer drying. The weight of the known volume of dried soil is used to determine the maximum dry
density. The diference between the wet and dry weights can be used to determine the moisture
content of the material tested.
The relative density is calculated by the relationship:
D
r
=
d
max
(
d
min
)
(6-11)
d
(
d
max
d
min
)
x
100
where:
D
r
= relative density (dimensionless)
d
max
= maximum dry density (force/length
3
)
dmin
= minimum dry density (force/length
3
)
d
= measured dry density of the sample (laboratory or in situ) (force/length
3
)
The relative density D
r
is used as a basis to confrm whether placement of coarse-grained soils in the
feld meets the minimum specifed compaction criteria. The minimum acceptable relative density for
coarse-grained soils typically ranges between about 70 to 85 percent depending upon performance
requirements.
6.5.4 Hydraulic Conductivity Tests
The hydraulic conductivity (or permeability) of a soil is a measure of the rate at which water will fow
through a soil under a particular pressure (or head). It is represented by the coefcient of hydraulic
< PREVIOUS VIEW
6-110
Chapter 6
MAY 2009
conductivity k, which is normally expressed in units of distance per time. Hydraulic conductivity is
essentially the volume of fow per unit time per unit of cross-sectional area for a unit pressure gradi-
ent. While laboratory measurement of hydraulic conductivity can be valuable in developing design
criteria, feld tests are generally more representative of in-situ materials if site conditions and access
allow performance of the tests.
Hydraulic conductivity is measured in the laboratory by percolating water through a soil sample of
known cross-sectional area and length. The constant head hydraulic conductivity test is conducted
for coarse-grained soils in accordance with ASTM D 2434, Standard Test Method for Permeabil-
ity of Granular Soils (Constant Head). This test method is most suitable for granular soils with a
hydraulic conductivity greater than 10
-4
cm/sec that might be used for drains or flter media and
that do not have more than 10 percent passing a No. 200 sieve. Soils tested using this procedure
are typically compacted in a permeameter to a density comparable to the relative density used for
feld placement.
For soils that have a hydraulic conductivity less than about 10
-4
cm/sec, laboratory hydraulic con-
ductivity tests should be conducted in accordance with ASTM D 5084, Standard Test Methods for
Measurement of Hydraulic Conductivity of Saturated Porous Materials Using a Flexible Wall Perme-
ameter. This standard permits hydraulic conductivity testing by either the constant- or falling-head
methods and is suitable for soils with hydraulic conductivities not less than about 10
-9
cm/sec. Thus,
the method and equipment are suitable for testing a wide range of soils. The test specimen is sealed
within a fexible membrane and enclosed within a pressure cell similar to that used for triaxial test-
ing (Section 6.5.7.4). This setup permits back pressuring to saturate the test sample and application
of high hydraulic pressures (gradients) needed for testing low hydraulic conductivity soils and fne
coal refuse within a reasonable time frame of several days to a few weeks.
Table 6.8 shows the probable range of hydraulic conductivity for various USCS soil classifcation
groups. For coal refuse, hydraulic conductivity data are less well documented. Based upon a very
limited number of feld tests and observations, hydraulic conductivity values for coarse coal refuse
range from 10
-6
to 10
-2
cm/sec (about 1.0 to 10,000 f/year). Hegazy et al. (2004) report the results of
falling head and rising head slug tests that were performed to estimate the hydraulic conductivity
of the coarse refuse at coal refuse disposal facilities in western Pennsylvania. The average horizon-
tal hydraulic conductivity k
h
was 3 x 10
-5
cm/sec (about 30 f/year), and the standard deviation and
coefcient of variation were 2.7 x 10
-5
and 0.9, respectively. Density is inversely related to hydraulic
conductivity; thus equipment trafcking across the refuse surface and weathering due to exposure
following placement tend to increase the density and lower the hydraulic conductivity. These envi-
ronmental and construction processes can result in a vertical hydraulic conductivity on the order of
10 times less than the horizontal hydraulic conductivity, as discussed in Section 6.6.2.1. For impound-
ing coal refuse disposal facilities, it is important to be aware of the efects of density and grain-size
distribution on the hydraulic conductivity of the refuse materials.
As with coarse refuse, the hydraulic conductivity of fne coal refuse varies greatly and is difcult
to predict. Typical values based on the results of piezocone dissipation tests performed at northern
Appalachian sites are presented in Table 6.36. The range of estimated k
h
indicates that fne coal
refuse behaves similarly to very fne sands, silts and mixtures of sand, silt and clay. Some designers
have reported the presence of moderate plasticity clay with very low vertical hydraulic conductiv-
ity at refuse disposal sites. In general, conservative assumptions should be made relative to the
hydraulic conductivity of coal refuse, with consideration of feld test data when available. Gener-
ally, conservative values are assumed based upon judgment and hydraulic conductivity values
from the high end of the range determined from available test data, material classifcation and
representative anisotropy values, resulting in either higher (more conservative) phreatic levels or
seepage rates.
< PREVIOUS VIEW
6-111
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
6.5.5 Geosynthetic Materials Tests
Geosynthetic materials are polymeric sheet materials used with soil, rock, or other geotechnically-
related material as an integral part of a civil engineering project, structure, or system. Geotextiles,
geomembranes and geosynthetic clay liners (GCLs) are types of geosynthetic materials that may be
used in the design and construction of a coal refuse disposal facility to convey or limit seepage or to
act as a flter medium.
A geotextile is a permeable geosynthetic made of textile materials. At refuse disposal facilities, geo-
textiles are used as flters in drainage applications, as well as for material separation applications
such as beneath spillway linings and haul roads.
Geomembranes are continuous polymeric sheets with very low hydraulic conductivity (typically less
than 10
-12
cm/sec) in contrast to GCLs, which are sheets of very low hydraulic conductivity, composite
barrier material. The geomembrane polymeric types used for refuse disposal applications include: (1)
chlorinated polyethylene (CPE), (2) chlorosulfonated polyethylene (CSPE), also called Hypalon, (3)
high-density polyethylene (HDPE) and (4) polyvinyl chloride (PVC). Of these types, PVC and HDPE
are the most commonly used because they are lowest in cost and widely available.
GCLs consist of dry bentonite clay soil between two geotextiles or on a geomembrane carrier. Geo-
membranes and GCLs are manufactured in sheets and delivered to the site in rolls. The geotextiles
used above and below the dry clay may or may not be connected with threads or fbers to increase
the in-plane strength. Geomembrane rolls are seamed in the feld using thermal methods or solvents,
and GCLs are overlapped in the feld to create a continuous barrier. At refuse disposal sites, geo-
membranes and GCLs are used as hydraulic barriers to limit seepage from impoundments into the
groundwater and underground mines.
Although not required by MSHA, some state agencies that regulate coal refuse disposal facilities may
require linings to control seepage. While placement of a low-hydraulic conductivity, compacted soil
liner is permited, some sites have insufcient material for constructing such a liner. Therefore, geo-
membrane and geosynthetic clay liners are more commonly used at these sites.
Leakage, rather than hydraulic conductivity, is the primary design concern for geomembrane-lined
containment structures. Leakage can occur through poor feld seams, poor factory seams, pinholes
from manufacture, and puncture holes from handling, placement, or in-service activity. Leakage of
geomembrane liner systems can be minimized by specifcation of an appropriate liner material and
implementation of QA/QC procedures during installation.
TABLE 6.36 ESTIMATED HORIZONTAL HYDRAULIC CONDUCTIVITY OF FINE
COAL REFUSE BASED ON PIEZOCONE DISSIPATION TESTS
Test
Depth
(m)
t
50
(seconds)
c
h
(cm
2
/s)
k
h
(cm/s)
PCPT1 7.9 800 1510
-3
1.2110
-5
PCPT1 19.2 30,000 0.410
-3
0.0310
-5
PCPT1 22.9 40 30110
-3
24.310
-5
PCPT3 13.3 128 9410
-3
7.5910
-5
PCPT4 21.9 300 4010
-3
3.2410
-5
Note: Pool level was approximately 3 m below the ground surface.
(HEGAZY ET AL., 2004)
< PREVIOUS VIEW
6-112
Chapter 6
MAY 2009
The quality of geosynthetic material installation can be controlled by testing. Tables 6.37, 6.38 and 6.39
identify applicable quality control test methods published by ASTM or the Geosynthetic Research
Institute (GRI) for geotextiles, geomembranes and GCLs, respectively. Laboratory testing such as the
gradient ratio test described in ASTM D 5101, Standard Test Method for Measuring the Soil-Geo-
textile System Clogging Potential by the Gradient Ratio, can be employed to check for clogging of
geotextiles used for fltration. Geotextile Filter Performance via Long Term Flow (LTF) Tests (GRI
Method GT1) should also be considered. In the design of a slurry impoundment, if a geotextile is to
be used instead of a granular flter in a location where clogging of the geotextile would adversely
afect the safety of the embankment, testing should be performed with site-specifc materials to dem-
onstrate that signifcant clogging of the geotextile will not occur during the design life of the flter.
Additionally, as discussed in Section 6.6.2.3.2, sufcient feld instrumentation to monitor the phreatic
level near critical drain and flter installations is recommended.
6.5.6 Consolidation Tests
Applying loads to coal refuse and underlying foundation materials causes compressive strains that
are either immediate or time-dependent. Immediate strains are usually the result of elastic deforma-
tion of solids and compression of voids that are not held open temporarily by trapped pore water.
Time-dependent strain is referred to as consolidation and is a function of the movement of pore water.
Immediate strain is common to soils with a low degree of saturation and/or a high hydraulic con-
ductivity. This type of deformation is usually not an important part of the design of an embankment
because the resulting movements are generally complete by the end of embankment construction.
Situations where immediate strain should be considered include:
Horizontal and vertical movement of a pipe within the embankment that could cause
opening of joints, cracking of the pipe material or changes in the slope of drainage
pipes. Special pipes are available that are fexible or have joints that allow longitudi-
nal movement and slight bending.
Movement beneath rigid structures, such as concrete spillways, that can be cracked
or hydraulically afected by diferential movement.
The behavior of saturated, fne-grained soils under stress can be very important in disposal facility
design. Compressive strain in these soils can occur only through drainage of water from the pores
because, at the stresses experienced during construction, both the soil particles and the pore water are
essentially incompressible. Therefore, most of the strain occurs slowly, possibly over several months
or years. This slow compressive strain is termed consolidation. With saturated fne-grained soils, the
major portion of the total strain is due to consolidation. Until the drainage occurs, the excess pore-
water pressures can reduce the efective strength of the material and cause instability.
Consolidation can be an important design consideration because, in addition to causing damage to
pipes and structures, it can afect the gradient of surface drainage structures and the integrity of a cap
following abandonment. The problems it creates may not become apparent until afer the disposal
facility begins operation, when corrections are most expensive. If an embankment is constructed over
sof clay deposits or previously setled fne refuse, the movements resulting from consolidation can
be especially large and the excess pore-water pressures can cause instability. Additional problems
that can result from consolidation include:
Setlement of the embankment crest below the design elevation
Diferential setlement that disrupts internal drains
Cracking of the embankment, particularly in areas where large diferential setle-
ments occur over small distances, such as where sof foundation materials abut
harder soil or rock at the base of a valley wall
< PREVIOUS VIEW
6-113
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
TABLE 6.37 METHODS FOR QUALITY CONTROL TESTING OF GEOTEXTILES
Description Test Method
Standard Test Method for Biological Clogging of Geotextiles or Soil/Geotextile Filters ASTM D 1987
Standard Practice for Sampling of Geosynthetics for Testing ASTM D 4354
Standard Test Methods for Water Permeability of Geotextiles by Permittivity ASTM D 4491
Standard Test Method for Trapezoid Tearing Strength of Geotextiles ASTM D 4533
Standard Test Method for Tensile Properties of Geotextiles by the Wide-Width Strip Method ASTM D 4595
Standard Test Method for Grab Breaking Load and Elongation of Geotextiles ASTM D 4632
Test Method for Determining the (In-plane) Flow Rate per Unit Width and Hydraulic
Transmissivity of a Geosynthetic Using a Constant Head
ASTM D 4716
Standard Test Method for Determining Apparent Opening Size of a Geotextile ASTM D 4751
Standard Test Method for Index Puncture Resistance of Geotextiles, Geomembranes and
Related Products
ASTM D 4833
Standard Test Method for Measuring the Soil-Geotextile System Clogging Potential by the
Gradient Ratio
ASTM D 5101
Standard Test Method for Permittivity of Geotextiles Under Load ASTM D 5493
Standard Test Method for Hydraulic Conductivity Ratio (HCR) Testing of Soil/Geotextile
Systems
ASTM D 5567
Standard Test Method for Biological Clogging of Geotextile of Soil/Geotextile Filters ASTM D 1987
Geotextile Filter Performance via Long Term Flow (LTF) Tests GRI Test Method GT1
Fine Fraction Filtration Using Geotextile Filters GRI Test Method GT8
Restrictions on the rate of fll placement over previously deposited fnes
Setlement of abandoned facilities that disrupts cap integrity and positive surface
drainage control
The conventional laboratory test for measuring the consolidation characteristics of a fne-grained
soil consists of trimming an approximately 1-inch-thick undisturbed soil sample in a metal ring that
prevents lateral expansion. This test is described in ASTM D 2435, Standard Test Methods for One-
Dimensional Consolidation Properties of Soils Using Incremental Loading. Porous stones above and
below the sample allow excess pore water to drain from the soil as it is subjected to a series of load
and unload cycles. Application of a constant vertical load to the sample permits measurement of com-
pression with time. When compression stops, the sample is said to be 100 percent consolidated under
the applied load. The load is subsequently removed and the sample undergoes a small rebound. The
load is then increased and held constant until the compression stops again. This procedure is contin-
ued for several cycles of loading and unloading, and a relationship is developed between the applied
load and the compression produced. It is convenient to express the relationship between sample
compression and load, as shown in Figure 6.37, such that the logarithm of the applied load is ploted
on the horizontal axis and the compression is ploted on the vertical axis in terms of the void ratio of
the sample (as the sample is compressed, the void ratio decreases).
< PREVIOUS VIEW
6-114
Chapter 6
MAY 2009
TABLE 6.38 METHODS FOR QUALITY CONTROL TESTING OF GEOMEMBRANES
Description
ASTM Test
Method
Standard Practice for Sampling of Geosynthetics for Testing D 4354
Standard Practice for Determining the Integrity of Field Seams Used in Joining Flexible Polymeric
Sheet Geomembranes
D 4437
Standard Practice for Determining the Integrity of Factory Seams Used in Joining Manufactured
Flexible Sheet Geomembranes
D 4545
Standard Test Method for Index Puncture Resistance of Geotextiles, Geomembranes and Related
Products
D 4833
Standard Test Method for Determining Performance Strength of Geomembranes by the Wide Strip
Tensile Method
D 4885
Standard Test Method for Determining the Coeffcient of Soil and Geosynthetic or Geosynthetic and
Geosynthetic Friction by the Direct Shear Method
D 5321
Standard Test Method for the Determination of Pyramid Puncture Resistance of Unprotected and
Protected Geomembranes
D 5494
Standard Test Method for Large Scale Hydrostatic Puncture Testing of Geosynthetics D 5514
Standard Test Method for Multi-Axial Tension Test for Geosynthetics D 5617
Standard Practice for Geomembrane Seam Evaluation by Vacuum Chamber D 5641
Standard Practice for Pressurized Air Channel Evaluation of Dual Seamed Geomembranes D 5820
Standard Test Method for Determining Tearing Strength of Internally Reinforced Geomembranes D 5884
Standard Guide for Selection of Test Methods to Determine Rate of Fluid Permeation Through
Geomembranes for Specifc Applications
D 5886
Standard Test Method for Measuring Core Thickness of Textured Geomembrane D 5994
Standard Test Method for Determining the Integrity of Field Seams Used in Joining Geomembranes
by Chemical Fusion Methods
D 6214
Standard Practice for the Nondestructive Testing of Geomembrane Seams using the Spark Test D 6365
Standard Test Method for Determining the Integrity of Non-reinforced Geomembrane Seams
Produced Using Thermo-Fusion Methods
D 6392
Standard Guide for the Selection of Test Methods for Flexible Polypropylene (FPP) Geomembranes D 6434
Standard Guide for Selection of Techniques for Electrical Detection of Potential Leak Paths in
Geomembrane
D 6747
Standard Practice for Leak Location on Exposed Geomembranes Using the Water Puddle System D 7002
Standard Test Method for Strip Tensile Properties of Reinforced Geomembranes D 7003
Standard Test Method for Grab Tensile Properties of Reinforced Geomembranes D 7004
Standard Practice for Ultrasonic Testing of Geomembranes D 7006
< PREVIOUS VIEW
6-115
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
As shown in Figure 6.37, the steeper slope of the void ratio versus log efective stress plot is defned as
the compression index C
c
. The fater slope of the void ratio versus log efective stress plot is defned
as the recompression index C
cr
. Typically, C
cr
is in the range of 0.1 to 0.2 C
c
. The compression index
is typically used for estimating consolidation setlement of normally consolidated soils and fnes.
Procedures for determining C
c
and C
cr
are presented in ASTM D 2435. Presenting the test results in
the form shown in Figure 6.37 simplifes computation of the estimated compression for a given incre-
ment of applied load. Most standard texts on soil mechanics include comprehensive discussions of
this test and procedures for using the test data to predict setlement.
Consolidation testing also provides information regarding the time rate of setlement as a function of
consolidation stress. The time rate of consolidation setlement is defned by the coefcient of vertical
consolidation c
v
and the coefcient of horizontal consolidation c
h
. The parameter c
v
can be used for
estimating vertical pore pressure dissipation with time, which is useful for evaluating the rate of con-
solidation setlement of flls that are large in areal extent. The parameter c
h
can be used for estimat-
ing horizontal pore pressure dissipation with time, which is important in the design of wick drains.
Procedures for determining c
v
and c
h
are presented in ASTM D 2435. Values for c
v
and c
h
determined
from laboratory consolidation results tend to be conservative (i.e., underpredict the rate of pore-
pressure dissipation) and can vary signifcantly from in-situ values. More reliable values of c
v
and c
h
can usually be obtained by conducting a dissipation test during piezocone testing (Section 6.4.3.7) or
by monitoring piezometers.
Description
ASTM Test
Method
Standard Practices for Electrical Methods for Locating Leaks in Geomembranes Covered with
Water or Earth Materials
D 7007
Standard Test Method for Determining the Tensile Shear Strength of Pre-Fabricated Bituminous
Geomembrane Seams
D 7056
Standard Specifcation for Non-Reinforced Polyvinyl Chloride (PVC) Geomembranes Used in Buried
Applications
D 7176
Standard Specifcation for Air Channel Evaluation of Polyvinyl Chloride (PVC) Dual Track Seamed
Geomembranes
D 7177
Standard Practice for Leak Location using Geomembranes with an Insulating Layer in Intimate
Contact with a Conductive Layer via Electrical Capacitance Technique (Conductive Geomembrane
Spark Test)
D 7240
TABLE 6.38 METHODS FOR QUALITY CONTROL TESTING OF GEOMEMBRANES
(CONTINUED)
TABLE 6.39 METHODS FOR QUALITY CONTROL TESTING OF GEOSYNTHETIC CLAY LINERS
Description
ASTM Test
Method
Standard Guide for Storage and Handling of Geosynthetic Clay Liners D 5888
Standard Test Method for Determining the Internal and Interface Shear Resistance of
Geosynthetic Clay Liner by the Direct Shear Method
D 6243
Standard Guide for Acceptance Testing Requirements for Geosynthetic Clay Liners D 6495
Standard Test Method for Tensile Strength of Geosynthetic Clay Liners D 6768
< PREVIOUS VIEW
6-116
Chapter 6
MAY 2009
Coarse coal refuse is normally placed at a moisture content below saturation and is usually sufciently
coarse-grained that consolidation is not a consideration. On the other hand, consolidation of fne coal
refuse is important, especially in the design of disposal facilities developed using the upstream method
of construction. Consolidation efects should be considered if the incremental increase in height of an
embankment constructed on setled fne coal refuse is greater than several feet and the embankment
supports drainage structures or seepage barriers that could be impacted by diferential setlement. Con-
solidation parameters may also be estimated from the results of triaxial shear strength tests.
Published data for typical ranges of consolidation parameters for coal refuse are limited. Almes and
Butail (1976) report that C
c
for saturated fne coal refuse varies between 0.2 and 0.3 at moisture contents
ranging between 30 and 45 percent. Hegazy et al. (2004) present values for the horizontal coefcient
of consolidation c
h
determined from piezocone testing in fne coal refuse deltas at disposal sites in
western Pennsylvania. The values of c
h
range from 15 x 10
-3
cm
2
/sec to 300 x 10
-3
cm
2
/sec. These results
are typical of coefcient of consolidation values reported for sandy silt to silty clay soils (Bardet, 1997).
6.5.7 Shear Strength and Related Tests
The shear strengths of soil and coal refuse materials used to construct an embankment, or used as the
embankment foundation, are needed for stability analysis of the embankment. Stability analyses are
discussed more extensively in Section 6.6.4. Shear strengths are also needed for determination of the
allowable bearing pressure for structures founded on or within the embankment and for the stability
of slopes cut during embankment construction.
Embankment and foundation stability may be evaluated using either total stress or efective
stress methods. The method selected depends on the:
Embankment material or materials
Foundation conditions
FIGURE 6.37 RESULTS OF CONSOLIDATION TEST ON FINE COAL REFUSE
0.1 1.0 10 100
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
V
O
I
D
R
A
T
I
O
,
e
EFFECTIVE CONSOLIDATION STRESS, v (TSF)
e
i
e
o
= e
m
Cc
1-cycle
Ccr
BORING B-1 DEPTH 7.8'
Ccr
1+eo
= 0.003
Cc = 0.23
Ccr = 0.015
OCR = 1
INITIAL STATE w = 49.3%
d = 79.73 PCF
CONSOLIDATION PARAMETERS
eo = 0.81 vo = 0.38 TSF
ei = 0.905 vm = 0.38 TSF
em = 0.81 Cc
= 0.127
1+em
vo
=
vm
= 0.38 TSF
1-cycle
FIGURE 6.37 RESULTS OF CONSOLIDATION TEST ON FINE COAL REFUSE
FIGURE 6.37 RESULTS OF CONSOLIDATION TEST ON FINE COAL REFUSE
< PREVIOUS VIEW
6-117
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
Magnitude of pore-water pressures within the embankment
State of construction or use for which embankment stability is to be evaluated
Total stress is a combination of the stress between the individual soil grains, termed efective stress,
and the pressure of the pore water, termed pore pressure. Because pore water has no shear resis-
tance, all shear resistance is represented by the efective stress. The shear stress at failure (ultimate
shear strength) on any surface within an embankment is directly related to the stress normal to the
failure surface, because the failure mechanism involves friction of one body moving on another and
apparent bonding. This relationship can be expressed as:
max
= c + ( - u) tan = c + tan (6-12)
where:
max
= shear stress on surface at failure (force/length
2
)
c = efective cohesion (force/length
2
)
= angle of efective internal friction (degrees)
= total stress acting normal to the failure surface (force/length
2
)
u = pore-water pressure acting on the failure surface (force/length
2
)
= efective stress acting normal to the failure surface (force/length
2
)
The following paragraphs describe the above two approaches and their application to refuse embank-
ment design.
For shear strength tests conducted for a total stress analysis, water is not allowed to drain from the
sample during shearing. This method of stability analysis and related types of analyses (Section 6.6.4)
are generally considered most appropriate for evaluating relatively short-term conditions that would
occur: (1) during and immediately afer construction, (2) immediately following rapid changes in
the impoundment level, (3) during pushouts, and (4) during seismic loadings. Tests typically used
to develop strength parameters for a total-stress analysis include the vane shear test, the unconfned
compression test and the unconsolidated-undrained (UU) triaxial test. For these tests, the strength
does not increase with increasing normal stress if the soil is saturated. Total stress analysis param-
eters can also be calculated from the consolidated-undrained (CU) triaxial shear test.
Shear strength tests for an efective-stress analysis either allow water to drain from the samples
during testing or provide for measurement of the pore pressures under loading and confning
conditions that are intended to simulate actual feld conditions. Efective-strength parameters
apply to all soil types, including gravels, sands, silts, and clays. This method of stability analysis
and related types of analyses (Section 6.6.4) are generally considered most appropriate for evalu-
ating long-term conditions afer the temporary efects of construction on pore pressures have
dissipated and seepage rates become steady. The tests typically used to develop efective-stress
strength parameters include: (1) consolidated-undrained triaxial shear tests with pore-water pres-
sure measurements (CU), (2) consolidated-drained triaxial tests at slow strain rates (CD), or (3)
drained direct shear tests. For these tests, the strength increases with increasing normal stress. For
long-term analyses, the drained test strength parameters are the efective cohesion intercept c and
efective friction angle from the efective stress Mohr-Coulomb envelope. The shear strength
max
is given by:
max
= c + tan (6-13)
< PREVIOUS VIEW
6-118
Chapter 6
MAY 2009
For analysis purposes, c is ofen assumed to be zero because laboratory tests are afected by loading
rate and duration efects. In this situation, the cohesion component of strength can be likened to a
bond that weathers with time (Mesri and Abdel-Ghafar, 1993).
Sample preparation for laboratory shear strength tests is an essential aspect of a testing program,
if the tests are to accurately refect feld conditions. Undisturbed, disturbed and remolded or com-
pacted samples may be tested depending on the soil and material conditions to be modeled in the
total- or efective-stress analyses. For shear strength tests on proposed embankment construction
materials, the materials are compacted to the densities and moisture contents that are anticipated to
occur within the embankment. Tests for the shear strength of coarse-grained foundation soils are per-
formed on samples that are reconstituted in the laboratory to simulate the in-situ conditions. Tests for
the shear strength of fne-grained foundation soils are performed on undisturbed samples obtained
from borings or test pits.
Triaxial and direct shear test results are presented either as a series of Mohrs circles or stress paths
that refect sample failure conditions. A Mohrs circle presentation is typically used for UU triaxial
and direct shear test results, while a stress path presentation is used for CU and CD triaxial test
results. An example Mohrs circle presentation is shown in Figure 6.38, and a stress path presentation
is provided in Figure 6.39. As shown in Figure 6.38, a Mohrs circle for each test is drawn on a plot of
shear stress versus normal stress by drawing a circle that connects the maximum and minimum
principal stresses on the sample at failure (i.e.,
1f
and
3f
, respectively). Failure is then defned by a
straight or curved line drawn tangent (or nearly tangent) to the series of circles on the plot. The inter-
FIGURE 6.38 EFFECTIVE STRESS MOHR'S CIRCLE FOR CU TRIAXIAL TEST
0 100 200 300 400 500 600 700 800
0
100
200
300
400
500
600
EFFECTIVE NORMAL STRESS, ' (KPA)
S
H
E
A
R
S
T
R
E
S
S
,
(
K
P
A
)
'
3f
'
1f
MOHR-COULOMB PARAMETERS
c' = 0 ' = 34
'
tan ' = 0.675
FIGURE 6.38 EFFECTIVE STRESS MOHRS CIRCLE FOR CU TRIAXIAL TEST
FIGURE 6.38 EFFECTIVE STRESS MOHRS CIRCLE FOR CU TRIAXIAL TEST
< PREVIOUS VIEW
6-119
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
cept of the line with = 0 is referred to as the total stress cohesion intercept c and the angle of the line
to the horizontal is the total stress angle of friction .
As shown in Figure 6.39, the stress path for each test is constructed on a plot of (
1
+
3
)/2, or p, versus
(
1
-
3
)/2, or q, where values of p and q are ploted for each load increment in each test. Failure is
defned by a straight or curved line connecting values of (q/p)max for each stress path. The intercept
of this line with the p axis (a) can be represented as a = c cos , where the angle to the horizontal of
the line connecting values of (q/p)max for each stress path is and tan = sin . Then the efective
angle of friction = arcsin (tan ) and c = a/cos .
Many soils exhibit stress-strain behavior that varies with confnement. This behavior is referred to as
stress dependency and can be characterized by the stress path method. A stress path is a numerical
and graphical representation of the past, present and future state of stress on a representative soil
element because it captures the geologic stress history of the element, the current stresses acting on
the element, and the anticipated future changes in stress on the element. The stress path of a material
is determined by ploting the efective strength from CU and CD triaxial tests for each load incre-
ment of the tests. Using the stress path method, the test results are then analyzed with respect to the
approximate feld stress and strain conditions before, during, and afer construction (Lambe, 1967;
Lambe and Marr, 1979).
Determining the appropriate strength parameters for evaluating the stability of any embankment,
regardless of size or location, should be performed by a person experienced in the engineering behav-
ior of soil, rock and refuse materials. The complexity of laboratory shear strength test procedures for
modeling expected feld conditions requires that laboratory shear strength testing be conducted with
FIGURE 6.39 EFFECTIVE STRESS PATHS AND FAILURE
ENVELOPE FOR CU TRIAXIAL TEST
p (PSI)
q
(
P
S
I
)
0 5 10 15 20 25 30 35 40 45 50 55 60
0
2
4
6
8
10
12
14
16
18
20
22
24
26
28
30
32
34
sin ' = tan
c' = a/cos '
a
1
-
3
p =
1
+
3
2
FIGURE 6.39 EFFECTIVE STRESS PATHS AND FAILURE ENVELOPE FOR CU TRIAXIAL TEST
FIGURE 6.39 EFFECTIVE STRESS PATHS AND FAILURE ENVELOPE FOR CU TRIAXIAL TEST
< PREVIOUS VIEW
6-120
Chapter 6
MAY 2009
greater care and more professional scrutiny than routine tests. Shear strength testing must be tailored
to site conditions by a qualifed geotechnical engineer familiar with the type of embankment to be
constructed and foundation conditions. The tests should be conducted by laboratories with appropri-
ate equipment and skilled technicians.
Standard shear strength test methods are: (1) vane-shear, (2) direct-shear, (3) unconfned-compres-
sion, and (4) triaxial-compression. The applicability of these tests to various soil types is presented in
Table 6.40. Descriptions of these tests are provided in the following sections. Head (1982, 1986) and
Bardet (1997) discuss other methods for determining shear strength.
6.5.7.1 Vane-Shear Test
The laboratory vane-shear test is used to determine the undrained shear strength s
u
using the test
method described in ASTM D 4648, Standard Test Method for Laboratory Miniature Vane Shear
Test for Saturated Fine-Grained Clayey Soil. Similar to the feld vane-shear test (Section 6.4.3.8), the
laboratory vane-shear test is conducted on very sof to stif, fne-grained, undisturbed, remolded or
reconstituted, cohesive soil by inserting a four-bladed vane into a soil sample and rotating it such
that shearing occurs along a cylindrical surface. The undrained shear strength is determined from the
resistance to rotation. The miniature vane is similar to the feld vane-shear device, except that it has a
smaller blade diameter (0.5 inch) and blade height (1 inch). Afer s
u
is determined, the residual (mini-
mum) shear strength s
ur
is determined by quickly rotating the vane 10 full rotations (to fully remold
the soil) and then conducting a second shear test. The ratio of peak to remolded undrained shear
strength is the sensitivity S
t
. The laboratory vane test is typically conducted on a vertically oriented
sample because that is the direction in which the soil sample is taken in the feld. If the sample is
rotated 90 degrees from the vertical, the laboratory vane test can be used to measure soil anisotropy.
Laboratory vane shear testing of fne coal refuse is not recommended. Instead, the strength of fne
coal refuse should be determined in situ using the CPT, PCPT methods or the feld vane-shear test,
as described in Section 6.4.3.8, or by laboratory testing using the direct-shear or triaxial-compression
test methods described in Sections 6.5.7.2 and 6.5.7.4, respectively.
6.5.7.2 Direct-Shear Test
The direct-shear test is the oldest and simplest form of shear test. Direct-shear tests are used for testing
reconstituted cohesionless soils and undisturbed cohesive soils. The test method is particularly useful
if the residual strength at large strain is desired. The test is conducted in accordance with ASTM D
3080, Standard Test Method for Direct Shear Test of Soils under Consolidated Drained Conditions.
As shown schematically in Figure 6.40, the direct-shear test is performed by placing a -inch-mini-
mum-thickness specimen into a cylindrical (2-inch-minimum-diameter) or square-shaped (typically
3 or 4 inches) shear box that is split along a horizontal plane. The test specimen is confned top and
botom by porous stones, and the shear box is placed in a container to permit submergence and
saturation of the specimen during testing. A vertical (normal) load is applied over the specimen and
allowed to consolidate. The test is conducted by holding the upper or lower part of the box stationary
and applying a horizontal load on the other part of the box to shear the specimen along a predefned
horizontal plane. The shearing load applied at failure divided by the cross-sectional area of the soil
sample is considered to be the shear stress at failure. The normal load divided by the cross-sectional
area of the sample is considered to be the normal stress at failure. Direct shear tests of cohesionless
soils are considered drained tests because of the high sample hydraulic conductivity. Depending on
the rate of shearing, direct-shear tests of cohesive soils can be either undrained or drained.
Afer the maximum shear strength has been determined, the residual shear strength (c
r
and
r
) can
be determined by performing repeated and rapid cycles of shearing (usually a minimum of fve full
forward and reverse cycles) along the plane of failure mobilized during the initial portion of the test.
< PREVIOUS VIEW
6-121
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
The repeated cycles of loading are intended to simulate large straining in the feld that would be typical
of a slope failure. Once the cycles of repeated shearing are complete, excess pore pressures in the test
specimen are allowed to dissipate under constant normal load. When the excess pore-water pressures
equilibrate (i.e., consolidation is complete), the test specimen is sheared as previously described.
As shown in Figure 6.41a, a series of direct-shear tests (typically three minimum) is conducted
using varying normal stress . Test results are ploted in the form of shear stress versus horizontal
displacement . A plot of the peak or failure shear stress versus is used to determine the angle
of internal friction and cohesion intercept, as shown in Figure 6.41b. For most soils, a line drawn
through the points for each test is approximately straight. This line is termed the failure envelope.
The angle that the line makes with the horizontal axis is a measure of the component of strength due
TABLE 6.40 LABORATORY TESTS FOR DETERMINING SOIL SHEAR STRENGTH
Type of Test
Relative Frequency of Use
(1)
Preparation
Prior to
Applying
Load
Drainage
Conditions
During
Test
Parameters
Determined
Remarks
Coarse-
grained
Soils
Fine-
grained
Soils
Coarse
Refuse
Fine
Refuse
Direct Shear
Drained 1 3 1 1
Consolidated
Under
Normal
Load
Drained
Effective
Stress
Diffcult
to control
rate of test
to assure
drained
condition
Undrained NA 3 NA NA
Consolidated
Under
Normal
Load
Undrained
Approximate
Total Stress
Diffcult to
conduct
quickly
enough to
assure no
drainage
Triaxial
Unconsolidated-
Undrained (UU)
NA 2 NA 3
Uncon-
solidated
Undrained
Approximate
Total Stress
Also called
Quick (Q)
test
Consolidated-
Undrained (CU)
1 1 1 1
Consolidated
Under
Isotropic
Pressure
Undrained
Total
Stress and
Effective
Stress
Pore
pressures
are
measured
to give
effective
stress
condition
Consolidated-
Drained (CD)
2 1 2 1
Consolidated
Under
Isotropic
Pressure
Drained
Effective
Stress
Also called
Slow (S)
test
Unconfned NA 2 2 NA Unconsolidated Undrained
Approximate
Total Stress
Sample
must have
suffcient
cohesion
to maintain
shape
without
support
Note: 1. 1 = frequently used 3 = applicable, but seldom used
2 = occasionally used NA = not applicable
< PREVIOUS VIEW
6-122
Chapter 6
MAY 2009
FIGURE 6.40 DIRECT SHEAR TESTING ARRANGEMENT
T
P
H
(MAYNE ET AL., 2002)
PISTON
LOWER FRAME
POROUS STONE
POROUS STONE
UPPER FRAME
FIGURE 6.40 DIRECT SHEAR TESTING ARRANGEMENT
to friction between the soil particles and is termed the angle of efective internal friction . When
the line intercepts the vertical axis at a value greater than zero, the intercept is referred to as the
efective cohesion c.
Direct-shear tests are simple and can be performed relatively quickly. However, the test has several
inherent shortcomings due to the forced plane of shearing:
The failure plane is predefned and horizontal and may not be the weakest plane in
the sample.
There is litle control over the drainage of the soil.
The height cannot be defned for calculating shear strains, so a stress-strain modulus
cannot be determined from the test.
Stress conditions on the failure surface are non-uniform and failure develops pro-
gressively (the entire strength of the specimen is not mobilized simultaneously at all
points on the failure surface), so measured strength values are lower than would be
obtained under uniform stress conditions.
Stress conditions are known only at failure.
6.5.7.3 Unconfned-Compression Test
The unconfned-compression test is conducted on cohesive soils in accordance with ASTM D 2166,
Standard Test Method for Unconfned Compressive Strength of Cohesive Soil. Cohesive soil speci-
mens are tested to failure by rapidly applying an axial load. Measurements of axial force and axial
deformation are made during the test, and the test results are presented as a plot of axial stress versus
axial strain as shown in Figure 6.42. The maximum measured force over the sample cross section q
u
is the axial stress, and the peak value of axial stress divided by 2 is the undrained shear strength s
u
.
For a total stress analysis, the unconfned-compression test provides an approximate measure of the
short-term, undrained strength of the soil at the density and moisture content of the sample. It pro-
vides no information about long-term strength properties appropriate for an efective stress analysis.
FIGURE 6.40 DIRECT SHEAR TESTING ARRANGEMENT
< PREVIOUS VIEW
6-123
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
NOTE: DATA BASED UPON DIRECT SHEAR TEST
OF TRIASSIC CLAY IN RALEIGH, NC
EFFECTIVE NORMAL STRESS ' (KPA)
S
H
E
A
R
S
T
R
E
S
S
,
(
K
P
A
)
6.41b SHEAR STRESS VS. EFFECTIVE NORMAL STRESS
FIGURE 6.41 RESULTS OF DRAINED DIRECT SHEAR TEST ON CLAY
(MAYNE ET AL., 2002)
DISPLACEMENT, (MM)
S
H
E
A
R
S
T
R
E
S
S
,
(
K
P
A
)
0 1 2 3 4 5 6 7 8 9 10
0
20
40
60
80
100
120
140
' = 214.5 KPA
' = 135.0 KPA
' = 39.3 KPA
6.41a SHEAR STRESS VS. DISPLACEMENT
0 50 100 150 250 200
0
20
40
60
80
100
120
140
STRENGTH PARAMETERS: c' = 0, ' = 26.0
o
tan ' = 0.488
'
FIGURE 6.41 RESULTS OF DRAINED DIRECT SHEAR TEST ON CLAY
FIGURE 6.41 RESULTS OF DRAINED DIRECT SHEAR TEST ON CLAY
< PREVIOUS VIEW
6-124
Chapter 6
MAY 2009
The unconfned-compression test can be performed using undisturbed, remolded or compacted soil
samples. The stress-strain curves and failure modes observed during testing provide an index value
of the soil properties in addition to strength. For example, bulging or yielding of the sample signifes
a relatively sof clay, while a sudden britle failure indicates a desiccated clay or cemented material.
The stress-strain curves developed from these tests should be used with caution when determining
the soil modulus for input to numerical analyses (e.g., fnite element analysis) because they are very
sensitive to minor variations of the modulus.
Test specimens with inclined fssures, sand and silt lenses or slickensides have a tendency to fail pre-
maturely along these weaker planes in unconfned compression tests. If these failure modes occur,
more sophisticated testing, such as triaxial tests, may be needed to obtain a more realistic determina-
tion of the in-situ strength.
6.5.7.4 Triaxial-Compression Test
The triaxial test is used to determine strength and stress-strain behavior of undisturbed, remolded
or reconstituted soil samples. To conduct a triaxial test, cylindrical samples are consolidated, usually
isotropically, and then sheared in axial compression. Undrained and drained testing can be con-
ducted, and pore-water pressures can be measured during undrained shear tests. Triaxial tests pro-
FIGURE 6.42 UNCONFINED COMPRESSION TEST RESULTS
AXIAL STRAIN,
a
(PERCENT)
A
X
I
A
L
S
T
R
E
S
S
,
a
(
K
P
A
)
0 1 2 3 4 5 6
0
1
2
3
4
5
6
(MAYNE ET AL., 2002)
UNDRAINED SHEAR STRENGTH
cu = su = 51 KPA
qu = 102 KPA = 2cu
FIGURE 6.42 UNCONFINED COMPRESSION TEST RESULTS
FIGURE 6.42 UNCONFINED COMPRESSION TEST RESULTS
< PREVIOUS VIEW
6-125
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
vide a reliable means for determining: (1) the undrained strength of cohesive soils, (2) the angle of
friction and cohesion intercept of undisturbed, reconstituted and compacted soils, and (3) the soil
modulus at intermediate to large strains.
Test specimens are typically 1.4 to 2.8 inches in diameter with a height to width ratio between 2 and
2.5. Selection of the sample diameter is governed by limitation of the maximum particle size in the
test specimen to not more than one-sixth of the sample diameter. Thus, the maximum particle size for
a 1.4-inch-diameter sample is about inch and for a 2.8-inch-diameter specimen about inch. If the
soil to be tested has large-size particles, then sufciently large-diameter specimens should be used
so that the sample diameter is more than six times the maximum particle size. Otherwise, the tested
sample should be modeled or scalped using the procedures recommended by Becker et al. (1972) and
summarized in Duncan and Wright (2005)
Modeling entails creating a modeled particle-size distribution that parallels the original gradation,
where the maximum particle size does not exceed one-sixth of the diameter of the tested sample.
Using this approach, Becker et al. (1972) determined that strength test results using the modeled gra-
dation were essentially the same as the strength of the original gradation that had been tested using
sufciently large diameter samples to meet the one-sixth criterion provided the test specimens were
prepared to the same relative density (Section 6.5.3.2).
Scalping entails using that portion of the sample that remains afer sieving to remove particle sizes
that exceed the one-sixth criterion. As with modeling, Becker et al. (1972) determined that strength
test results using a scalped gradation were essentially the same as the strength of the original grada-
tion provided that the test specimens were prepared to the same relative density. However, if either
modeling or scalping is used to achieve an acceptable particle-size distribution for testing, additional
testing to determine the minimum and maximum densities of both the original and modeled or
scalped materials must be conducted to verify that the relative density of the modeled or scalped
material is approximately equal to the relative density of the original material.
Of the options for achieving an acceptable particle-size distribution for testing, scalping is probably
the simpler and less costly approach. Scalping also does not result in an appreciable shifing of the
fnes content, which could afect the strength test results if a substantial portion of over-size mate-
rial must be removed for testing. While these procedures can be used for coal refuse, they should be
applied with caution, particularly in cases where the characteristics (e.g., material type, angularity,
surface roughness) of the smaller particles difer materially from those of the larger particles that
were removed from the sample.
To conduct the triaxial test, a sample is enclosed by a thin rubber membrane and placed inside a
cylindrical pressure chamber that is usually flled with water. The sample is subjected to a uniform
confning pressure
3
by compression of the fuid in the chamber acting on the membrane. The range
of
3
for a triaxial test series is generally selected so that the confning pressures are higher, lower and
about equal to anticipated in-situ value of
3
at the end of construction. Using a range of
3
where all
confning pressures are less than or equal to the anticipated in-situ value of
3
can result in overesti-
mation of the strength and compressibility of the tested material, as compared to the in-situ material.
Depending on the type of triaxial test conducted, a backpressure may be applied to the specimen
through the end platens to saturate the specimen. The test sample is sheared to failure by applying an
axial stress, typically referred to as the deviator stress (
1
-
3
), through a vertical loading ram. Axial
stress can be applied at a constant deformation rate (strain controlled) or by means of dead weight
increments or hydraulic pressure (stress controlled) until the sample fails.
Triaxial tests can be used to simulate various in-situ loading conditions. The types of triaxial tests
typically employed for this purpose include:
< PREVIOUS VIEW
6-126
Chapter 6
MAY 2009
Unconsolidated-undrained (UU or Q) test
Consolidated-undrained (CU test)
Consolidated undrained (CU or R) test with pore-pressure measurement
Consolidated-drained (CD or S) test
UU (also referred to as quick or Q) triaxial tests are conducted on cohesive soils in accordance with
ASTM D 2850, Standard Test Method for Unconsolidated-Undrained Triaxial Compression Test on
Cohesive Soils. The UU test provides only an approximation of the short-term or undrained strength
for a total stress analysis and typically is conducted in order to provide data for preliminary analy-
ses and for designing the fnal test program. The UU test does not provide data about the efective
stress or long-term properties of the material. In a UU test, the specimen is not allowed to consolidate
during application of
3
or to drain during the testing, so the strength measured is the undrained
shear strength s
u
. The rate of axial deformation during shear is comparable to the rate used for uncon-
fned compression tests. The results of undrained tests depend on the degree of saturation S
r
of the
specimens. If S
r
100 percent, testing of similar samples will provide similar values of s
u
, because the
shear strength of the test sample will not increase with increasing confning pressure. However, if S
r
95 percent, increasing
3
may result in increasing values of s
u
until the air voids compress and the
sample becomes completely saturated.
If water is allowed to completely drain from a test sample when
3
is applied, the sample becomes
uniformly consolidated. Two types of tests can be performed on consolidated samples. In the con-
solidated-undrained (CU) test, the consolidated sample is sheared by application of an axial test load
without allowing any additional water to drain during the loading. Because the sample does not
drain during loading, the results are suitable only for total stress analyses. The rate of strain used for
the CU triaxial test is similar to that for the UU triaxial test.
If the pore pressure is measured while the sample is loaded, efective stress parameters can be calcu-
lated. This test is called a consolidated undrained with pore-pressure measurement (CU or R) triaxial
test. The CU triaxial test permits determination of both total stress and efective stress (c and ) param-
eters. The rate of strain used for the CU triaxial test is much slower than for the UU or CU triaxial tests
so that excess pore pressures equilibrate throughout the test specimen and pore-water pressures can be
reliably measured at the ends of the test specimen. The CU test is performed in accordance with ASTM
D 4767, Standard Test Method for Consolidated Undrained Triaxial Compression Test for Cohesive
Soils. The rate of strain used for the CU triaxial test is prescribed in ASTM D 4767 and is much slower
than the rate of strain for UU triaxial tests. This permits equalization of excess pore pressures during
undrained shearing so that the excess pore-water pressures measured at the end of the test specimen
are not less than 95 percent of the excess pore-water pressure along the sample shear plane.
Consolidated-drained (CD) triaxial (also referred to as slow or S) tests also yield the efective stress
parameters c and . The primary diference between the CU and CD triaxial tests is that the sample
is allowed to drain during the CD test. The rate of strain used for CD tests is usually much slower
than the rate used for CU tests, so that the development of excess pore pressures in the test sample is
less than a few percent of the initial efective confning pressure
3
. The CD test measures only the
efective stress parameters c and .
Triaxial test results are typically presented in Mohrs Circle or p-q plots, as described previously.
Typical values of the shear strength from triaxial testing of fne coal refuse, as reported by Hegazy
et al. (2004) for sites in northern Appalachia, are summarized in Table 6.41. Drained-shear-strength
parameters were determined using consolidated isotropic undrained compression (CIUC) triaxial
tests with pore-pressure measurements and consolidated isotropic drained compression (CIDC) tri-
< PREVIOUS VIEW
6-127
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
axial tests. In this study, fne coal refuse samples from Shelby tubes were collected from beneath the
upstream stages of a refuse embankment or from working platforms built over the fne coal refuse
in the impoundment. Table 6.41 indicates that the variability of is low to moderate, while the vari-
ability of c is relatively high. The shear strength parameters presented in the table are peak values
and were found to decrease with increasing fnes contents, as shown in Figure 6.43.
Residual shear strength values at large strains determined in accordance with the cited ASTM stan-
dards are sometimes considered for the design of impoundments with upstream construction.
The test results in Table 6.41 and Figure 6.43 are presented for illustration purposes only, and strength
characteristics will vary depending on the geology and coal extraction, processing, and disposal
practices.
TABLE 6.41 SUMMARY OF FINE COAL REFUSE SHEAR STRENGTH
PARAMETERS BASED ON TRIAXIAL TEST RESULTS
Parameter Average Standard Deviation Coeffcient of Variation
33 degrees 4 0.12
c 11 kPa 14 1.30
(c = 0) 35 degrees 4 0.11
(HEGAZY ET AL., 2004)
0 20 40 60 80 100
25
30
35
40
45
' = -0.09(%-#200) + 40.5
r
2
= 0.25
PERCENT FINES PASSING NO. 200 SIEVE
E
F
F
E
C
T
I
V
E
A
N
G
L
E
O
F
I
N
T
E
R
N
A
L
F
R
I
C
T
I
O
N
(
D
E
G
R
E
E
S
)
FIGURE 6.43 VARIATION OF STRENGTH OF FINE COAL REFUSE
WITH INCREASING FINES CONTENT
(HEGAZY ET AL., 2004)
FIGURE 6.43 VARIATION OF STRENGTH OF FINE COAL REFUSE WITH FINES CONTENT
FIGURE 6.43 VARIATION OF STRENGTH OF FINE COAL REFUSE WITH FINES CONTENT
< PREVIOUS VIEW
6-128
Chapter 6
MAY 2009
6.5.8 Seismic Property Characterization
Laboratory testing for determination of the seismic properties of soil or refuse materials generally
involves the evaluation of potential strength loss associated with earthquake loading. This section
presents testing methods for: (1) cyclic-triaxial testing, (2) cyclic loading followed by monotonic load-
ing, and (3) resonant-column testing. Chapter 7 presents details of seismic stability and deforma-
tion analyses, including the application of the laboratory strain-based approach developed by Castro
(1994), sometimes referred to as the residual- or steady-state-strength approach. Section 7.4.3.2 pres-
ents guidance for laboratory testing and application of the laboratory tests below.
6.5.8.1 Cyclic-Triaxial Test
The cyclic-triaxial test can be used to evaluate the cyclic strength (or liquefaction potential) of primar-
ily cohesionless, free-draining soils in undrained shear. The samples tested are either undisturbed, or
they are reconstituted to simulate the relative density of the in-situ soil. The test apparatus consists of
a regular triaxial cell and a cyclic (ofen sinusoidal) loading machine atached to the loading piston.
The sample is isotropically consolidated in the triaxial cell and then subjected to a cyclic axial load in
extension and compression. The cyclic loading generally causes an increase in pore-water pressure
and a decrease in the efective confning pressure with increasing cyclic deformation of the sample.
Failure occurs when the excess pore-water pressure equals the initial efective confning pressure
(sometimes called initial liquefaction) or when some limiting cyclic or permanent strain is mobilized.
Details regarding the test method are described in ASTM D 5311, Standard Test Method for Load
Controlled Cyclic Triaxial Strength of Soil.
There are limitations to use of cyclic triaxial tests for representing feld conditions during earthquake
loading, including:
Non-uniform stress conditions imposed on the test sample by the end platens can
cause a redistribution of the void ratio.
There would be a continuous reorientation of the principal stresses in the feld
whereas the reorientation angle is either 0 or 90 degrees in the laboratory test.
The laboratory test sample is isotropically consolidated, whereas the material
sampled would be in an at-rest lateral earth pressure (K
o
) condition in the feld (i.e.,
lateral stress = K
o
times vertical stress).
Cyclic shear stress is applied on a horizontal plane in the feld but on a 45-degree
plane in the triaxial test.
The mean normal stress in the feld is constant while the mean normal stress in the
laboratory varies cyclically.
Despite these limitations, the cyclic triaxial test has been used with reasonable success since the early
1960s.
6.5.8.2 Cyclic Loading Followed by Monotonic Loading
The purpose of this type of testing is to evaluate the post-earthquake, residual, steady-state, und-
rained strength of clay-like materials for post-earthquake stability analyses. A limited number of
loading cycles is applied to a test sample to model the straining induced by earthquake loading.
Monotonic loading is then applied in order to determine the post-earthquake strength. The initial
portion of the test consists of a cyclic triaxial test (Section 6.5.8.1) followed monotonic loading using
the CU triaxial test (Section 6.5.7.4). Selection of the cyclic stress ratio and number of cycles to be
applied during cyclic loading depends on requirements discussed in Section 7.4.3.2. Typically, there
is a holding period between the end of cyclic loading and the beginning of monotonic loading to
permit equilibration of excess pore-water pressures so that measurement of pore-water pressures at
< PREVIOUS VIEW
6-129
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
the ends of the sample during CU testing are representative of pore-water pressures along the sample
shear plane. In reality, the time needed to transition between cyclic and CU testing is sufcient to
permit pore-water pressure equilibration in the sample. Application of this type of testing program
for seismic design is discussed in Section 7.4.3.2.
6.5.8.3 Resonant-Column Test
Evaluation of the response of foundation and embankment soils to seismic ground amplifcations
requires information regarding shear modulus (G
max
or G
o
) and damping D. While feld geophysical
methods such as the crosshole, downhole, and surface wave techniques can provide direct in-situ
measurements of shear wave velocity (Section 6.4.4), the resonant-column test permits an evaluation
of the variation (decrease) of shear modulus with increasing shear strain
s
and the increase of D with
s
under controlled efective stress states. The test may yield lower values than those obtained from
feld testing due to the efect of soil aging.
The resonant-column test is conducted in accordance with ASTM D 4015, Standard Test Methods for
Modulus and Damping of Soils by the Resonant-Column Method. The undisturbed or reconstituted
test specimen is sealed in a fexible membrane and enclosed in a pressure cell similar to that used for
triaxial testing (Section 6.5.7.4). This setup permits the use of back pressure to saturate the test speci-
men. The resonant-column device excites one end of the test sample in a fundamental mode of vibra-
tion by means of torsional or longitudinal motion. Both solid and hollow specimens can be used in
the apparatus. Either a sinusoidal torque or a vertical compressional load is applied to the top of the
sample through the top cap. The deformation of the top of the sample is measured, and the excitation
frequency is adjusted until the sample resonates. The wave velocity or modulus is computed from the
resonant frequency and the geometric properties of the sample and driving apparatus. Damping is
determined by switching of the current to the driving coil at resonance and recording the amplitude
of decay of the vibrations. The decay of the amplitude with time is used to determine the logarithmic
decrement (the percentage decay over one log cycle of time), which is directly related to the viscous
damping ratio. Typical test results are presented in Figure 6.44. Figure 6.44a shows the decrease in
shear modulus with increasing strain amplitude, and Figure 6.44b shows the increase in damping
with increasing strain amplitude for a clay soil.
The resonant-column test is generally limited to small to intermediate shear strains by the applied
force requirements and resonant frequencies. At larger strains, hollow samples must be used to main-
tain a relatively constant shear strain across the sample. For these reasons, resonant column testing is
primarily used to estimate shear modulus associated with small strains.
6.5.9 Rock Property Tests
Common laboratory tests for engineering properties of intact rocks and index testing of rock frag-
ments include measurements of strength, stifness, and durability. Table 6.42 presents a summary list
of ASTM standards and procedures for laboratory rock testing. Additional discussion of the testing
of coal is presented in Section 8.4.2.2.
6.5.9.1 Point-Load Index Test
Determination of rock strength is typically determined in the laboratory using specially prepared
rock core and specialized test equipment. Because of the extensive sample preparation and equip-
ment requirements, the point-load test was developed so that rock specimens from drilled core,
cut blocks or irregular lumps could be tested using portable equipment suitable for the feld or
laboratory. The point-load test is conducted using an apparatus (Figure 6.45) that applies a con-
centrated load through a pair of spherically truncated, conical platens. The distance between the
opposing specimen-platen contact points is recorded, and the load is steadily increased until the
specimen fractures and the failure load is recorded. The point-load test is conducted in accor-
dance with ASTM D 5731, Standard Test Method for Determination of the Point Load Strength
< PREVIOUS VIEW
6-130
Chapter 6
MAY 2009
FIGURE 6.44 TYPICAL RESULTS FOR RESONANT COLUMN TEST ON FINE COAL REFUSE
FIGURE 6.44 TYPICAL RESULTS FOR RESONANT COLUMN TEST
ON FINE COAL REFUSE
10
-5
10 10
-4
10
-3
10
-2
10
-1
0
0.2
0.4
0.6
0.8
1.2
1.0
N
O
R
M
A
L
I
Z
E
D
S
H
E
A
R
M
O
D
U
L
U
S
G
/
G
M
A
X
SHEAR STRAIN,
(%)
e = 0.80
c
= 1 KG/CM
2
, GMAX = 280 KG/CM
2
c
= 2 KG/CM
2
, GMAX = 450 KG/CM
2
c
= 4 KG/CM
2
, GMAX = 714 KG/CM
2
c
= 5.5 KG/CM
2
, GMAX = 842 KG/CM
2
RANGE OF
RESULTS
(ELLISON AND CHO, 1976)
6.44a DECREASE IN SHEAR MODULUS WITH STRAIN
0
2
4
6
8
10
12
e = 0.80
c
= 1 KG/CM
2
c
= 2 KG/CM
2
c
= 4 KG/CM
2
c
= 5.5 KG/CM
2
SHEAR STRAIN,
(%)
D
A
M
P
I
N
G
,
D
(
%
)
10
-5
10 10
-4
10
-3
10
-2
10
-1
6.44b INCREASE IN DAMPING WITH STRAIN
FIGURE 6.44 TYPICAL RESULTS FOR RESONANT COLUMN TEST ON FINE COAL REFUSE
< PREVIOUS VIEW
6-131
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
Index of Rock. The test is used to classify and characterize rock that has a compressive strength
greater than 2,200 psi.
ASTM recommends that test samples conform to size and shape requirements. In general, for diam-
etral tests, core specimens with a length-to-diameter ratio of 1.0 are adequate, while for axial tests,
core specimens with length-to-diameter ratio of 0.3 to 1.0 are suitable. Specimens for the block and
the irregular lump test should have a length of 50 35 mm and a length-to-width ratio between 0.3
and 1.0 (preferably close to 1.0). Samples are typically tested at their natural moisture content.
Size corrections are applied to obtain the point-load strength index I
s(50)
of a rock specimen. A strength
anisotropy index I
a(50)
is determined when I
s(50)
values are measured perpendicular and parallel to
TABLE 6.42 STANDARDS AND PROCEDURES FOR LABORATORY TESTING OF INTACT ROCK
Test Category Name of Test
ASTM Test
Method
Point Load
Strength
Standard Test Method for Determination of the Point Load Strength Index of
Rock and Application to Rock Strength Classifcation
D 5731
(1)
Compressive
Strength
Standard Test Method for Compressive Strength and Elastic Moduli of Intact
Rock Core Specimens under Varying States of Stress and Temperatures
D 7012
(1)
Creep Tests
Standard Test Method for Creep of Rock Core Under Constant Stress and
Temperature
D 7070
Tensile Strength
Standard Test Method for Direct Tensile Strength of Intact Rock Core
Specimens
D 2936
Standard Test Method for Splitting Tensile Strength of Intact Rock Core
Specimens
D 3967
(1)
Direct Shear
Standard Test Method for Performing Laboratory Direct Shear Strength Tests of
Rock Specimens under Constant Normal Force
D 5607
(1)
Hydraulic Cond.
Standard Test Method for Permeability of Rocks by Flowing Air D 4525
Durability
Standard Test Method for Slake Durability of Shales and Similar Weak Rocks D 4644
(1)
Standard Test Method for Testing Rock Slabs to Evaluate Soundness of Riprap
by Use of Sodium Sulfate or Magnesium Sulfate
D 5240
Standard Test Method for Evaluation of Durability of Rock for Erosion Control
under Freezing and Thawing Conditions
D 5312
Standard Test Method for Evaluation of Durability of Rock for Erosion Control
under Wetting and Drying Conditions
D 5313
Deformation
and Stiffness
Standard Test Method for Compressive Strength and Elastic Moduli of Intact
Rock Core Specimens under Varying States of Stress and Temperatures
D 7012
Standard Test Method for Laboratory Determination of Pulse Velocities and
Ultrasonic Elastic Constants of Rock
D 2845
Specimen
Preparation
Standard Practices for Preparing Rock Core as Cylindrical Test Specimens and
Verifying Conformance to Dimensional and Shape Tolerances
D 4543
Standard Practice for Preparation of Rock Slabs for Durability Testing D 5121
Note: 1. The rock test procedure associated with this ASTM standard is described in the Chapter 6 text. Additional
discussion of the testing of coal is presented in Section 8.4.2.2.
(ADAPTED FROM MAYNE ET AL., 2002)
< PREVIOUS VIEW
6-132
Chapter 6
MAY 2009
planes of weakness. The test can be performed in the feld or in the laboratory. The point-load index
is used to estimate the unconfned compressive strength q
u
using a relationship of the form:
q
u
= K I
s(50)
(6-14)
The value of the constant K has been reported to vary from 15 to 50 (especially for anisotropic rocks)
depending upon the specifc rock formation. Rusnak and Mark (2000) determined that K 21 for
rocks associated with coal seams in the eastern, mid-western and western U.S. Additional discussion
of the limitations of testing of coal is presented in Section 8.4.2.2.
6.5.9.2 Unconfned Compressive Strength Test
The unconfned compressive strength serves as an initial index of the competency of intact rock and
represents the most direct method for determining rock strength. The test method is described in
ASTM D 7012, Standard Test Method for Compressive Strength and Elastic Moduli of Intact Rock
Core Specimens under Varying States of Stress and Temperatures. In this test, cylindrical rock speci-
mens are tested in compression without lateral confnement. The test procedure is similar to the
unconfned compression test for soils and concrete. The test specimen should be a rock cylinder of
length-to-width ratio in the range of 2 to 2.5 and should have fat, smooth, and parallel ends cut per-
pendicular to the cylinder axis in accordance with ASTM D 4543, Standard Practices for Preparing
Rock Core Specimens and Determining Dimensional and Shape Tolerances. The peak stress during
unconfned loading is the uniaxial compressive strength q
u
. The results may be afected by: (1) mois-
ture content, (2) rate of loading, (3) condition of both ends of the rock sample, and (4) the presence of
FIGURE 6.45 POINT-LOAD TEST APPARATUS
FLEXIBLE
HYDRAULIC
HOSE
PRESSURE
GAGE
ROCK CORE
BEING TESTED
LOADING
FRAME
CONICAL
ENDS
HYDRAULIC
JACK
FIGURE 6.45 POINT-LOAD TEST APPARATUS
GRADUATED
SCALE
(ADAPTED FROM MAYNE ET AL., 2002)
FIGURE 6.45 POINT-LOAD TEST APPARATUS
< PREVIOUS VIEW
6-133
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
inclined fssures, intrusions, and other anomalies that may cause premature failures on the associated
planes. These conditions should be noted so that other tests such as triaxial or direct shear tests can be
performed, as appropriate. Because the tests are of necessity conducted using intact rock samples, the
test results may not be representative of rock mass behavior due to the efects of in-situ discontinuities
(e.g., bedding planes and joints), weathering and moisture. For these reasons, unconfned compressive
strength testing should be limited to rock types and strata where tests on intact rock are reasonably
representative of the rock mass. This is especially a problem in determining the compressive strength
of coal through laboratory testing of small-diameter core (e.g., 2 to 3 inches) because the presence of
defects in the coal results in unconservative strengths as compared to the strengths determined from
tests on larger-diameter core (e.g., 6 to 12 inches) and in-situ tests on feld-scale test volumes (Pariseau,
2006). Additional discussion of the limitations of testing of coal is presented in Section 8.4.2.2.
The stress-strain behavior of intact rock samples can be measured during an unconfned-compres-
sion test in accordance with ASTM D 7012, Standard Test Method for Compressive Strength and
Elastic Moduli of Intact Rock Core Specimens under Varying States of Stress and Temperatures. For
this test, specimen deformations are measured using strain gauges applied to the test specimen or
linear voltage displacement transducers atached to the top and botom load platens.
6.5.9.3 Indirect Tensile Strength Test
Rock is relatively weak in tension; thus, the tensile strength To of intact rock is approximately 5 per-
cent of its compressive value (Mayne et al., 2002). Because of the difculties involved in proper end
preparation (Jaeger et al., 2007), the direct tensile strength testing of rock specimens is not a common
laboratory procedure. Therefore, the tensile strength of rock is usually determined by indirect meth-
ods such as the indirect tensile (Brazilian) test. The indirect tensile strength of intact rock core
T
is
determined in accordance with ASTM D 3967, Standard Test Method for Spliting Tensile Strength
of Intact Rock Core Specimens. Core specimens with length-to-diameter ratios between 2 and 2.5
are placed in a compression loading machine with the load platens placed diametrically across the
specimen. The maximum load to fracture the specimen is recorded and used to calculate the indirect
tensile strength.
Alternatives to the indirect tensile strength test are the direct tensile strength test and the bending test.
The direct tensile strength of intact rock core is determined in accordance with ASTM D 2936, Stan-
dard Test Method for Direct Tensile Strength of Intact Rock Core Specimens. The core specimens for
direct tensile testing are prepared as described in ASTM D 4543 for compressive strength testing, but
the test is conducted by cementing the specimen ends to the test load apparatus and applying tensile
loads on the sample until it fails in tension. While there is no ASTM standard for bending tests, the
test specimens need to be long relative to their thickness which makes their preparation difcult and
expensive, especially when samples are taken from very jointed rock strata (Pariseau, 2006).
Thus, the indirect tensile strength test is signifcantly more convenient and practical for routine
measurements than the direct tensile strength test and the bending test, and it provides very similar
results to those obtained from direct tension tests (Jaeger et al., 2007). In many situations, the indi-
rect tensile test provides a more fundamental measurement of rock material strength than compres-
sion tests because the failure mode is more representative of failures that occur (e.g., overstressing
of roof strata) in underground mines. It should be noted that the point-load index test discussed in
Section 6.5.9.1 is actually a variation of the indirect tensile strength test with results correlated to
compressive strength.
6.5.9.4 Rock Durability
Rock used for slope protection and aggregate used for drainage purposes must have high durability
to atain suitable long-term performance when exposed to construction and in-service environments.
< PREVIOUS VIEW
6-134
Chapter 6
MAY 2009
Material deterioration mechanisms include cracking, spalling, delaminating or spliting, disaggregat-
ing, dissolving and disintegration (USACE, 1990a). The following ASTM test methods can be used to
evaluate the durability of these types of materials:
D 5240, Rock Slab Testing For Riprap Soundness, Using Sodium/Magnesium Sulfate
D 5312, Rock-Durability For Erosion Control Under Freezing/Thawing
D 5313, Rock-Durability For Erosion Control Under Weting/Drying
Additional rock durability test methods that have been used for evaluation of the slake durability of
surface mine spoils and earthen embankments constructed using shale and other slake prone sedi-
mentary rocks are described in Andrews et al. (1980) and Michael and Superfesky (2007). Some of
these test methods involve use of acid solutions.
Rock materials used for constructing earthen dam and coal refuse embankments are generally
obtained from borrow sources near the construction area. Because coal is found in geologic set-
tings where shales and other weak rocks are encountered as part of mining operations, the dura-
bility of these rock materials as compacted fll needs to be evaluated. The most problematic rock
types are certain shales, mudstones, claystones and other weak rocks that degrade rapidly soon
afer they are exposed to atmospheric conditions. These materials can degrade rapidly, afecting
the stability of an embankment fll, a rock cut, or foundation on which a refuse embankment is
constructed.
The slake durability test was devised in the early 1970s (Franklin and Chandra, 1972) to provide
a qualitative measure of these materials in service. As described in ASTM D 4644, Standard Test
Method for Slake Durability of Shales and Similar Weak Rocks, representative rock fragments are
subjected to cycles of weting and drying, and the weight loss is measured afer two cycles. Figure
6.46 provides an illustration of the slake durability test apparatus.
FIGURE 6.46 ROTATING DRUM ASSEMBLY AND SETUP
OF SLAKE DURABILITY EQUIPMENT
1
4
0
M
M
4
0
M
M
2
0
M
M
100 MM
WATER-FILLED TROUGH
DRUM CONTAINING
SAMPLE
(MAYNE ET AL., 2002)
FIGURE 6.46 ROTATING DRUM ASSEMBLY AND SETUP OF SLAKE DURABILITY EQUIPMENT
FIGURE 6.46 ROTATING DRUM ASSEMBLY AND SETUP OF SLAKE DURABILITY EQUIPMENT
< PREVIOUS VIEW
6-135
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
Ten dried fragments of rock of known weight are placed in a drum fabricated with 2.0-mm-square
mesh wire cloth. The drum is rotated in a horizontal position along its longitudinal axis while par-
tially submerged in distilled water to promote weting of the sample. The specimens and the drum
are dried at the end of the rotation cycle (10 minutes at 20 rpm) and weighed. Afer two cycles of
rotating and drying, the weight loss and the shape and size of the remaining rock fragments are
recorded, and the slake-durability index (SDI) is calculated. As a qualitative measure of durability,
Gamble (1971) proposed the classifcation system presented in Table 6.43 using the SDI determined
afer two cycles. Both the SDI and the description of the shape and size of the remaining particles are
used to determine the durability of sof rocks. The application of the test for design and construction
with mine spoil is presented in Section 8.8.2.
Limestone and calcite cemented sandstone can degrade relatively quickly where acidic leachates are
present and should not be used in the construction of flters, underdrains and internal drains. A fzz
test of aggregate and rock can be conducted by placing a dilute solution of hydrochloric acid on the
material. If an efervescent or fzzing reaction occurs, the material is generally considered to be unac-
ceptable. Laboratory testing may also be performed on aggregate materials to evaluate the calcium-
carbonate portion in terms of percent of total content. The presence of sulfde minerals in aggregates
can lead to the formation of sulfuric acid and sulfate minerals and should also be avoided. Generally,
aggregate materials should contain less than 1 to 5 percent calcium carbonate and much less than 1
percent sulfde materials in order to be acceptable for drain applications.
6.5.10 Structural Material Tests
This section identifes the quality control tests that should be considered in developing plans and
specifcations for disposal facilities when using portland cement concrete and controlled low-strength
material (CLSM).
6.5.10.1 Concrete
Portland-cement concrete is used at coal refuse disposal facilities for construction of drainage control
structures and structure foundations. Structural design of reinforced concrete is beyond the scope of
this Manual. For additional guidance, the most current version of ACI 318, Building Code Require-
ments for Structural Concrete and Commentary, should be used. Table 6.44 presents ASTM test
methods for quality control of fresh portland-cement concrete.
6.5.10.2 Controlled Low-Strength Materials (CLSM)
CLSM is used as a replacement for compacted earth fll and consists of various mixtures of pozzolan
(e.g., fy ash), portland cement, aggregates (typically fne aggregate such as sand and/or botom ash),
select cohesionless soil, water, and occasionally chemical admixtures (Howard and Hitch, 1998). A
low percentage of bentonite or atapulgite can also be included in CLSM for reduced hydraulic con-
ductivity and enhanced plasticity, where such characteristics are important. CLSM has been used for
general backflls, shallow cutof trenches, structural flls, insulating and isolating flls, pavement bases,
erosion control, pipe bedding, cradles and backfll, and void flling. It is the later two applications
that have found the greatest application in mining and mine refuse disposal. CLSM is processed and
mixed much like fresh concrete and delivered as a fuid in ready-mix trucks or as a dry material in a
dump truck. The quality control test methods used for CLSM are adapted from those used for fresh
concrete and grout, accounting for the various constituents in the CLSM mix and its lower compres-
sive strength (i.e., 1,200 psi or less). Table 6.45 presents test methods used for quality control of CLSM.
CLSM must be designed in accordance with performance criteria appropriate for the specifc appli-
cation. At coal refuse disposal facilities, CLSM is used primarily for bedding and backflling of pipe.
When pipe is used as part of a decant or other structure that extends through the embankment cross
section, the bedding and backfll should have adequate strength to provide support for the pipe and
< PREVIOUS VIEW
6-136
Chapter 6
MAY 2009
low hydraulic conductivity and suitable stifness and shrinkage characteristics to limit the potential
for seepage along the pipe. The bedding should also perform as intended relative to the desired pipe
behavior (i.e., fexible versus rigid system). Typical CLSM mix designs have a target maximum com-
pressive strength in the range of 50 to 200 psi for fexible pipe installations and up to 1,200 psi for struc-
tural applications. When CLSM is used in fexible pipe installations, a lower strength CLSM is specifed
in order to produce high quality, soil-like bedding, cradle, and backfll zones to preserve the fexible
behavior of the pipe-backfll system while capitalizing on the benefts of CLSM (i.e., ease of placement,
beter protection of the pipe under construction trafc loadings, improved haunch support and seepage
control due to more intimate contact with the pipe, resistance to piping and erosion, and beter qual-
ity control). Testing of the uniaxial strength of CLSM samples for design of fexible pipe installations
should follow ASTM C109 with the addition of strain measurements for determination of the modulus
of elasticity in accordance with ASTM D7012. Table 6.46 provides a mix design guide for CLSM. It
should be noted that the CLSM strength is also dependent on the cement, ash and mixing water charac-
teristics. Application-specifc design and testing is recommended for use of CLSM as pipe backfll, with
strengths evaluated at 7, 14 and 28 days (similar to concrete) and also at 56 and 90 days.
TABLE 6.45 QUALITY CONTROL STANDARDS FOR CLSM
Description
ASTM Test
Method
Standard Test Method for Preparation and Testing of Controlled Low Strength Material (CLSM)
Test Cylinders
D 4832
Standard Practice for Sampling Freshly Mixed Controlled Low-Strength Material D 5971
Standard Test Method for Unit Weight, Yield, Cement Content, and Air Content (Gravimetric) of
Controlled Low Strength Material (CLSM)
D 6023
Standard Test Method for Flow Consistency of Controlled Low Strength Material (CLSM) D 6103
TABLE 6.43 SLAKE DURABILITY CLASSIFICATION
Durability Classifcation Percent Retained after Two 10-Minute Cycles
Very High > 98
High 95 to 98
Medium High 85 to 95
Medium 60 to 85
Low 30 to 60
Very Low < 30
(GAMBLE, 1971)
TABLE 6.44 QUALITY CONTROL STANDARDS FOR PORTLAND-CEMENT CONCRETE
Description
ASTM Test
Method
Standard Test Method for Compressive Strength of Cylindrical Concrete Specimens C 39
Standard Test Method for Slump of Hydraulic Cement Concrete C 143
Standard Test Method for Air Content of Freshly Mixed Concrete by the Volumetric Method C 173
Standard Test Method for Air Content of Freshly Mixed Concrete by the Pressure Method C 231
< PREVIOUS VIEW
6-137
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
6.5.11 Verifcation of Test Results
Unverifed or unvalidated test data or engineering properties (e.g., parameters that are assumed
from published or other available sources or adopted from other sites without evaluation of their
appropriateness) should not be used in the design of refuse disposal facilities. The only exceptions
are analyses for facilities with very simple embankments with low hazard potential and for embank-
ments where very conservative slopes are acceptable. Therefore, it is important to verify test data,
particularly those for shear strength.
If site access, time and budget were not an issue, the design engineer could obtain as many samples
as deemed necessary and conduct as many laboratory or in-situ tests as desired to obtain a complete
assessment of subsurface soil and rock conditions. Engineering properties would be quantifed, and
any inconsistent data would be set aside and additional testing would be initiated, as needed. Unfortu-
nately, site access, schedule and budgets are major factors that designers must consider in making criti-
cal decisions throughout the design process to obtain the most reliable and realistic soil, rock and coal
refuse property data. As described previously, a critical step in determining these properties lies in the
selection of specifc tests and proper interpretation of the test results. For a number of reasons (e.g., cost,
sampling difculties or lack or representative values), it may be difcult to determine values for specifc
parameters of interest. Fortunately, designers can sometimes use established and/or site-specifc cor-
relations to obtain values for the desired parameters. Correlations are useful for evaluating test results
that do not appear to be representative, and they can also be very useful in preparing test programs,
performing preliminary calculations, and serving as a quality assurance check on the test results.
Engineering property correlations come in many forms, but all have a common theme. Specifcally, a
useful correlation is developed from a large database of results based on past experience. In the best
case, the correlation and experience have been developed or calibrated using specifc site construc-
tion materials, or the correlation may be based on reportedly similar construction materials. The
reliance upon or use of correlations to obtain soil, rock and coal refuse properties may be justifed
TABLE 6.46 CLSM MIX DESIGN GUIDE
Properties and Criteria Type A Type B Type C Type D
Mix Design
(1)
(per yd
3
)
Cement (lb)
Pozzolans (lb)
Bottom ash or coarse aggregate
or fne aggregate (lb)
Air entrainment
100
2000
0
50
300
2600
150 to 200
300
2600
300 to 700
100 to 400
(2)
Slump (in) ASTM C 143 7 (min.) 7 (min.) 7 (min.) 7 (min.)
Density (lb/ft
3
) ASTM D 6023 NA
(3)
NA NA
30 to 70 or as
specifed
(4)
Water absorption of aggregate
AASHTO T 85
20% max.
Compressive strength (psi) 125 (max.) 125 (max.) 800 (max.) 90 to 400
Note: 1. Quantities may vary in order to adapt mix to density and strength requirements or to adapt to site
conditions and material characteristics.
2. Requires use of suitable light-weight aggregate or air-entraining admixture.
3. Not applicable.
4. Approximate value; use of air-entraining agent may reduce this value.
(ADAPTED FROM PennDOT, 2007)
< PREVIOUS VIEW
6-138
Chapter 6
MAY 2009
in the following cases: (1) specifc data are simply not available and the only possibility is indirect
comparison to other properties, (2) a limited amount of data for the specifc property of interest are
available and the correlation will complement these limited data, or (3) the validity of certain data is
in question and a comparison to previous test results allows the accuracy of the data to be evaluated.
It is strongly emphasized that correlations should never be used as a substitute for an adequate site
exploration program, but rather should complement and verify available test data. Examples of each
of the three cases listed above are provided in the following:
Specifc data are unavailable Several examples of this type exist. Most notable is the
strength of uncemented clean sands. Undisturbed sampling may be problematic and
prohibitively expensive, and correlations to SPT, CPT, and other in-situ tests results
have been shown to be quite reliable.
Limited data are available If few high-quality consolidation tests are performed
and compression properties are found to correlate well with the results of Aterberg
limits tests, it may be concluded that Aterberg limits tests can be used for assess-
ment of compression properties.
Data validation If results from tests on two similar materials are inconsistent, com-
paring the results to previous data for similar soils may allow determination as to
whether the data are simply inconsistent or if some of the data are inaccurate.
There are several sources of correlation data for geotechnical materials and properties. Many geo-
technical textbooks and references provide correlations, including Holtz and Kovacs (1981), DOD
(2005) and Carter and Bentley (1991). The Electric Power Research Institute (EPRI) commissioned
the preparation of useful documents that include several correlations for laboratory and in-situ tests
(Kulhawy and Mayne, 1970). For information regarding the engineering properties of coal refuse
materials, designers should refer to Hegazy et al. (2004), Chen (1979), DiGioia and Gray (1979), and
Coates and Yu (1977).
The following comments on applicability and use of correlations are noted:
A correlation is only as good as the data upon which it is based.
There may be some scater in the correlation data. The efects of data scater should
be accounted for by using upper and lower bound (i.e., best case/worst case) scenar-
ios and factors of safety in the design calculations.
A correlation will be most accurate if it is calibrated to site soil/material conditions.
If calibration to site conditions and design-phase laboratory test data cannot be con-
ducted in sufcient detail, consideration should be given to an expanded program of
feld inspection, performance monitoring and maintenance (Chapter 12) and instru-
mentation (Chapter 13).
6.6 GEOTECHNICAL ANALYSES
Sections 6.1 through 6.5 have addressed planning and design considerations, site exploration, and
material property characterization required for development of a functional, safe and environmen-
tally acceptable coal refuse disposal facility. This section discusses geotechnical analyses associated
with refuse disposal facility design. Analytical procedures for seepage, setlement, slope stability,
rock excavation, and conduit design are presented. These discussions are necessarily limited in both
scope and depth. More extensive treatments of these subjects are available in references cited herein.
6.6.1 Analytical and Numerical Methods in Geotechnical Engineering
Many of the important areas of geotechnical engineering analysis are governed by diferential
equations for solving problems of elastic equilibrium, consolidation and steady-state seepage.
< PREVIOUS VIEW
6-139
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
Limit analysis using classical techniques such as Rankines earth pressure theory, Terzaghis bear-
ing capacity equation, or various methods of slices is another important approach for the solution
of geotechnical problems involving the failure of soil masses for slope stability analysis. These
traditional methods of analysis, as applied to the design of coal refuse disposal facilities, are
discussed in the following text. Seismic analysis and design issues for refuse disposal facilities are
addressed in Chapter 7.
The availability of powerful personal computers, augmented by improved tools for problem setup,
constitutive modeling of material properties, and post-processing of computational results, has facili-
tated the use of numerical methods to solve complex geotechnical problems. Sophisticated methods,
such as the fnite element (FE) method, are available to perform complex analyses related to seepage,
deformation, and slope stability. The FE method is a numerical method for obtaining solutions to dif-
ferential equations, given appropriate boundary condition data. Some of the most useful features of
the FE method are that it can:
Simulate one-, two- and three-dimensional problems
Accommodate complex geometries and construction sequencing
Model soil property variability (e.g., nonlinear stress-strain behavior and anisotropic
hydraulic conductivity)
Nearly all areas of geotechnical analysis can be solved using the FE method. However, the engineer
must recognize the inherent uncertainty and variability in material properties. The possible efects of
simplifying assumptions for material behavior in FE models must be carefully evaluated. An under-
standing of empirical relationships and the ability to apply the lessons of experience are key factors
in successfully using FE models and interpreting the results obtained.
6.6.2 Seepage Analysis
6.6.2.1 Basic Assumptions, Conditions Requiring Seepage Analysis
All earth and coal refuse structures are to some degree pervious to water and therefore suscep-
tible to seepage. Seepage is a concern in earthen dams and coal refuse embankments for three
reasons: (1) the pore-water pressures in an embankment and its foundation afect the stability of
the embankment, (2) an excessive hydraulic gradient on an embankment slope, at drain or fne/
coarse material interfaces, or at the toe can lead to piping, internal erosion, and destructive uplif
pressures, and (3) water lost through or under a structure may require treatment or may be a valu-
able water source worth recirculating through the preparation plant. Seepage control systems are
incorporated into embankments to minimize the potential for excessive or uncontrolled seepage.
Practical concepts for seepage control in coal refuse embankments and their foundations are dis-
cussed in Section 6.3. This section discusses analytical methods for assessing the need for seepage
control and for selecting materials and design dimensions to achieve that control. Additionally,
operational controls to mitigate seepage, such as the deposition of slurry within impoundments,
are also discussed.
Under steady-state conditions, seepage through an earth or coal refuse embankment and its founda-
tion can be estimated from the relationship for fow through porous media, which is based on the
following assumptions:
The fow occurs through incompressible media.
The fow is caused by gravity forces only.
The media through which fow occurs is always saturated.
The boundary conditions of the fow are known.
< PREVIOUS VIEW
6-140
Chapter 6
MAY 2009
For unsaturated fow and transient seepage analyses, additional assumptions may be required.
Whether or not the boundary conditions are actually known depends on the specifc fow condition
being analyzed. For fow through a pervious foundation under a relatively impervious embankment,
the boundary conditions are readily apparent. However, not all boundary conditions are easily deter-
mined, and many boundary conditions must be approximated or determined by iterative analysis
procedures (e.g., the seepage and phreatic condition in an embankment with internal drains).
Steady-state fow through porous media such as soils or coal refuse is normally laminar and therefore
follows the Darcy equation:
Q = k i a (6-15)
where:
Q = fow rate (volume/time)
k = coefcient of hydraulic conductivity (length/time)
i = hydraulic gradient (dimensionless)
a = cross-sectional area through which fow occurs (area)
Selection of the coefcient of hydraulic conductivity should be based on feld and laboratory testing,
although evaluation of other factors may also be required, including:
Published information (e.g., soil surveys) along with the results of site-specifc index
and classifcation tests for natural soils and coal refuse.
Potential variability of existing fll or mine spoil, especially when there are no avail-
able data related to in-situ material or placement.
Potential for anisotropic conditions (if not documented based upon site-specifc
data), particularly for coal refuse embankments where the horizontal hydraulic
conductivity may be an order of magnitude greater than the vertical hydraulic
conductivity.
Sensitivity of fow rate and hydraulic gradient (and phreatic surface) to variation of
assumed parameters.
Observed site groundwater levels in natural soils and existing embankments (to
assist in validating assumed values, particularly if multiple zones or anisotropic
conditions are present).
Seepage analysis for coal refuse embankments should be based on conservative estimates of hydrau-
lic conductivity, with values selected based on available test data (preferably feld tests), material
classifcations and anisotropy ratios resulting in conservative (higher) phreatic levels and seepage
rates. The ratio of horizontal to vertical hydraulic conductivity (anisotropy ratio) can vary due to
the material type, placement and compaction in lifs, and for embankment dams has been reported
to range from 1 to 100 (Fell et al., 2005). The U.S. Army Corps of Engineers (USACE, 1993) suggests
an anisotropy ratio between 2 and 10 or greater for embankment dams constructed in compacted
lifs, and MSHA (2007) recommends a ratio of at least 9 for coarse coal refuse or as based upon site-
specifc conditions and materials. For an existing embankment, back calculation of monitored pool
and piezometer levels to estimate anisotropy is recommended.
If fne coal refuse deposits are accounted for in the seepage analysis, an anisotropy ratio of between
1 and 10 is usually applied. In typical situations where coarse coal refuse is more permeable than
< PREVIOUS VIEW
6-141
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
the fne refuse, the anisotropy ratio of the fne refuse deposit may have litle efect on the computed
phreatic levels and seepage rates in downstream coarse refuse embankments. It is likely that the
anisotropy ratio of hydraulically-placed fne refuse is quite variable (beyond the range cited above)
such that design of seepage control and collection measures critical for stability of the facility should
be efective over the likely range of anisotropy.
For impounding coal refuse embankments, seepage analyses should be performed for critical inter-
mediate stages along with the fnal stage of development. The pool level used in the seepage analysis
should be the maximum normal pool level for the stage (usually based on the decant invert level) or,
if there is potential for saturation of the embankment due to retention of the design storm, the maxi-
mum design storm pool level.
Unsteady-state seepage analysis can be employed for evaluation of the extent of saturation of an
embankment or to assess the impact of rapid drawdown on the upstream embankment slope. Also,
unsaturated fow analysis is sometimes coupled with saturated fow analysis (generally using numeri-
cal models) to evaluate internal drainage systems, particularly where there is signifcant anisotropy.
Turbulent fow regimes may occur in rockfll and rock drains used in embankments. Based on ear-
lier work by Cedergren (1989), the U.S. Army Corps of Engineers (1993) provides guidance for esti-
mating the reduction in hydraulic conductivity of narrow size range aggregate caused by turbulent
fow at large hydraulic gradients in underdrains (Figure 6.47). As indicated in Cedergren (1989),
the reduction in hydraulic conductivity is of relatively litle importance for fat-lying underdrains
if hydraulic gradients are less than 0.02. Furthermore, the efect of increasing the hydraulic gra-
dient 100 times (from 0.01 to 1.0) reduces the hydraulic conductivity to one-tenth of the laminar
value. Alternatively, Leps (1973) developed the following empirical relationship for turbulent fow
through rock which is sometimes applied:
Q = a W m
0.5
i
0.54
[e / (1 + e)] (6-16)
where:
Q = fow rate (length
3
/time)
a = fow cross section area (length
2
)
W =
empirical constant for rockfll material, dependent on the shape and roughness
of rock particles and the viscosity of water (length-time units) (Wilkins, 1956)
m = hydraulic mean radius (length)
i = hydraulic gradient (dimensionless)
e = void ratio (dimensionless)
Leps (1973) provides the following guidance on the determination of Wm
0.5
based on rock size:
Rock size (in) 3/4 2 6 8 24 48
W m
0.5
(in/sec) 10 16 28 32 58 84
While Equation 6-16 applies to uniformly-sized rock, Leps suggests that it can be adapted for graded
materials by using the 50-percent rock size to compute Wm
0.5
, provided that the minus 1-inch material
is less than 30 percent, and preferably less than 10 percent by weight, of the rockfll. If the percentage
of minus 1-inch material is greater than 30, Equation 6-16 may not be applicable.
< PREVIOUS VIEW
6-142
Chapter 6
MAY 2009
6.6.2.2 Seepage Analysis Methods
There are several approaches for conducting seepage analyses. Prior to the widespread availabil-
ity of computers, graphical hand-solutions (i.e., hand-drawn fow nets) and hand calculations were
commonly used for modeling embankment and foundation seepage. However, the use of numeri-
cal methods (primarily fnite element modeling) has become the most common method for seepage
analyses. These approaches are discussed in the following sections.
Seepage in homogeneous coal refuse embankments can be readily analyzed using graphical and
analytical methods, provided that the hydraulic conductivity and any anisotropy are known. Zoned
embankments, variable foundation conditions, and the presence of the setled fne coal refuse intro-
duce additional complexity, and numerical methods are beter suited for detailed analysis of these
situations. Coal refuse embankments are developed in stages, each with a specifc confguration and
internal drainage system designed on the basis of the projected fne coal refuse level and pool level.
Thus, in addition to the fnal impounding stage, intermediate stages will likely need to be analyzed as
part of the design of internal drainage structures and to establish the stage confguration in conjunc-
tion with the stability analysis.
With either the graphical, analytical or numerical modeling approach, it is essential that the
boundary conditions for the analysis be accurately defned. These boundary conditions typically
include:
Appropriate reservoir/impoundment pool level
Foundation conditions, including possible artesian conditions
Location(s) and capacity(s) of internal drains
Geometry of intermediate and post-construction cross sections
Maximum and minimum potential infow rate(s) to the reservoir/impoundment
FIGURE 6.47 APPROXIMATION FOR ESTIMATED REDUCTION IN HYDRAULIC
CONDUCTIVITY UNDER TRUBULENT FLOW
0.025
EFFECTIVE SIZE (IN)
0.3
0.5
1.0
0.10
5.0
LIMITING
ENVELOPE
COMPLETE
TURBULENCE
FIGURE 6.47 APPROXIMATION FOR ESTIMATED REDUCTION IN HYDRAULIC
CONDUCTIVITY UNDER TURBULENT FLOW
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
HYDRAULIC GRADIENT, i
R
E
L
A
T
I
V
E
H
Y
D
R
A
U
L
I
C
C
O
N
D
U
C
T
I
V
I
T
Y
,
k
(CEDERGREN, 1989)
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
FIGURE 6.47 APPROXIMATION FOR ESTIMATED REDUCTION IN HYDRAULIC
CONDUCTIVITY UNDER TRUBULENT FLOW
< PREVIOUS VIEW
6-143
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
Selection of the appropriate elevations for water and fne coal refuse in an impoundment are impor-
tant if meaningful results are to be realized. Typically, normal pool or the maximum decant inlet
level for a specifc stage is selected. If the reservoir pool is above the setled fne coal refuse level and
against the upstream slope of the impoundment, the efect on the phreatic surface elevation of the
fne coal refuse will be limited, even if it has low hydraulic conductivity relative to the coarse refuse.
If the impoundment is designed to retain a design storm that would cause saturation of the embank-
ment due to an elevated reservoir level, then seepage associated with that reservoir level should be
analyzed. Cedergren (1989) provides a method for estimating the approximate time for embankment
saturation to occur for simple embankment geometry and homogeneous hydraulic conductivity.
Unsteady seepage analyses can be performed for complicated geometries and zoned embankments,
although it is recommended that the designer exercise caution and consider redundant seepage con-
trol measures in critical situations.
Site conditions may necessitate the evaluation of other features and boundary conditions. As one
of the primary purposes of seepage analyses is to determine seepage conditions for use in stability
analyses, the locations of the seepage cross sections should be consistent with the locations of stabil-
ity sections (for 2D analyses). Also, the size, lateral extent and continuity of drainage features should
be evaluated when deciding whether and how to incorporate them into the analysis. For example, a
blanket drain that is limited in lateral extent to the center of a valley may not signifcantly afect the
phreatic level at the abutments of an embankment. In such a case, it may be necessary to also analyze
the infuence of a higher saturation level in cross sections close to the abutments.
6.6.2.2.1 Graphical (Flow Net) Approach
A fow net is a set of orthogonal lines graphically representing the seepage conditions in an embank-
ment. The seepage at any cross section of an earth or coal refuse embankment can be determined by
constructing a fow net using free-hand trial-and-error sketching. One group of fow net lines repre-
sents paths of fowing water, and the other group represents lines of equal head (equipotential lines).
The fow net must satisfy the fow boundary conditions. The fow lines and equipotential lines must
intersect at right angles, and for each element the mean dimension parallel to the fow lines must
equal the mean dimension parallel to the equipotential lines (or remain in the same proportion over
the extent of the fow net). These dimensional requirements can be checked by drawing a circle in
elements and determining if it is tangent to the mid-sides of the adjacent fow and equipotential lines.
Figure 6.48 shows a typical fow net analysis.
To estimate the rate of seepage fow from a two-dimensional fow net of unit thickness, the Darcy
equation can be writen in the following form:
q = k h (N
f
/N
d
) (6-17)
where:
q = fow rate per unit width (length
2
/time)
k = coefcient of hydraulic conductivity (length/time)
h = total head loss across the system (length)
N
f
= number of fowpaths through the system (dimensionless)
N
d
= number of potential drop (divisions of head loss) across the system (dimensionless)
Examples of the application of this equation are provided in numerous references, including Cede-
gren (1989), Harr (1962) and USACE (1993).
< PREVIOUS VIEW
6-144 REV.
Chapter 6
AUG. 2010
An example of the fow net graphical approach is illustrated in Figure 6.49a for a homogeneous dam
section. Figure 6.49b illustrates the graphical transformations necessary to develop fow nets for a
homogeneous dam embankment with anisotropic hydraulic conductivity. When a dam is composed
of a number of diferent soils with varying anisotropic hydraulic conductivity, the graphical approach
becomes considerably more complicated. In practice, the use of graphical methods may require that a
number of simplifying assumptions be made to render the problem manageable. The construction of
fow nets and their application are described in USDA (1973a) and Reddi (2003).
6.6.2.2.2 Analytical Solutions
The seepage and associated phreatic surface for an embankment dam with an underdrain resting on
an impervious base, generally referred to as Kozenys basic parabola, can be determined analytically
(Harr, 1962). The phreatic surface and minimum width of drain a
0
(Figure 6.48) can be estimated
from the following equations:
x = ( ky
2
/2q) + (q/2k) y
0
= (d
2
+ h
2
)
0.5
d (6-18)
q = k [(d
2
+ h
2
)
0.5
d] (6-19)
where:
k = coefcient of hydraulic conductivity (length/time)
q = seepage rate per unit length of drain (length
2
/time)
Also, the d, h, x, and y dimensions (length) are as shown on Figure 6.48 and a
0
= y
0
/
2
.
Harr (1962) provides analytical solutions for embankments with foundations at great depth (as well
as fnite depth) and with variable upstream slopes. These procedures allow for treatment of anisot-
ropy using transformed sections (simple expansion or contraction of spatial coordinates to convert an
anisotropic zone to an isotropic zone).
FIGURE 6.48 FLOW NET ANALYSIS
F
Y
C
X
D
Z
y
0
E A B
a
0
IMPERMEABLE FOUNDATION
= 0
= 0
=
q
d
= -kh
h
(HARR, 1962)
FIGURE 6.48 FLOW NET ANALYSIS
FIGURE 6.48 FLOW NET ANALYSIS
< PREVIOUS VIEW
6-145
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
Kashef (1986) and Mishra and Singh (2005) point out simplifying assumptions for Kozenys basic
parabola and their impact on seepage rate and drain length. The results are compared with numerical
methods for zoned embankments.
Coal refuse embankments frequently employ limited width and height internal drains positioned
above the base of the dam to intercept seepage and control the phreatic surface. Van Zyl and Rob-
ertson (1980) have analyzed a drain of limited width on an impervious base. They estimate the
efective width of the drain as approximately 0.1h for the condition where d/h > 2, provided the
drain flter has sufcient capacity. They also recommend that the drain width be sized 50 percent
larger than the width determined based on the capacity of the flter, with a minimum dimension
of 2 meters.
Analytical solutions provide a means to estimate seepage and phreatic levels for some specifc bound-
ary conditions. They can be useful for initial development of fow nets and for interpreting results
obtained with numerical modeling, as subsequently discussed.
6.6.2.2.3 Numerical Modeling Approach
As noted previously, the advent of powerful computers has facilitated the use of numerical model-
ing for evaluating seepage. There are a number of fnite element programs that can be used to model
seepage for two- and three-dimensional, anisotropic, unconfned and confned fow. The use of such
computer programs has several advantages over the use of graphical methods to analyze seepage.
Frequently, computer programs support the use of hydraulic conductivity functions (relationships
between hydraulic conductivity and pore-water pressure) and thus allow the inclusion of unsatu-
rated fow into the seepage model. Computer programs can more easily accommodate anisotropic
FIGURE 6.49 FLOW NET ANALYSIS FOR ANISOTROPIC CONDITIONS
FIGURE 6.49 FLOW NET ANALYSIS FOR ANISOTROPIC CONDITIONS
PHREATIC SURFACE FOR kh = 16kv
(TRANSPOSED BACK FROM BELOW)
PHREATIC SURFACE FOR kh = kv
NOTE: FLOW NET IS FOR ISOTROPIC
CASE (kh = kv)
DRAIN k ~ 0
h
6.49a ACTUAL CROSS SECTION
(USDA, 1973a)
6.49b TRANSFORMED CROSS SECTION
DRAIN
h
k ~ 0
NOTE: HORIZONTAL TRANSFORMATION FACTOR
kv
kn
=
1
=
16
0.25 =
WHERE: kv = VERTICAL HYDRAULIC CONDUCTIVITY
kh = HORIZONTAL HYDRAULIC CONDUCTIVITY
FIGURE 6.49 FLOW NET ANALYSIS FOR ANISOTROPIC CONDITIONS
< PREVIOUS VIEW
6-146
Chapter 6
MAY 2009
hydraulic conductivities and complicated soil/rock geometries, including varying anisotropy ratios.
Transient analyses are also possible using numerical methods if water content data for each material
in the model are available.
Computer programs utilizing numerical methods frequently provide several alternatives for model-
ing boundary conditions, including the ability to model internal drainage systems of limited width
and height located above the base of the embankment. However, not all of these alternatives will
result in a realistic model. Therefore considerable judgment must be used if realistic results are
to be achieved. As noted previously, accurate geometry and realistic representation of boundary
conditions are vitally important when seting up a seepage model. The resulting model should be
checked using a more simple approach (e.g., fow nets) to determine whether the resulting fow
paths and related directional orientation of velocity vectors, predicted total head, and predicted
pressure heads are realistic.
The application of numerical methods for seepage analyses at fne coal refuse impoundments has
been presented in publications by Snow et al. (2000) and Thacker (2000). These studies present
models for steady-state fow conditions with anisotropic hydraulic conductivity for various mate-
rial property zones, and they use specifed heads in the embankment or as boundary conditions
to simulate impoundment levels or drains. For numerical models of this type, fow vectors and
hydraulic gradients should be checked to verify that values are reasonable and that mass balance
is maintained.
6.6.2.3 Seepage Control Measures
6.6.2.3.1 Granular Drain and Filter Requirements
Drains and flters made of aggregate materials are used within and under coal refuse embankments
to collect seepage and to lower pore-water pressures in parts of embankment where internal ero-
sion or unstable conditions might otherwise develop. The use of drains and flters in coal refuse
embankments is also discussed in Section 6.3. Technical requirements for drains are discussed in the
following pages.
For a drainage system to be efective, the following criteria must be satisfed:
The hydraulic conductivity must increase in the direction of fow.
The particles in the system must be stable against the seepage forces.
Material segregation during construction must be prevented so that
grain-size characteristics are uniform throughout the drain.
The materials must not be susceptible to clogging over time.
The materials must be durable and not decompose over time.
Filter criteria must be satisfed for transitions between a drain and adjacent portions of an embank-
ment and between zones of fner and coarser materials within the embankment. As water perco-
lates through soil or refuse materials, seepage forces in the direction of fow are exerted on the
soil/material particles. Where water fows from a fne soil into a coarse soil, it may transport fne
particles into the voids in the coarser material. This erosion of the fne material can lead to piping,
a phenomenon where a path of much higher hydraulic conductivity that could ultimately cause
a failure is developed within an embankment. When the diference in grain-size distribution for
adjacent embankment zones is too great to prevent piping, a flter or layer of material of intermedi-
ate particle size needs to be placed between the two zones. When adjacent embankment and drain-
age system zones difer greatly in particle size, two or more flter zones with graduated particle
sizes may be needed.
< PREVIOUS VIEW
6-147
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
The U.S. Army Corps of Engineers (USACE, 1993) guidance for selecting materials for successive
layers of a flter and drainage system, stated as rules, is provided in the following:
To prevent particles of a fne soil from washing into an adjacent coarser soil:
Rule 1: The criteria presented in Table 6.47 should be followed.
Rule 2: Ideally, grain-size curves for the base and flter soils should be approximately
parallel. For gap-graded base materials (sometimes the case with coarse coal
refuse), flter criteria that exclude the coarser portion of the base soil should be
applied. Also adjustments can be made to exclude particles larger than the No.
4 (4.75-mm) sieve. The flter should be designed to protect the fne matrix. For
gap-graded and broadly-graded materials, considerable care and judgment
must be employed in identifying the grain-size portion of the base soil to be
flter-protected so that sufcient capacity to transmit seepage fow from the
base soil is provided (USBR, 2007a). Alternatively, tests can be conducted in
the laboratory to select the appropriate flter (Sherard and Dunnigan, 1985).
For hydraulic conductivity consistent with adequate discharge capacity:
Rule 3:
D
15 flter soil
3 to 5 (6-20)
D
15 base soil
Rule 4: The portion passing the No. 200 sieve should be less than or equal to 5 percent,
and the fnes should be cohesionless.
To prevent segregation during placement:
Rule 5: The restrictions on flter grain size presented in Table 6.48 should be followed.
Rule 6: D
max
3 inches.
The USACE (1993) provides guidance for flters within gap-graded and broadly-graded soils. The
later condition may be particularly relevant because the grain-size distribution associated with
coarse coal refuse embankments is frequently either gap-graded or broadly-graded. Additionally,
the above guidelines (Rule 6) may require adjustment in some project-specifc situations where large
mine rock overburden is used in downstream zones.
Where a perforated drainpipe is the fnal element of a drainage system, a well-graded gravel that
will not wash through the perforations or slots should be placed around the pipe. The D
85
size of
the gravel should be greater than the width of the perforations or slots. Experience has shown that
drainpipes placed with perforations directed downward are less likely to clog. Geotextiles should
not be placed directly against a perforated or sloted pipe due to the potential for clogging. Instead, a
zone of sand or aggregate should be placed around the pipe and then the flter (geotextile or another
granular layer) should be placed around the aggregate (France, 2004)
The discharge capacity of all portions of a drainage system must be checked for adequacy, particu-
larly for thin horizontal drains. This can be accomplished by very simple approximate calculations
using the Darcy equation presented earlier in this section. Cedergren (1989) contains several useful
examples of this calculation.
The NRCS (1994) and the USBR (2007a) also provide guidance on flter criteria, including the use of
perforated pipes in drains. These criteria may be applied to coal refuse disposal facilities. McCook
< PREVIOUS VIEW
6-148
Chapter 6
MAY 2009
(2006) presents an overview of design and construction issues for pipes in drain systems, comparing
criteria published by the USACE, NRCS and USBR and methods for estimating infow capacity. Slot-
ted pipe generally is less prone to clogging and has a much larger infow capacity than circular per-
forated pipe. However, perforated pipe probably has higher strength than sloted or screened pipe,
especially when the perforations are relatively small in diameter and widely spaced. Care should be
exercised in selecting pipe materials, recognizing that: (1) sloted single-wall corrugated tubing has
been vulnerable to crushing in some drain installations and (2) single-wall corrugated pipe is seldom
used in the design of signifcant- and high-hazard-potential dams. Pipes in drains should generally
be designed to withstand the maximum height of backfll over the pipe, and they should be protected
against damage during construction. Care should be exercised in evaluating the load-carrying capac-
ity of perforated pipe, as the perforations can afect the strength of the pipe. Since the outer rings of
TABLE 6.48 RESTRICTIONS ON FILTER PARTICLE SIZE TO LIMIT SEGREGATION
Minimum D
10 flter
(mm)
Maximum D
90 flter
(mm)
< 0.5 20
0.5 to 1.0 25
1.0 to 2.0 30
2.0 to 5.0 40
5.0 to 10 50
10 to 50 60
(USACE, 1993)
TABLE 6.47 FILTER CRITERIA AS A FUNCTION OF BASE SOIL AND PERCENT
PASSING NO. 200 SIEVE
Base Soil
Category
Base Soil
Description
Percent Finer Than
No. 200 (0.075-mm)
Sieve
(1)
Filter Criteria
(2)
1 Fine silts and clays > 85
D
15
flter soil
9 D
85
base soil
D
15 flter soil
0.2 mm
2
Sands, silts, clays,
and silty to clayey
sands
40 to 85 D
15 flter soil
0.7 mm
3
Silty and clayey
sands and gravels
15 to 39
D
15
flter soil
0.7 mm + [4(D
85 base soil
) - 0.7mm]
x [(40 - A) / (40 - 15)]
(3,4)
4 Sands and Gravels < 15 D
15
flter soil
4 to 5 D
85
base soil
(5)
Note: 1. Category designation for soil containing particles larger than 4.75 mm is determined from a gradation
curve of the base soil that has been adjusted to 100 percent passing the No. 4 (4.75-mm) sieve.
2. Filters should have a maximum particle size of 3 inches (75 mm) and a maximum of 5 percent passing
the No. 200 (0.075-mm) sieve with a PI = 0. To provide suffcient hydraulic conductivity, flters must have
D
15 flter soil
4 D
15 base soil
, but not less than 0.1 mm.
3. A is the percent passing the No. 200 sieve using the adjusted grain-size curve.
4. When 4 D
85 base soil
< 0.7 mm, use D
15 flter soil
= 0.7 mm.
5. In Category 4, the D
15
flter soil
4 D
85
base soil
criterion should be used for flters beneath riprap
subject to wave action and for drains that may be subjected to violent surging and vibration.
(USACE, 1993)
< PREVIOUS VIEW
6-149
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
corrugated pipe carry the majority of the load, the efect of perforations on the inner rings is thought
to be negligible (< 1 percent). Double-wall, corrugated and sloted HDPE and PVC pipe is available
in a variety of diameters.
It should also be noted that all materials used in drainage systems for a coal refuse embankment
should be resistant to corrosion from exposure to seepage leachates. Guidelines for selecting such
materials are presented in Chapter 11.
Drain Capacity and Grade
Drains should be sized with sufcient dimensions and grade (typical minimum slope of 1 percent)
to convey the estimated drain demands based on the seepage analyses multiplied by a safety factor.
It is recommended that drains have the capacity to convey 10 times the estimated drain demand
(Cedergren, 1989). It is also recommended that drains installed in trenches should include a mini-
mum width determined on the basis of capacity requirements and constructability considerations
(typically 3 feet) of the drain core material surrounded by specifed minimum thicknesses for the
surrounding flter aggregate(s). For drains at the ground surface or the coal refuse working surface,
the minimum width of the core material should be determined on the basis of constructability. Addi-
tional discussion of drain and flter dimensioning is available from the USBR (2007a).
Detailed review of the hydraulic gradients in the vicinity of a drain may lead to adjustments in drain
dimensions. McCook (2002) provides a discussion of critical hydraulic gradients and notes that cohe-
sive soils such as clays will sustain signifcant gradients when confned. Indraratna and Radampola
(2002) provide further guidance relative to critical gradients at drain flters.
Drain Confgurations
The above criteria for drain capacity and grade apply to aggregate drains of various confgurations,
including trench drains, fnger drains, blanket drains, chimney drains and discrete longitudinal
drains. In determining the thickness of drainage layers, the layer inclination and the method of place-
ment in addition to hydraulic conductivity requirements must be considered. Drainage layers placed
by machine should generally have a thickness of at least 18 inches. Layers placed very carefully by
hand in confned spaces or machine-placed layers for which there is construction oversight and QC
compliance monitoring should have a minimum thickness of 12 inches. If adjacent materials could
migrate into drainage layers during placement, the thicknesses should be increased. Where machine
placement is used for constructing steeply inclined drainage layers within an embankment, each layer
should be sufciently thick to permit efcient operation of the construction equipment. Filters and
drainage structures that are so thin that a small amount of contamination during construction would
reduce the efective size to below design requirements are generally considered to be inadequate.
Another factor that may impact the thickness of drainage layers is the source and grain size of the gran-
ular material. Granular drains should be compatible with the selected flter material. The minimum
thickness of a well-graded rock drain must be more than the maximum size of the rock, and gradation
requirements should be strictly monitored so that there is no concentration of large rock. If uniformly-
graded rock is used for a drain, then an increase in the minimum thickness should be considered. A
minimum thickness equal to twice the maximum rock size should provide predictable fow capacity.
6.6.2.3.2 Geotextile-Wrapped Drains
At some coal refuse embankments, geosynthetic materials have been used to separate granular drain
material from soils or coal refuse as a replacement for or supplement to granular flters. In addition
to some state dam safety agencies, the Nation Dam Safety Review Board currently recommends
that geotextiles not be used as flters in locations where they would be critical to the safety of an
embankment dam, citing concerns about the long-term capability of the geotextile to function with-
< PREVIOUS VIEW
6-150
Chapter 6
MAY 2009
out deterioration or clogging. Critical drain applications and areas of concern include internal drains
for controlling the phreatic surface for embankment stability and drains and flters that are designed
to minimize the potential for piping of susceptible soils. FEMA, in a publication planned for 2009, is
addressing the use of geotextiles in embankment dams.
MSHA has generally permited the use of geotextiles as flters in slurry impoundments provided that
testing using site-specifc materials demonstrates acceptable behavior with respect to clogging and there
is sufcient instrumentation to monitor the phreatic level. In addition to monitoring of the phreatic level,
the seepage quantity from geotextile-wrapped drains should be monitored as an additional indicator
of how the geotextile is performing. In recognition of the potentially more critical seepage conditions
that can exist in other dams with signifcant hydraulic head and narrow cross section (e.g., fresh water
dams), and until the use of geotextiles is accepted in this application, MSHA advocates that granular
flters be used in such signifcant- and high-hazard-potential dams where flters are critical to safety.
As discussed in Section 6.5.5, a geosynthetic is a planar polymeric material used with soil or rock
as an integral part of a civil engineering project, structure, or system. Geotextiles are a subcategory
of geosynthetics and are made from woven or nonwoven fabric that allows the passage of water. At
refuse disposal facilities, geotextiles have ofen been used as flters in internal drains. The geotextiles
restrict movement of soil particles as water fows into the drain. Typically, non-woven geotextiles are
used for fltration purposes. However, woven monoflament geotextiles have performed well in flter
applications, and knit geotextiles have been used around perforated pipes as part of a two-stage flter
in combination with a primary sand flter layer.
AASHTO M288, Standard Specifcation for Geotextile Specifcation for Highway Applications,
(AASHTO, 2008) provides reference information concerning material properties, applications related
to highway use (including subsurface drainage) and construction guidance for the use of geotextiles
in drainage applications. FHWA Publication No. HI-95-038, Geosynthetic Design and Construction
Guidelines, (Holtz et al., 1998) is a comprehensive reference providing information on the retention
criterion, hydraulic conductivity criterion, clogging resistance criterion and survivability criterion, as
discussed in subsequent paragraphs.
Geotextiles, like graded granular flters, require engineering design if they are to perform as desired.
Unless fow requirements, piping and clogging resistance, and constructability requirements are
accurately specifed, the geotextile/soil fltration system may not perform properly. Also, as with
graded granular flters, construction using geotextiles must be monitored to verify that installation is
performed correctly and that the geotextiles are not damaged during installation.
Design of geotextile flters is comparable to design of graded granular flters. A geotextile is similar
to a soil in that it has openings (voids) and flaments and fbers (particles). However, the geometric
relationship between flaments and openings is more complex than the relationship between par-
ticles and voids. In geotextiles, opening size can be measured directly, whereas for soils pore size is a
function of particle size. Because pore size can be directly measured, relatively simple relationships
have been developed between the pore sizes and particle sizes of the soil to be retained. Three fltra-
tion concepts are used in the design process:
1. If the size of the largest opening in the geotextile flter is smaller than the larger par-
ticles of embankment material, the flter will retain the embankment material.
2. If the smaller openings in the geotextile are sufciently large enough to allow smaller
particles of embankment material to pass through, then the geotextile will not clog.
3. The number of openings in the geotextile should be such that adequate fow can be
maintained even if some of the openings become clogged.
< PREVIOUS VIEW
6-151
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
REV. AUG. 2010
The above simple concepts and analogies to soil flter design criteria have been used to establish
design criteria for geotextiles (Holtz et al., 1998). Specifcally, these criteria require that geotex-
tiles must:
Be capable of retaining soil or other embankment material (retention criterion)
Allow water to pass (hydraulic conductivity criterion)
Be functional throughout the life of the structure (clogging resistance criterion)
Survive the installation process (survivability criterion)
Retention Criterion
For steady-state fow conditions:
AOS BD
85 soil
(6-21)
AOS O
95
(6-22)
where:
AOS = apparent opening size (length)
O
95
= opening size in the geotextile for which 95 percent are smaller (length)
B = coefcient (dimensionless)
D
85
= soil particle size for which 85 percent are smaller (length)
B ranges from 0.5 to 2 and is a function of the type of soil to be fltered, its density, the uniformity coef-
fcient C
u
(for granular soils), the type of geotextile (woven or nonwoven), and the fow conditions.
For sands, gravelly sands, silty sands, and clayey sands (with < 50 percent passing the No. 200 sieve),
B is a function of C
u
as defned below:
C
u
2 or 8: B = 1
2 C
u
4: B = 0.5 C
u
4 < C
u
< 8: B = 8 / C
u
where:
C
u
= D
60
/D
10
(6-23)
For silts and clays, B is a function of the type of geotextile:
Woven B = 1 O
95
D
85
Nonwoven B = 1.8 O
95
D
85
For both AOS or O
95
0.3 mm
Due to their random pore characteristics and, for some types, their felt-like nature, nonwoven
geotextiles will generally retain fner particles than a woven geotextile of the same apparent open-
ing size. Therefore, a value of B = 1 is more conservative for nonwoven geotextiles than it is for
woven geotextiles.
< PREVIOUS VIEW
6-152
Chapter 6
MAY 2009
Hydraulic Conductivity Criterion
For non-critical applications (less severe conditions):
k
geotextile
k
retained material
(6-24)
For critical applications (severe conditions):
k
geotextile
10 k
retained material
(6-25)
Guidelines for determining critical nature or severity for drainage applications are provided in
Table 6.49. Geotextile permitivity can be defned as:
= k/ t (6-26)
where:
k = Darcy coefcient of hydraulic conductivity (length/time)
t = geotextile fabric thickness (length)
Geotextile permitivity should meet the following requirements:
0.5/sec for < 15 percent passing No. 200 sieve
0.2/sec for 15 to 50 percent passing No. 200 sieve
0.1/sec for > 50 percent passing No. 200 sieve
Clogging Resistance Criterion
For actual fow capacity, the hydraulic conductivity criterion for non-critical applications is conserva-
tive because an equal quantity of fow through a relatively thin geotextile takes signifcantly less time
than fow through a thick granular flter. Where fow reduction is judged not to be a problem, such as
in clean, medium to coarse sands and gravels, Equation 6-24 may be used. Even so, some pores in the
geotextile may become blocked or plugged with time. Therefore, for critical or severe applications,
Equation 6-25 should be used to provide an additional level of conservatism. FEMA, in a document
to be published in 2009, suggests even greater conservatism with k
geotextile
= 10 to 100 k
retained material
and
indicates that the French (Degoute and Fry, 2002) use 100 k
retained material
for dams.
The required fow rate through the system q should also be determined, and the geotextile and drain-
age aggregate should be selected to provide adequate capacity. As indicated previously, fow capac-
ities should not be a problem for most applications. In some situations, however, such as where
geotextiles are used in multiple stage flters around a perforated or sloted pipe (pipe wraps), portions
of the geotextile may not function efectively. For these applications, the following criteria should be
used together with the hydraulic conductivity criteria:
q
required
= q
geotextile
(A
g
/A
t
) (6-27)
where:
A
g
= geotextile area available for fow (length
2
)
A
t
= total geotextile area (length
2
)
< PREVIOUS VIEW
6-153
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
TABLE 6.49 GUIDELINES FOR EVALUATING CRITICAL NATURE OR SEVERITY
FOR DRAINAGE APPLICATIONS
Category Critical Less Critical
A. Critical Nature of the Project
1. Risk of loss of life and/or structural damage due to
drain failure
High None
2. Repair costs versus installation costs of drain
Much Higher Equal or Less
3. Evidence of drain clogging before potential catastrophic
failure
None Yes
B. Severity of the Conditions
1. Soil to be drained
Gap-graded, pipable or
dispersible
Well-graded or uniform
2. Hydraulic gradient High Low
3. Flow conditions
Dynamic, cyclic or
pulsating
Steady-state
(ADAPTED FROM CARROLL, 1983)
Clogging resistance for less critical/less severe conditions should be designed to meet:
O
95 geotextile
3 D
15 retained material
(6-28)
This relationship applies to soils with C
u
> 3. For C
u
< 3, a geotextile with the maximum AOS value
should be selected. In situations where clogging is possible (e.g., gap-graded or silty soils), the fol-
lowing optional qualifers may be applied:
For nonwovens porosity of the geotextile n 50 percent
For woven monoflament and slit-flm wovens percent open area (POA) 4 percent
Most common non-woven geotextiles have porosities much greater than 70 percent. While most
woven monoflaments easily meet the criterion of POA 4 percent, tightly woven slit flms do not
and are therefore not recommended for subsurface drainage applications.
For critical/severe conditions, geotextiles should be selected in accordance with the retention and
hydraulic conductivity criteria described previously. A fltration test should be conducted to check
for clogging using samples of on-site materials and hydraulic conditions such as long-term fow tests
or the gradient ratio test, which is described in ASTM D 5101, Standard Test Method for Measur-
ing the Soil-Geotextile System Clogging Potential by the Gradient Ratio. Dam safety agencies have
expressed concern about the use of geotextiles in critical flters and drains (e.g., structures critical for
the control of internal erosion and piping failures) due to susceptibility to:
Excessive clogging from buildup of fnes at the face of the geotextile
or from biological, fungal or mineral mater buildup
Separation at interfaces, junctions, connections and boundaries
Undetected damage during construction.
< PREVIOUS VIEW
6-154
Chapter 6
MAY 2009
These performance problems are related to the following mechanisms: (1) inability to support the
seepage discharge face, (2) excessive clogging or piping, (3) stress-induced distortion, (4) environ-
mental degradation, (5) slope instability, and (6) rupture. Designers should consider these concerns
and performance issues when using geotextiles and, as part of the planning process, should discuss
the acceptability of any proposed application with MSHA and state agencies.
Talbot et al. (2000) describe concerns about the use of geotextiles in dams, including the propensity
for excessive clogging because seepage forces move the geotextile away from the base soil and into
the voids of the adjacent drain. Thus, separation can occur between the base soil and geotextile allow-
ing fne particles from the base soil to accumulate at the interface, thus blinding the flter. Blinding
can also occur if there are open voids in the base soil or if the base soil surface is irregular and has
poor contact with the geotextile. These concerns can be addressed through: (1) use of fne gravel
(about -inch to 1-inch maximum size) for the drainage layer to improve contact with the base soil
(Giroud, 1997; van Zyl and Robertson, 1980), (2) grading the base soil surface smooth and placing the
geotextile in contact with the base soil with a minimum of wrinkles, and (3) minimizing vertical or
steeply inclined slopes (FEMA, to be published in 2009).
The potential for particulate clogging can be addressed through application of flter criteria, and
chemical and biological clogging can be addressed based on evaluation of the drain environment
and/or testing, as discussed subsequently. Stress-induced distortion, environmental degradation,
and stability can be addressed with design and laboratory testing procedures.
Chemical clogging of geotextiles at coal refuse disposal sites will most likely be associated with iron
and manganese oxidation and deposition, although it is possible that calcium carbonate precipita-
tion may also be encountered. These processes are afected by chemical reactions and biological
activity that varies depending on whether the environment is anaerobic and aerobic. Factors that
may lead to problems with clogging of drains are discussed by Smith and Hosler (2006). Research
into predictive methods has generally concentrated on biological clogging and the role of micro-
bial activity, and more work is necessary for prediction of potential chemical clogging. Long-term
hydraulic conductivity testing procedures can be used to evaluate the potential efect of chemical
clogging.
Koerner and Koerner (2005) provide drainage reduction factors for determination of the design fow
rate or transmissivity of geotextiles. These reduction factors address soil clogging and blinding,
reduction in void space, and chemical and biological clogging. Koerner and Koerner recommend
reduction factors to the design fow capacity of a geotextile of between 1.2 and 1.5 for landfll flters
to account for chemical clogging, and they suggest adoption of a high reduction factor where total
suspended solids in the permeating liquid is greater than 5,000 mg/l. Similarly, they recommend
reduction factors of between 2 and 5 to the design fow capacity of a geotextile to account for bio-
logical clogging and suggest adoption of a high reduction factor where biochemical oxygen demand
(BOD) is greater than 5,000 mg/l.
The potential for biological clogging can be examined in accordance with ASTM D1987, Standard
Test Method for Biological Clogging of Geotextile or Soil/Geotextile Filters. However, before using
this test, Mackey and Koerner (1999) recommend that to facilitate interpretation of the results of the
test the physical, chemical and biological processes at the site be evaluated and understood. If clog-
ging is a concern, a higher-porosity geotextile can be used, and/or the drain design and operation can
include an inspection and maintenance program for fushing the drainage system. For nonwoven
geotextiles, Luetich et al. (1992) recommend using the largest porosity available, but not less than
30 percent; for woven geotextiles they recommend using the largest POA, but not less than 4 per-
cent. Because of concerns for clogging, geotextiles with the largest opening sizes that satisfy piping
requirement should generally be used.
< PREVIOUS VIEW
6-155
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
Survivability Criterion
For fltration and drainage applications, if a geotextile is to survive the construction process, cer-
tain minimum strength and endurance properties are required. Table 6.50 provides these minimum
requirements for less critical/less severe applications. It is important to understand that these mini-
mum survivability parameters are based on empirical data from geotextiles that have performed sat-
isfactorily in drainage applications. These parameters serve as guidelines for selecting geotextiles for
most projects. The guidelines are not intended to replace site-specifc evaluation, testing, and design.
Geotextile material should be covered subsequent to installation as soon as possible to prevent degrada-
tion from sunlight or in accordance with the manufacturers recommendations. Geotextile endurance
TABLE 6.50 GEOTEXTILE STRENGTH PROPERTY REQUIREMENTS
FOR DRAINAGE GEOTEXTILES
(1,2,3,4)
Property
ASTM Test
Method
Units
Geotextile Class 2
(5)
Elongation
< 50%
(6)
50%
(6)
Grab Strength D 4632 N (lb) 1100 (247) 700 (157)
Sewn Seam Strength
(7)
D 4632 N (lb) 990 (223) 630 (142)
Tear Strength D 4533 N (lb) 400
(8)
(90) 250 (56)
Puncture Strength D 4833 N (lb) 400 (90) 250 (56)
Burst Strength D 3786 kPa (lb/in
2
) 2700 (392) 1300 (189)
Note: 1. Acceptance of geotextile material shall be based on ASTM D 4759, Standard Practice for
Determining the Specifcation Conformance of Geosynthetics.
2. Acceptance shall be based upon testing of either conformance sampler obtained using
Procedure A of ASTM D 4354, Standard Practice for Sampling of Geosynthetics for Testing,
or on manufacturers certifcations and testing of quality assurance samples obtained using
Procedure B of ASTM D 4354.
3. Values apply to minimum strength; use value in weaker principal direction. All numerical values
represent minimum average roll value (i.e., test results from any sampled roll in a lot shall meet
or exceed the minimum values in the table). Lot samples according to ASTM D 4354.
4. Woven slit flm geotextiles will not be allowed.
5. AASHTO Geotextile Class 2 is the default geotextile selection. The engineer may specify
AASHTO Class 3 geotextiles for trench drain applications based on one or more of the following:
a) The engineer has found Class 3 geotextiles to have suffcient survivability based on feld
experience.
b) The engineer has found Class 3 geotextiles to have suffcient survivability based on
laboratory testing and visual inspection of a geotextile sample removed from a feld test
section constructed under anticipated feld conditions.
c) Subsurface drain depth is less than 2 m, drain aggregate diameter is less than 30 mm,
and the compaction requirement is 95 percent of ASTM D 698, Standard Test Methods for
Laboratory Compaction Characteristics of Soil Using Standard Effort (12,400 ft-lbf/ft
3
(600
kN-mm
3
)).
6. As measured in accordance with ASTM D 4632, Standard Test Method for Grab Breaking Load
and Elongation of Geotextiles.
7. When seams are required. Values apply to both feld and manufactured seams.
8. The required MARV tear strength for woven monoflament geotextiles is 250 N (56 lb).
(ADAPTED FROM AASHTO STANDARD SPECIFICATIONS FOR TRANSPORTATION MATERIALS AND METHODS OF TESTING, PART I SPECIFICATIONS, 2007, BY PERMISSION OF THE
AMERICAN ASSOCIATION OF STATE HIGHWAY AND TRANSPORTATION OFFICIALS, WASHINGTON, D.C. USED BY PERMISSION. DOCUMENT MAY BE PURCHASED FROM THE AASHTO
BOOKSTORE AT 1-800-231-3475 OR ONLINE AT HTTP://BOOKSTORE.TRANSPORTATION.ORG)
< PREVIOUS VIEW
6-156
Chapter 6
MAY 2009
is related to longevity. Geotextiles have been shown to be basically inert materials in most environ-
ments and applications. However, some applications may expose the geotextile to chemical or bio-
logical activity that could dramatically afect fltration properties or durability. For example, in drains,
granular flters and geotextiles can become chemically clogged by iron or carbonate precipitates and
biologically clogged by algae and mosses. Biological clogging is a potential problem when flters and
drains are periodically inundated then exposed to air. Excessive chemical and biological clogging can
signifcantly afect flter and drain performance, and monitoring with piezometers is recommended.
6.6.2.3.3 Seepage Control along Conduits
Seepage along conduits that pass through coal refuse embankments should be controlled. Two meth-
ods for seepage control are flter diaphragms and anti-seep collars (also known as cutof collars).
Since the mid-1980s, the use of anti-seep collars has become less common and flter diaphragms have
become more widely used (Van Aller, 1998). Many dam safety agencies require the use of flter dia-
phragms because of the many instances where seepage problems have occurred along conduits even
when anti-seep collars were used. The primary advantage of flter diaphragms is the relative ease of
construction as compared to anti-seep collars, particularly with respect to compaction around con-
duits. The presence of anti-seep collars complicates compaction around conduits, and accordingly
they are more likely to function poorly due to construction faws. Also, flter diaphragms are con-
sidered to be a beter measure for mitigating the consequences of embankment cracking associated
with the presence of a conduit. Filter diaphragms should be constructed with suitable flter materials,
and careful placement is required during construction. The following sections present basic design
considerations for both flter diaphragms and anti-seep collars.
Filter Diaphragms
A flter diaphragm is used for intercepting seepage through backfll pores or cracks and to pre-
vent internal erosion of the backfll materials along buried conduit installations. To meet fltration
and drainage requirements, flter diaphragms may consist of a single zone or multiple zones of
granular material. The guidance for dimensioning flter diaphragms provided in the following text
is taken from the USDA (1985) and NRCS (2007b). Van Aller (1998) discusses various aspects of
flter diaphragm design, and McCook (2002) discusses site-specifc conditions that should also be
considered.
Filter diaphragms should be located approximately parallel to the centerline of the embankment and
approximately perpendicular to the direction of seepage fow, and should extend horizontally and
vertically into adjoining portions of the embankment and foundation. In homogeneous dams, flter
diaphragms should be located using the following criteria:
Downstream of the cutof trench
Downstream of the centerline of the dam if there is no cutof trench
Upstream of a point where the embankment cover (from the upstream face of the
diaphragm to the downstream face of the embankment) is at least one-half of the
diference between the elevation of the top of the flter diaphragm and the maximum
potential impoundment water level
In zoned embankments, the flter diaphragm should be located downstream of the core zone and/or
cutof trench, in accordance with the minimum cover guidance for homogeneous dams. In instances
where the downstream shell of a zoned embankment is more pervious than the diaphragm material,
the diaphragm should be located at the downstream face of the core zone.
Provisions should be made for discharging seepage and groundwater collected by the flter diaphragm.
These provisions could include tying the diaphragm into other drainage systems in the foundation,
< PREVIOUS VIEW
6-157
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
tying into internal embankment drains or designing/constructing a separate outlet for the flter dia-
phragm. Such an outlet should be designed with the assumption that the hydraulic conductivity in
the zone upstream of the flter diaphragm is 100 times the hydraulic conductivity in the compacted
embankment material. This zone should have a cross-sectional area equal to the area of the flter dia-
phragm, and the length of the seepage path should equal the distance from the embankment upstream
toe to the flter diaphragm. This higher hydraulic conductivity is intended to account for partially-
flled cracks and openings in the upstream zone. An advantage of having a separate outlet for the flter
diaphragm is that the seepage outfow can be monitored separately from other seepage fows and will
provide feedback as to the performance of the diaphragm system over time.
For rigid conduits, the NRCS (2007b) recommends that flter diaphragms should extend the follow-
ing minimum distances from the conduit surface (Figure 6.50):
1. Horizontally and vertically upward 3 times the outside diameter of circular conduits
or the vertical dimension of rectangular box conduits except that:
a. The vertical extension need be no higher than the maximum potential
impoundment level.
b. The horizontal extension need be no further than 5 feet from the sides
and slopes of any excavation for installation of the conduit.
2. Vertically downward from the conduit:
a. Filter diaphragms should extend from the pipe support 1.5 times the out-
side diameter of circular conduits or outside vertical dimension of box
conduits or to the top of rock, whichever is shallower.
b. Alternatively, for conduit setlement ratios of 0.7 and greater, flter
diaphragms should extend the greater of 1 foot beyond the botom of
the trench excavation for the conduit or 2 feet. The diaphragm should
be terminated at the surface of bedrock when it occurs within this dis-
tance. Additional control of general seepage through an upper zone of
weathered rock may be needed. The conduit setlement ratio is defned
in NRCS (2007a) and Technical Release-5 (TR-5) by the USDA (1958) and
requires a complex computation. On frm foundations with = 0.7 or
greater (the conduit setlement ratio for pipe on rock is 1.0), the flter dia-
phragm should extend below the pipe to rock, or at least 2 feet.
For fexible conduits, NRCS recommends that flter diaphragms be designed to extend in all direc-
tions a minimum of 2 times the outside diameter from the surface of the conduit, except that the dia-
phragm need not extend beyond the limits in 1a and 1b described above or beyond a bedrock surface
beneath the conduit.
Filter diaphragms should have a minimum thickness of 3 feet, but the thickness specifed should
be appropriate for the level of quality control and supervision during construction. If a multi-zone
system is employed to satisfy flter criteria, a minimum thickness of 1 foot should be used for any
single zone. Greater thickness may be required as dictated by: (1) fow capacity requirements, (2)
the need to tie the flter diaphragm into embankment internal or foundation drains, or (3) the need
to accommodate construction methods. An example design for a flter diaphragm associated with a
decant pipe is shown in Figure 6.51.
Some state regulatory agencies have developed specifc guidance on flter diaphragms. Also, McCook
(2002) discusses site-specifc conditions that may warrant enlarging the flter diaphragm relative to
< PREVIOUS VIEW
6-158
Chapter 6
MAY 2009
the minimum guidance cited above. These conditions include foundations with varying rock sur-
faces, sof soils, and situations where there is potential for diferential setlement and related strain.
Anti-seep Collars
Cutof or anti-seep collars are intended to minimize seepage along the contact between the outside
surface of a conduit and an embankment. Van Aller (2004) and FEMA (2005a) provide compelling
reasons to use flter diaphragms rather than anti-seep collars. For low-hazard-potential structures
with dam heights of 35 feet or less and storage volume less than 3,000 acre-feet, the NRCS (2002)
allows consideration of anti-seep collars. In coal refuse embankment dams, there may be site-specifc
reasons to use anti-seep collars. For example, during early stages of construction, installation of anti-
seep collars may be preferable to an internal drainage structure that, in the case of downstream con-
struction, could ultimately be located relatively far upstream and perhaps beneath the impoundment
during later stages of construction. In such situations, the pipe and collar backfll should be designed
so as to minimize the potential for concentrated seepage zones and internal erosion.
Cutof collars have been fabricated using concrete, steel, and plastic for consistency with conduit
materials. The intent of their use is to increase the length of percolation along the conduit contact
FIGURE 6.50 FILTER DIAPHRAGM DESIGN FOR CONDUIT
FIGURE 6.50 FILTER DIAPHRAGM DESIGN FOR CONDUIT
PROFILE THROUGH COAL REFUSE EMBANKMENT
COMPACTED COARSE
COAL REFUSE
DROP INLET
PRINCIPAL SPILLWAY
(PIPE OUTLET)
A
A
FILTER DIAPHRAGM
FILTER DIAPHRAGM
SECTION A - A
FLEXIBLE CONDUIT
2D
2D 2D
STOP AT ROCK SURFACE IF LESS THAN
2D BENEATH STRUCTURE
2D
3D
D
1.5D
3D
STOP AT ROCK SURFACE IF LESS THAN
1.5D BENEATH STRUCTURE
SECTION A - A
RIGID CONDUIT
FILTER DIAPHRAGM
(USDA, 1985)
FIGURE 6.50 FILTER DIAPHRAGM DESIGN FOR CONDUIT
< PREVIOUS VIEW
6-159
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
FIGURE 6.51 EXAMPLE DESIGN CONFIGURATION FOR FILTER DIAPHRAGM AND DECANT PIPE
FIGURE 6.51 EXAMPLE DESIGN CONFIGURATION FOR
FILTER DIAPHRAGM AND DECANT PIPE
SECTION B - B
FLOWABLE
BACKFILL
DECANT PIPE
PIPE SUPPORT
ON PERIODIC
SPACING
2.5 D (MAX)
4" SCHEDULE 80
DIAPHRAGM OUTLET PIPE
3"
C DECANT PIPE
D
L
PROTECTIVE COVER
AASHTO NO. 8
COARSE AGGREGATE
C DECANT PIPE
L
PROJECTED LIMITS OF TRENCH
EXCAVATION FOR PIPE UPSTREAM
AND DOWNSTREAM OF FILTER
DIAPHRAGM
GEOTEXTILE
ANGLE OF
REPOSE
SECTION A - A
2D OR 5' (MIN)
AASHTO NO. 57
COARSE AGGREGATE
2D (MIN)
SAND
2D
(OR TO ROCK)
PROTECTIVE COVER
FLOWABLE BACKFILL
A
A
SAND
4" O PERFORATED SCHEDULE 80 PVC PIPE
GEOTEXTILE
AASHTO NO. 8
COARSE AGGREGATE
ANGLE OF
REPOSE
COMPACTED
COAL
REFUSE
PROTECTIVE COVER
AASHTO NO. 57
COARSE AGGREGATE
........................................................................................................................
........................................................................................................................
CAP
B
B
FIGURE 6.51 EXAMPLE DESIGN CONFIGURATION FOR FILTER DIAPHRAGM AND DECANT PIPE
< PREVIOUS VIEW
6-160
Chapter 6
MAY 2009
surface by 20 to 30 percent for signifcant- and high-hazard-potential dams (USBR, 1987a). For a
conduit on an earth foundation, the cutof collar should completely encircle the conduit. Where
the foundation is sound rock and good contact along the base is expected, cutof collars will need
to extend only to a depth sufcient for keying into the rock foundation. Cutof collars should be
separated from rigid conduits using watertight fllers (gasket or seal) to avoid introducing concen-
trated stresses into the walls of the conduit. Non-rigid collars can be atached or clamped to fexible
conduits using gaskets that accommodate conduit deformation such that a watertight connection
is achieved.
For small, low-hazard-potential dams, the NRCS (2002) recommends the use of flter diaphragms
unless it is determined that anti-seep collars will adequately serve the purpose. If anti-seep collars are
used, the NRCS recommends increasing the fow path along the conduit by at least 15 percent, with
10- to 25-foot spacing between collars.
6.6.2.3.4 Relief Wells
If a pervious layer underlies a relatively impervious layer beneath the toe of an embankment slope,
pore-water pressures can build up in the pervious layer to produce an artesian efect if drainage
downstream is impeded and the layer is recharged upstream by groundwater, seepage from the
impoundment, or rainfall. If artesian water pressure must be reduced for stability, relief wells can be
used, as shown in Figure 6.52.
Water collected by relief wells is usually conducted through a horizontal overfow pipe at the ground
surface and discharged into a lined drainage ditch near the toe of the embankment. With this collec-
tion method, the phreatic surface can be lowered to the ground surface. If a greater lowering of head
is required, the outfalls from the wells and the collection drain can be lowered below the ground sur-
face provided that they can be connected to the collection drain through a discharge pipe. To allow
for inspection and maintenance, relief well casings should extend to the ground surface.
The spacing required for relief wells depends on the geology and groundwater hydrology at the site. As
with embankment seepage modeling, an analysis should be performed to evaluate the impact of relief
wells on foundation pressures. Numerical modeling is well suited to such an analysis, but requires
an understanding of groundwater conditions, including geometry, pre-construction piezometric levels
and the connection of the pervious strata with recharge features such as the impoundment or local
aquifers. In instances where the model can be sufciently simplifed, fow nets can be used successfully.
As a check, an estimate of the required spacing of relief wells can be made using methods presented by
Leonards (1962). Given the limitations inherent in this type of modeling, the adequate function of the
relief wells should be monitored with piezometers. A relief well system can be readily expanded if the
initial confguration fails to produce the desired reduction in piezometric head.
Atention should be given to relief well design details so that the wells meet performance require-
ments and can be properly inspected and maintained. The chemical content of the water to be recov-
ered should be evaluated to determine if any special precautions are needed for preventing corrosion
of any part of the relief well system. The diameter of the internal perforated pipe depends on the
anticipated fow volume, but should be no less than 6 inches. To facilitate infow and to prevent clog-
ging, relief wells should be surrounded by a flter designed according to the requirements discussed
earlier in this section.
The annulus around the portion of the relief well above the pervious layer should be sealed with an
impervious material (such as hydrated bentonite or a cement-bentonite mixture) or concrete to pre-
vent upward fow of water around the pipe. It may be necessary to temporarily lower the water level
in the relief well during the construction of this seal.
< PREVIOUS VIEW
6-161
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
GROUND
SURFACE
C
L
R
I
S
E
R
3
'
-
0
"
FLANGED RISER COVER
OUTFALL
FLAPGATE
OUTFALL
GUTTER
1' - 0"
CONCRETE BACKFILL
CONCRETE OR
BENTONITE BACKFILL
SAND BACKFILL
TOP OF FILTER PACK
4" (MIN.) FILTER PACK
SURROUNDING PIPE
6" (MIN.) I.D. SLOTTED SCREEN
MAY CONTAIN SECTIONS OF
RISER THROUGH VERY FINE
SAND OR SILT STRATA
BOTTOM PLUG
F
I
L
T
E
R
P
A
C
K
L
E
N
G
T
H
V
A
R
I
E
S
FIGURE 6.52 TYPICAL RELIEF WELL INSTALLATION
FIGURE 6.52 TYPICAL RELIEF WELL INSTALLATION
6.6.2.3.5 Horizontal Drains
Horizontal drains perform a similar function to relief wells, but provide more efective drainage
either in the foundation under the main body of an embankment or within the embankment. Most
ofen, horizontal drains are used to reduce excessive pore-water pressure within or beneath an exist-
ing embankment. A typical horizontal drain installation and details are shown in Figure 6.53. Hori-
zontal drains are normally sloped toward the discharge end.
Although horizontal drains 600 feet long and longer have been installed, lengths of 400 feet or less are
more common. The drains, which are usually sloted plastic pipe, are normally installed inside steel
drill rods, which are subsequently retracted. Because there is no soil flter around the sloted pipe, the
slots must be sized to prevent infltration of fnes based upon the rules discussed in Section 6.6.2.3.1.
In some cases, porous plastic pipe has been placed around the sloted pipe to limit the infltration of
fnes. The spacing required between horizontal drains is difcult to determine accurately. Pore-water
pressure should be monitored with piezometers to check the performance of an installation. A system
FIGURE 6.52 TYPICAL RELIEF WELL INSTALLATION
< PREVIOUS VIEW
6-162
Chapter 6
MAY 2009
FIGURE 6.53 EXAMPLE HORIZONTAL DRAIN INSTALLATION
FIGURE 6.53 EXAMPLE HORIZONTAL DRAIN INSTALLATION
DETAIL 2 - DRAIN COLLAR AND DRAINAGE COLLECTION
DETAIL 1 - 1 1/2" I.D. PVC PIPE WITH 0.02"
MACHINE-CUT SLOTS
MACHINE CUT SLOTS
EXTERNAL OR
INTERNAL COUPLING
NOMINAL 2" O.D.
SECTION A - A
(SEE DETAIL 2)
GROUTED ANNULUS
3/4" I.D.
PVC PIPE
6" CASING
GROUT BACKFILL
1 1/2" I.D.
PVC PIPE
DAM
2'
3/4" PVC PIPE
WITH 0.01" SLOTS
BENTONITE AND BURLAP
ANNULAR PACKER
GROUTED ANNULUS
SURROUNDING 6" CASING
BACKFILL GROUT
3
1
SURFACE CASING FOR
DRILLING COLLAR
1 1/2" I.D. SOLID
PVC PIPE DRAIN
3/4" I.D. PVC BLEEDER PIPE
2'
4'
COLLECTION
TANK
RUBBER
GASKETS
FLANGE
PLATE
FLANGED
OUTLET PIPE
STEEL FENCE
POSTS
A
A
NATIVE SOIL
DAM
SAND DRAINS
(MAY BE CONSTRUCTED
LATER, AS NEEDED)
140' LENGTH OF SLOTTED 1 1/2" I.D.
PVC PIPE (SEE DETAIL 1)
APPROXIMATE PHREATIC SURFACE
COLLECTOR PIPE AND DRAIN
(SEE DETAIL 2)
SOLID 1 1/2" I.D.PVC PIPE
BEDROCK
~ 3 SLOPE
EMBANKMENT CROSS SECTION
NATIVE SOIL
FIGURE 6.53 EXAMPLE HORIZONTAL DRAIN INSTALLATION
< PREVIOUS VIEW
6-163
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
of horizontal drains can be readily expanded if the capacity of the initial installation fails to produce
the required reduction in piezometric head.
The installation of horizontal drains can be dangerous in situations where large volumes of seepage
could lead to a piping failure at the drain borehole during installation. Horizontal drain systems
should be installed by an experienced contractor under expert supervision.
6.6.2.3.6 Impoundment Liners
To protect the groundwater some state agencies may require that seepage from fne coal refuse be
limited. This can be accomplished by using a layer of low-hydraulic-conductivity soil or geomem-
brane liners. Geomembranes are manufactured, low-hydraulic-conductivity synthetic materials that
function as barriers to liquids and vapors. When used as a liner on the botom and sides of a refuse
disposal impoundment, a geomembrane can impede leachate migration from overlying refuse into
the underlying soil and groundwater and can be used to collect the leachate for treatment. When a
geomembrane is used as a cap in the fnal cover over the impoundment, it prevents precipitation
from infltrating the coal refuse, thus minimizing or eliminating leachate generation.
Minimum requirements for geomembrane liners are ofen specifed in state regulations. Minimum
strength properties are provided in Table 6.51. However, if the application is on a slope or there is a
possibility that diferential setlement could occur, increasing stress and strain on the geomembrane, a
more conservative choice of membrane thickness may be appropriate. The strength of the liner is usually
reduced at seams. Standard tests for shear and peel strength should be performed on both factory and
feld seams. A determination must be made as to the minimum percentage of geomembrane material
strength that the seam itself must possess. This minimum seam strength criterion is typically incorpo-
rated into the quality assurance/quality control (QA/QC) program for liner installation. For evaluation
of survivability, the minimum seam strength should be determined early in the design stage if it is likely
that the geomembrane will be subjected to ultraviolet radiation for a signifcant period of time. If that is
the case, a material with a high resistance to ultraviolet light deterioration should be used.
TABLE 6.51 RECOMMENDED MINIMUM PROPERTIES FOR GENERAL
GEOMEMBRANE INSTALLATION SURVIVABILITY
Property and Test Method
Required Degree of Survivability
Low
(1)
Medium
(2)
High
(3)
Very High
(4)
Thickness ASTM D 1593 (mils) 20 25 30 40
Tensile ASTM D 882 (1-in strip) (lb/in) 30 40 50 60
Tear Resistance D 1004 (Die C) (lb) 5 7.5 10 15
Bursting Strength D 3787 (lb) 20 25 30 35
Impact Resistance D 3998 (ft-lb) 10 12 15 20
Note: 1. Low refers to careful hand-placement on very uniform, well-graded subgrade with light loads of a static
nature typical of vapor barriers beneath building foor slabs.
2.
Medium refers to hand- or machine-placement on machine-graded subgrade with medium loads
typical of canal liners.
3. High refers to hand- or machine-placement on machine-graded subgrade of poor texture with high loads
typical of landfll liners and covers.
4. Very high refers to hand- or machine-placement on machine-graded subgrade of very poor texture with
high loads - typical of reservoir covers and liners for heap leach pads.
(ADAPTED FROM KOERNER, 2006)
< PREVIOUS VIEW
6-164
Chapter 6
MAY 2009
6.6.2.3.7 Foundation Seepage Cutofs
In addition to measures to control seepage in embankments, foundation treatments such as cutof
trenches backflled with compacted low-hydraulic-conductivity materials are also routinely incorpo-
rated into refuse facility designs. Foundation cutof trenches are discussed in Section 6.3.
6.6.2.3.8 Impoundment Water and Slurry Deposition Management
Management of impoundment water and slurry deposition at a coal refuse disposal facility are oper-
ational measures for controlling and mitigating the efects of seepage. Maintenance of a low level
of clarifed water in an impoundment reduces the hydraulic head and the source volume for seep-
age through the embankment. Additionally, deposition of coal refuse slurry at or near the upstream
embankment face and the resulting build up of a delta of fne refuse against the embankment will
reduce seepage into the embankment if the fne refuse has a lower hydraulic conductivity than the
embankment material. To develop and maintain an efective fne refuse delta across the upstream
face of the embankment, periodic relocation of the slurry discharge point or use of multiple discharge
points may be required. At some facilities, small cells have been constructed near the upstream end
of a refuse impoundment for storage of water for recirculation to the preparation plant. Such provi-
sions reduce the likelihood of water being impounded directly against a coarse refuse embankment
and associated seepage concerns.
6.6.2.4 Seepage Measurements
When seepage rates or piezometric levels are an important factor in the performance of a disposal
facility, estimates of the phreatic surface, magnitude and rate of dissipation of pore pressures,
and quantity of seepage collected in internal drains should be made. These predictions should
be checked during and afer construction by instrumenting the refuse embankment (and founda-
tion if necessary) and measuring changes in the phreatic surface and pore pressures. Procedures
for monitoring groundwater levels and pore-water pressures are discussed in Chapter 13. Field
piezometer data provide a basis for updating performance predictions and for possible modifca-
tion of design or construction procedures. Weirs or parshall fumes at the outlets of internal drains
can be used to measure the seepage collected from an embankment for comparison to results of
seepage analyses associated with the internal drain design. In addition, instrumentation can pro-
vide a check on the in-situ embankment hydraulic conductivity, as compared to that assumed in
the analyses and/or determined by laboratory or feld hydraulic conductivity tests. Monitoring
of seepage conditions is also important for detection of unanticipated changes in the saturation
level or seepage quantity and can provide an early indication of a problem such as clogging of an
underdrain.
6.6.3 Settlement Analysis
6.6.3.1 Conditions Requiring Deformation Analysis
Setlement of a coal refuse embankment occurs as a result of embankment compression, foundation
compression, plastic deformation, diferential setlements, mine subsidence or a combination of these
efects. Setlement is usually important if an embankment will impound water or if it will serve as
the foundation for future construction. The design of any impounding embankment must limit setle-
ment of the crest so that the freeboard is not reduced below the allowable limit. Embankments should
be cambered so that the crest is elevated relative to the abutments to compensate for the increased
setlement that typically occurs near the center of the embankment where the foundation overburden
and embankment height are the greatest. It is also important that the embankment does not setle so
much that the hydraulic conductivity characteristics of the embankment are signifcantly changed or
the potential for piping or internal erosion due to cracking is created. Embankment and foundation
setlement can also afect the performance and structural integrity of conduits and other structures.
The setlement that can be tolerated by a structure depends upon its function.
< PREVIOUS VIEW
6-165
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
The magnitude of setlement under self-weight experienced by a coal refuse embankment cannot
be accurately estimated from fundamental stress-strain properties. The most useful information for
computing embankment setlement is performance data from instrumented earth and rockfll dams.
These data indicate that setlement of well-compacted earth dams due to embankment compression
ranges from less than 1 percent for dams constructed of non-plastic soils to more than 4 percent for
dams constructed of highly-plastic, fne-grained soils. Measured setlements of rockfll dams have
ranged from essentially no setlement for well-compacted and sluiced rockfll placed in thin lifs to 10
percent or more for unsluiced rock placed in thick lifs. Therefore, the amount a coal refuse embank-
ment will setle due to compression can range from 1 percent or less for well-compacted materials to
8 to 10 percent for uncompacted materials. However, a coal refuse embankment is ofen constructed
over 20 to 30 or more years, as compared to one to three years for an earth or rockfll dam. Thus,
nearly all coal refuse embankment setlement will likely occur during construction, and additional
setlement afer abandonment will be minimal.
If an embankment foundation consists of dense glacial till, dense sand and gravel, or rock, it will
deform only slightly under the weight of the embankment because, at the stress levels experienced,
both the foundation materials and any existing pore-water are essentially incompressible. Thus, set-
tlement of such embankments due to foundation setlement will be minimal. If the foundation con-
sists of saturated fne-grained soils, there may be signifcant setlement from consolidation due to
the time-dependent expulsion of pore water from the soil. In such cases, the development of excess
pore-water pressures can also afect foundation stability and these excess pressures need to be taken
into account in embankment stability analyses.
Upstream construction of coal refuse disposal embankments over fne coal refuse deposits typically
results in consolidation setlements that can give rise to elevated pore pressures if construction occurs
rapidly. To prevent localized instability, construction procedures should be carefully planned. Moni-
toring of pore pressure may be needed as part of monitoring of embankment stability and controlling
the rate of construction. These issues are discussed in Chapter 11.
Consolidation occurs slowly, sometimes over several months or years. Consolidation setlement can
be a very important design factor because, in addition to causing damage to drainage pipes and
structures, it can afect aspects of the entire disposal facility. Problems created by consolidation may
not become apparent until afer the disposal facility begins operation, at a time when the setlement
could create a safety hazard and when corrections are most expensive. If consolidation is associ-
ated with embankment construction over sof clay deposits, the total setlement could be substantial.
When a slurry impoundment is to be eliminated by backflling the remaining reservoir, the setlement
of the cap material must be considered in the design. Sufcient cap material must be placed such that
following long-term setlement positive surface drainage will be maintained and a depression that
will collect water is not created.
The magnitude of the foundation consolidation setlement depends on the weight of the embank-
ment, the depth and thickness of the compressible strata in the foundation, and the compression
indices of these compressible strata. Compression indices can be obtained from laboratory consoli-
dation tests on undisturbed samples taken from the compressible strata. Compression indices from
feld setlement records for other disposal facilities underlain by similar compressible strata with
similar moisture contents and index properties can also be used to predict the potential range of
setlement.
The rate at which setlements occur is a function of the rate of change in vertical stress (directly pro-
portional to the rate of construction), the hydraulic conductivity of the compressible material, and the
drainage characteristics of the foundation. The rate of setlement is much more difcult to estimate
than the magnitude of setlement. Computations based on laboratory consolidation test data may be
< PREVIOUS VIEW
6-166
Chapter 6
MAY 2009
inaccurate because the rate at which foundation setlements occur is ofen controlled by minute geo-
logical characteristics that may not be detected even by carefully conducted foundation studies.
Consolidation and conventional laboratory tests used for measuring the consolidation characteristics
of a soil are discussed in Section 6.5.6.
Plastic deformation of an embankment and its foundation represents only a portion of the observed
setlement at a coal refuse embankment. Plastic deformation is the lateral spreading of an embank-
ment at the base coincident with setlement. Although difcult to determine, it is important to allow for
the amount of plastic deformation that will occur because it can cause extension of pipes constructed
through the embankment. Although the magnitude of such an extension may be as much as several
percent of the total length of the pipe, damage can usually be avoided if the extension occurs uniformly
over the length of the pipe. However, if the extension is concentrated or if the pipe cannot tolerate sig-
nifcant extension, it may separate and leak or fracture and collapse. Such an outcome could seriously
afect embankment stability. To minimize the risk of this type of failure, pipes should either be placed in
the foundation below the plastic zone or should be designed to tolerate the anticipated extension.
Diferential vertical setlements can also cause damage to a coal refuse embankment and to interior
structures. Diferential foundation setlements, particularly if they occur over small distances, can
result in embankment cracks and lead to subsurface erosion. Diferential setlements can also cause
damage to pipe drains and decant lines installed within or beneath an embankment. When consider-
ing the potentially damaging efects of foundation setlements beneath a planned coal refuse embank-
ment, it should be assumed that some areas of the foundation may have localized setlements at least
twice as great as the predicted overall setlement. Therefore, if substantial foundation setlements are
expected, placing pipelines within or beneath an embankment should be avoided. If a decant pipe is
placed in an embankment, it should be located as far as possible from the areas where the setlement
is anticipated to be greatest.
Embankments constructed above or adjacent to underground mine voids can experience setlement
and lateral distortion caused by subsidence. Depending on the characteristics of the overburden rock,
the depth and type of mining, and whether the mining is pre-existing or is occurring during dis-
posal operations, the subsidence can have a signifcant efect on the stability of the embankment and
the retained coal refuse. Additional discussion of subsidence issues can be found in the following
Manual sections:
Section 5.4 for mine subsidence considerations during facility design
Section 8.4 for analysis of mine subsidence
Section 8.5 for mine breakthrough potential evaluations
Sherard (1973) reports that earth dam cores of almost all soil types have experienced cracking.
Therefore, designers should carefully consider the surface topography in which, and the founda-
tion condition and materials on which, the embankment will be constructed. For dams in narrow
valleys, cracking paterns such as those shown in Figure 6.54 have been observed. Transverse ver-
tical cracks ofen develop within about 15 to 50 feet from the end of the crest, and in the central
portion where the embankment is the highest, vertical cracks ofen develop on or near the crest,
as shown in Figure 6.54.
Based on the results of laboratory testing reported by Leonards and Narain (1963) and statisti-
cal evaluation of earth dam performance by Biarez et al. (1970) and Londe (1970), as reported by
Sherard (1973), the tensile strain at which frst cracking occurs is in the range of 0.1 to 0.3 percent.
In most of these case histories, the embankments were placed with a moisture content from 1 to
< PREVIOUS VIEW
6-167
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
4 percent dry of optimum (Wilson and Squier, 1969). Tensile strains in earth dams should be con-
sidered, and precautionary measures to limit the magnitude of tensile strains should be incorpo-
rated, as necessary. Precautionary measures could include: (1) incorporating more crack resistant
soils in the dam cross section where cracking is most likely, (2) taking added care in foundation
preparation to avoid/minimize large changes in foundation grades, or (3) increasing the width of
the embankment cross section. Regardless, designers should assess the potential efects of factors
that can cause diferential setlement and incorporate appropriate measures to limit the develop-
ment of cracks during construction and long-term operations.
Broadly-graded coarse coal refuse, placed at or above optimum moisture content and at relatively
low construction rates that allow setlement to occur as fll is gradually added, is generally consid-
ered to be less susceptible to cracking than earthen materials, which are frequently subject to longer-
term consolidation, afer completion of construction.
6.6.3.2 Settlement and Deformation Analysis
Embankment and foundation setlement and deformation are typically evaluated using principles of
elastic behavior and/or Terzaghi consolidation theory. For most setlement and deformation analyses
associated with coal refuse disposal facilities, stresses in foundations and embankments and changes in
stress due to load application (e.g., embankment loading on the foundation) can be estimated using elas-
tic theory. Closed-form equations and graphical methods for estimating stresses for a variety of geomet-
ric loading cases, homogeneous and layered subsurface profles, and isotropic and anisotropic subgrade
conditions are available in Poulos and Davis (1974). Methods for estimating setlement due to elastic
compression and consolidation are described in numerous foundation engineering textbooks (Terzaghi
et al., 1996; Hunt, 1986; Holtz and Kovacs, 1981) and design manuals (CGS, 2007; DOD, 2005).
Since the early 1970s, fnite element (FE) methods that permit realistic deformation analysis of earthen
embankments and foundations have been developed. As summarized by Duncan (1996), some of the
special features of FE methods with application to embankments and foundations are:
FIGURE 6.54 CRACKING PATTERNS OBSERVED FOR DAMS IN NARROW VALLEYS
OPEN CRACKS
PLAN VIEW A
OPEN CRACKS
PLAN VIEW B
FIGURE 6.54 CRACKING PATTERNS OBSERVED FOR DAMS IN NARROW VALLEYS
(FELL ET AL., 2005)
LONGITUDINAL SECTION A
EXAGGERATED CREST
SETTLEMENT
ROCK
PROJECTION
LONGITUDINAL SECTION B
OPEN CRACK
OPEN
CRACKS
FIGURE 6.54 CRACKING PATTERNS OBSERVED FOR DAMS IN NARROW VALLEYS
< PREVIOUS VIEW
6-168
Chapter 6
MAY 2009
Versatile tool for analysis of stresses and movements in earth masses, including:
Stresses, deformations and pore pressures in embankments and foundations
Conditions during construction such as consolidation and embankment
compression due to self-weight
Potential for cracking and hydraulic fracturing
Can model nonlinear stress-strain behavior, non-homogeneous conditions, and
changes in geometry such as embankment construction
Sofware are available with graphical pre- and post-processors to facilitate data input
and evaluate analysis results.
FE analyses require input data such as: (1) defnition of the in-situ stress conditions of the foundation
materials, (2) the stress-strain properties of the foundation and embankment materials, and (3) the
sequence of construction.
For problems that involve a natural soil or rock deposit or an existing fll, the state of stress in the soil
mass prior to the beginning of construction or loading must be specifed because:
For incremental analyses, the changes in stress calculated during each increment
are added to the stresses at the beginning of the increment in order to evaluate the
stresses at the end. To begin this process, it is necessary to know the initial in-situ
stresses.
The stifness of the soil is a function of the stresses in the soil.
The in-situ stresses can be measured, but are usually estimated. For level ground where at-rest
pressure conditions would be expected, the vertical stresses are usually assumed to be equal to the
overburden pressure, and the horizontal stresses are assumed to be the at-rest lateral earth pressure
coefcient K
o
times the overburden pressure. The value of K
o
is usually estimated based upon empiri-
cal relationships (Jaky, 1944; Mayne and Kulhawy, 1982). For initially sloping ground, one procedure
that has been used is performing a gravity turn-on analysis (i.e., applying vertical forces representing
the weight of the material to an initially unstressed mesh) and then changing the horizontal stresses
to K
o
times the calculated vertical stresses.
The stress-strain properties of modeled materials play a critical role in fnite element analyses.
For most deformation analyses of embankments and foundations where constituent materials
are not stressed close to failure and strains are small, stress-strain behavior can be represented
by a linear elastic model. However, for rock foundations, the modulus should refect the defor-
mation characteristics of the rock mass through modifcation of the deformation of intact (i.e.,
unfractured) rock using a rock mass classifcation system such as that proposed by Bieniawski
(1989). The Bieniawski geomechanics classifcation system involves modifcation of the modulus
of intact rock as determined from unconfned compression tests (Section 6.5.9.2) using the fol-
lowing relationship:
E
m
= 2 RMR 100 (6-29)
where:
E
m
= modulus of the rock mass (force/length
2
)
RMR =
rock mass rating; this parameter accounts for the efects of intact rock
strength, rock quality designation (RQD), joint spacing, joint condition,
joint orientation and groundwater (dimensionless)
< PREVIOUS VIEW
6-169
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
RMR increases with rock quality with a range from 0 to 100. Equation 6-29 is valid for RMR 55. For sofer
rocks, the following relationship, which was proposed by Serafm and Pereira (1983), can be used:
E
m
= 10
(RMR-10)/40
(6-30)
For soils it is ofen necessary to use stress-strain relationships that account for nonlinear behavior
and the variation of soil modulus with confning pressure. Table 6.52 summarizes the types of stress-
strain models typically used and their respective advantages and limitations.
TABLE 6.52 STRESS-STRAIN RELATIONSHIPS USED FOR FINITE ELEMENT
DEFORMATION ANALYSES OF EMBANKMENTS AND FOUNDATIONS
Stress-Strain Relationship Advantages Limitations
Linear Elastic Simplicity
Can only model real soil behavior at low
stress levels and small strains
Multi-linear Elastic
Can model any shape stress-strain
curve for ductile materials
Must be developed on a case-by-case
basis to approximate stress-strain
behavior
Hyperbolic
Can model nonlinear behavior; para-
meters have physical signifcance and
can be evaluated by triaxial testing
Inherently elastic; does not model plastic
deformations in a fully logical manner
(ADAPTED FROM DUNCAN, 1996)
Analyses should simulate as closely as possible the actual construction or loading sequence asso-
ciated with the structure being analyzed. This can be accomplished by adding elements to simu-
late fll placement, removing elements to simulate excavation, and applying loads in increments.
Other processes that can be modeled in FE analyses include raising or lowering phreatic levels and
consolidation.
Comparison of the results of FE analyses with feld measurements shows there is a tendency for cal-
culated deformations to be larger than measured deformations. Duncan (1996) ofered the following
reasons for this diference:
Soils/materials in the feld tend to be stifer than laboratory test samples at the same
density and moisture content due to aging efects.
Average feld densities are higher than the specifed minimum dry density, which is
ofen used as the target for preparing samples for laboratory testing.
Soils/materials sampled in the feld for laboratory testing are disturbed by the sam-
pling process.
Most feld conditions approximate plane strain whereas triaxial tests are routinely
used for laboratory characterization.
2D fnite element analyses overestimate deformations of embankments constructed
in V-shaped valleys with steep walls.
Analysis of embankments using the FE method has demonstrated the considerable potential of the
approach and has identifed the sources of uncertainty that engineers should be aware of. These
uncertainties are primarily associated with difculties in accurately predicting the in-place density
and moisture content of soils/materials in the feld and difculties anticipating the sequence of opera-
tions that will be followed during construction.
< PREVIOUS VIEW
6-170
Chapter 6
MAY 2009
6.6.3.3 Deformation Control Measures
In a slurry impoundment, fne coal refuse typically has a low unit weight (as compared to typical soils),
low hydraulic conductivity and low coefcient of consolidation. It is placed by peripheral discharge of low-
solids-content slurry, either by single-point discharge or using multiple discharge locations. Because of this
method of placement, fne coal refuse tailings can be very sof and susceptible to long-term setlements.
The rate of consolidation at existing refuse disposal sites can be expedited by the installation of pre-
fabricated vertical (PV) drains into the tailings at close vertical spacing to drain excess pore-water
pressures. PV drains consist of a high-fow polymeric core wrapped in a non-woven geotextile. The
drains are installed using static, vibratory or jeting methods. Guidelines for the engineering design
of PV drains are presented in Rixner et al. (1986). Brown and Greenaway (1999) describe instances
where PV drains were used to expedite consolidation of uranium mill tailings before construction
of a clay cap required for abandonment. Prefabricated drains have also been placed horizontally on
the surface of a coal refuse impoundment prior to upstream construction to speed consolidation and
dissipation of pore pressures (Thacker et al., 1988). The efectiveness of the drain installations was
demonstrated by the control of pore pressures and consolidation. Drilled horizontal drains have also
been used to lower pore-water pressures in tailings impoundments.
Drainage measures to accommodate dissipation of pore pressures and to consolidate fne materials
such as fne coal refuse can be incorporated into facility design, as discussed in Section 6.3.
Adaptations of shallow and deep soil mixing technologies developed for improving loose and sof
soil deposits have been used for the in-place solidifcation of fne coal refuse. Bazn-Arias et al. (2002)
describe the use of these methods to stabilize fne coal refuse using custom-designed equipment to
blend a cement/fy ash slurry with coal refuse for supporting a highway embankment over a slurry
pond. QC testing of cured samples resulted in a 28-day, unconfned compressive strength greater
than 100 psi and peak shear strength parameters of > 45 and c > 13 psi.
Lime has also been used to stabilize fne coal refuse. Lime stabilization causes the refuse to behave
as an overconsolidated material. However, when load is applied that exceeds the apparent maxi-
mum past consolidation pressure, the stabilized refuse tends to collapse and return to its previous
unstabilized behavior. Therefore, it is recommended that potential use of lime stabilization of fne
coal refuse be carefully evaluated through laboratory and feld performance testing before imple-
mentation in the feld.
6.6.3.4 Deformation Measurements
When predicted setlements are an important factor in the design of a coal refuse disposal facility,
estimates of the magnitude and the rate of setlement should be made using sophisticated analyses
with the best available data. The performance of a coal refuse disposal facility should be monitored
through installation of an instrumentation system and comparison of observed data to predicted
deformations. Depending on site conditions and project requirements, deformations and deforma-
tion rates can be monitored by:
Surface monuments (vertical and horizontal displacements)
Setlement gages and extensometers (vertical displacements)
Inclinometers (lateral displacements in slopes)
Piezometers (piezometric heads in the embankment and foundation)
As described in Chapter 13, instrumentation should be installed to verify that acceptable levels of
performance are being achieved and to provide a check on design assumptions.
< PREVIOUS VIEW
6-171
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
6.6.4 Slope Stability Analysis
6.6.4.1 Conditions Requiring Stability Analysis
The design of a new coal refuse disposal facility or the expansion or modifcation of an existing
facility requires that the stability of compacted embankments and natural soil and rock slopes and
foundations be assessed. The most critical cross sections and cases must be analyzed. Factors consid-
ered in selecting the most critical cross sections include: (1) slope of the embankment, (2) height of
the embankment, (3) foundation conditions, (4) pore-pressure conditions, and (5) presence of lower
strength material zones within the embankment/impoundment (e.g., upstream construction cross
sections). Critical cross sections can include upstream or downstream embankment slopes. Impor-
tant cases include the following:
Long-term or fnal embankment confguration
Intermediate stages of development that may include critical cross sections
related to slope, height, foundation, embankment material properties, or
phreatic levels or pore-pressure conditions
Short-term, end-of-construction conditions for stages on sof, compressible
materials
Rapid loading of fne coal refuse deposits during initiation of upstream
construction that possibly leads to an unacceptably low factor of safety
against bearing failure of the fne refuse underlying the area of embank-
ment raising
Rapid drawdown of the impoundment
Seismic loading and strength loss or increases in pore pressure
Phreatic levels and pore-pressure conditions should be determined on the basis of seepage analyses
and, for existing facilities, by correlations with piezometric measurements.
In selecting an upstream construction cross section for analysis, the interface between the coarse
refuse embankment and fne refuse deposits should be determined based on subsurface exploration
(borings and cone penetration tests are recommended) if an existing facility is being analyzed. For
new facilities, analysis cross sections should be determined based on staging calculations considering
upstream construction procedures and fne coal refuse behavior. Staging calculations will identify
the approximate level of the setled fne refuse at the initiation of upstream construction. The follow-
ing facility-specifc issues must be considered:
Type of setled fne coal refuse that will support the upstream stage (e.g., sandy or
clayey fne refuse)
Presence and location of impoundment pool relative to the extent of the upstream
push out
Staging area for coarse coal refuse to be used for the push out
Presence of excess pore-water pressures in areas where the fne refuse is not fully
consolidated
Equipment and lif thickness that can be used for the push out
Monitoring program that can be implemented and used in controlling the push out
The behavior of fne coal refuse in response to upstream construction may include consolidation,
mixing with the coarse refuse during the initial push out, and development of a zone of assimi-
lation. Appropriate material strengths and levels of excess pore-water pressure need to be used
< PREVIOUS VIEW
6-172
Chapter 6
MAY 2009
in analyses. Experience at similar facilities can provide valuable insight into material behavior,
construction procedures to be employed, and the resulting zone of mixing. The strength in this
zone of mixing should be determined based on the relative properties of the coarse and fne refuse
materials at the site. In some cases, an exploration program at an existing similar facility might be
undertaken for determining the desired properties. Typically, this zone of mixing does not play
a signifcant role in the static analysis of downstream embankment slopes, but can be critical for
seismic analysis and upstream slopes. The following guidance has been developed for deciding
whether potential upstream failure surfaces are critical to the seismic stability and deformation of
the embankment (MSHA, 2007):
Potential upstream slope stability failure surfaces that terminate on the crest of the
embankment stage should provide an acceptable factor of safety such that the integ-
rity of the embankment and impounding capacity of the facility are maintained (i.e.,
if a portion of the embankment becomes unstable, a sufcient section of the crest will
remain intact to prevent release from the impoundment).
Potential deformation of the crest of the embankment should not result in the threat
of a release from the impoundment (i.e., sufcient freeboard must be available to
compensate for the maximum amount of crest setlement).
Another critical case where fne coal refuse characteristics are important is related to recovery or
re-mining of fne coal refuse within an impoundment. This situation typically involves the devel-
opment of an excavation plan with interim and fnal coal refuse slopes that must have acceptable
factors of safety.
Additional conditions requiring slope stability analyses may arise due to situations in which the
shear strength decreases or stress level in an embankment increases. Duncan and Wright (2005) cite
the following causes for a decrease in shear strength:
Increased pore pressure (reduced efective stress) due principally to a rise in ground-
water level or increased seepage during periods of heavy rainfall
Cracking near the crest of a slope due to tension and factors such as soil desiccation
Swelling of highly plastic and heavily overconsolidated clays
Development of slickensides due to shear in highly plastic clays resulting from shear
on distinct slip planes
Decomposition of rock in flls due to inadequate breakdown during compaction and
weathering as a result of weting and drying
Creep of highly plastic clays under sustained load
Leaching of chemical constituents in the soil matrix
Strain sofening of britle soils leading to progressive failure
Weathering of rocks and indurated soils due to physical, chemical and biological
processes
Cyclic loading and loss of strength due to liquefaction (Chapter 7)
Possible causes for an increase in shear stress include:
Increased loads at the top of a slope
Water pressure in tension cracks at the crest of a slope
Increase in soil weight due to increase in moisture content
< PREVIOUS VIEW
6-173
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
Excavation at the botom of a slope
Rapid drawdown of an impoundment (upstream slope)
Earthquake loading
Traditionally, slope stability has been evaluated using limit equilibrium analyses whereby the forces
tending to decrease stability are compared to the forces tending to increase stability. These types of
analyses are generally conducted using limit equilibrium slope stability computer programs. Since
the early 1970s, fnite element methods of analysis have improved to the point where realistic stabil-
ity/deformation analysis of soil slopes is possible. The following section provides an overview of
stability analyses for coal refuse disposal facility embankments.
6.6.4.2 Methods of Stability Analysis
6.6.4.2.1 Limit Equilibrium Stability Analysis
The stability of refuse embankments is usually solved by limit equilibrium methods of analysis. These
analyses are conducted by calculating the minimum factor of safety (FS) for a slide surface through
the slope as follows:
FS =
Available shear strength
=
s
(6-31)
Equilibrium shear stress
If a large number of potential slide plane surfaces are assumed, the surface with the minimum factor
of safety is a numerical representation of the relative safety of the slope. If FS = 1, a slope is in a state of
just-stable limit equilibrium. Because of the uncertainty related to the geometry of the actual slide
plane surface, the controlling soil properties, the pore-pressure distribution in the slope, and other
factors that may afect the stability of a slope, slopes for water impounding embankments should
be designed with FS equal to at least 1.5. A higher factor of safety should be used where factors that
afect slope stability (e.g., limited testing has been performed) are less certain. The available shear
strength (s) is defned in terms of the angle of friction ( or ) and cohesion (c or c) of the soil along
the slide plane surface using soil properties determined from in-situ or laboratory tests.
Two approaches can be used to satisfy static equilibrium of a slope. The frst and much less com-
monly used approach is to assume a single free-body bounded by the face of the slope and slide
plane surface. Examples of this approach are the infnite slope, log spiral and Swedish slip circle
methods. The second approach involves dividing the slope into a number of vertical slices that
extend between the face of the slope and slide plane surface. Examples of this approach are the
ordinary method of slices, simplifed Bishop method and Spencers method. Regardless of the
approach used there are more unknowns (e.g., forces, location of forces, FS) than equilibrium equa-
tions, so the problem is statically indeterminate. Therefore, assumptions must be made to render
the problem determinate. Examples of such assumptions include inclination of interslice forces, the
location of the normal force at the base of a slice and the relationship of interslice shear force to the
interslice normal force.
Some slope stability analysis methods are based solely upon force-equilibrium principles, while other
methods involve satisfaction of all conditions of equilibrium. The characteristics of various equilib-
rium methods for slope stability analysis are summarized in Table 6.53. If force-equilibrium methods
are used (Lowe and Karafath, 1960; USACE, 2003), the factor of safety is afected signifcantly by the
assumed inclinations of the side forces between slices. Thus, force-equilibrium procedures are not
as accurate as methods that satisfy all conditions of equilibrium (Janbu, 1973; Spencer, 1967; Mor-
genstern and Price, 1965). The maximum range of results for methods that satisfy all conditions of
equilibrium is generally less than 12 percent. Thus, with an average accuracy of about plus or minus
6 percent, a factor of safety calculated using procedures that satisfy all conditions of equilibrium can
< PREVIOUS VIEW
6-174
Chapter 6
MAY 2009
be considered to be acceptably accurate, because for practical purposes key parameters such as slope
geometry, unit weight, shear strength, and pore-water pressure cannot be defned with an accuracy of
plus or minus 6 percent. Therefore, any method that satisfes all conditions of equilibrium should be
sufciently accurate for impoundment design and analysis. Additional information relative to selec-
tion of a slope stability analysis method can be found in Duncan and Wright (1980).
For slopes composed of nearly homogeneous materials, both analysis and observation of actual
failures indicate that the failure surface can be approximated with sufcient accuracy by a circular
arc. For such cases, procedures that do not satisfy all conditions of equilibrium may be acceptably
TABLE 6.53 CHARACTERISTICS OF EQUILIBRIUM PROCEDURES
FOR SLOPE STABILITY ANALYSIS
Procedure Application
Infnite Slope
Homogeneous cohesionless slopes where stratigraphy restricts slip surface to
shallow depths and parallel to slope face. Very accurate where applicable.
Logarithmic Spiral
Applicable to homogenous slopes; accurate. Potentially useful for developing
slope stability charts and used in some software for design of reinforced slopes.
Swedish Circle
Applicable to slopes where = 0 and for relatively thick zones of weaker
materials where slip surface can be approximated by a circle.
Ordinary Methods of Slices
(Fellenius, 1922)
Applicable to non-homogeneous slopes and c - soils where slip surface
can be approximated by a circle. Very convenient for hand calculations, but
inaccurate for effective stress analyses with high pore pressures.
Simplifed Bishop
(Bishop, 1955)
Applicable to non-homogeneous slopes and c - soils where slip surface can
be approximated by a circle. More accurate than OMS, especially for effect tive
stress analyses with high pore pressures. Calculations can be performed by
hand or spreadsheet.
Force Equilibrium Methods
(Lowe and Karafath, 1960; USACE,
2003)
Applicable to virtually all slope geometries and soil profles. The only pro cedure
suitable for hand calculations with non-circular slip surfaces. Less accurate
than complete equilibrium procedures, and results are sensitive to assumed
inclination for interslice forces.
Janbu Generalized
Procedure of Slices
(Janbu, 1973)
Satisfes all conditions of equilibrium. Applicable for any shape of slip surface.
Numerical problems are encountered more frequently than with some other
methods.
Spencer
(Spencer, 1967)
Satisfes all conditions of equilibrium. Accurate procedure applicable to virtually
all slope geometries and soil profles. Simplest complete equilibrium procedure
for computing factor of safety.
Morgenstern and Price (1965)
Satisfes all conditions of equilibrium. Accurate procedure applicable to virtually
all slope geometries and soil profles. Rigorous, well-established complete
equilibrium procedure.
Chen and Morgenstern (1983)
Satisfes all conditions of equilibrium. An updated Morgenstern and Price
procedure. Rigorous and accurate procedure applicable to any slip surface
shape and slope geometry, loads and soil profles.
Sarma (1973)
Satisfes all conditions of equilibrium. Accurate procedure applicable to
virtually all slope geometries and soil profles. Convenient complete equi librium
procedure for computing seismic coeffcient required to produce a given factor
of safety. Side force assumptions are diffcult to implement for all but simple
slopes.
(DUNCAN AND WRIGHT, 2005)
< PREVIOUS VIEW
6-175
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
accurate. For non-homogeneous slopes or embankments, embankments supported on a foundation
with a weak zone and impoundments lined with geomembranes or geosynthetic clay liners, limit-
equilibrium procedures that are suitable for any shape slip surface and that satisfy all conditions of
equilibrium must be used. Designers should be especially alert to the presence of a weak layer or
layers upon which sliding may occur. In such cases, the factor of safety for wedge-type surfaces coin-
cident with the weak layer must be evaluated in addition to the usual failure surfaces. If movement
has already occurred in a zone of material that is included in a stability analysis, then residual shear
strength may be applicable.
An implicit assumption in equilibrium analyses of slope stability is that the stress-strain behav-
ior of the constituent material is ductile (i.e., it does not have a britle stress-strain curve where
the shearing resistance drops of afer reaching a peak). This limitation results from the fact that
limit-equilibrium methods provide no information regarding the magnitudes of the strains within
a slope, nor any indication about how they may vary along the slip surface. Therefore, unless the
strengths used in the analysis can be mobilized over a wide range of strains (i.e., the soil exhibits
ductile stress-strain behavior) there is no guarantee that the peak strength can be mobilized simul-
taneously along the full length of the slip surface. Where multiple embankment or impoundment
zones are traversed by the slip surface, strain compatibility for each material should be evaluated.
For instance, coarse coal refuse typically mobilizes peak strength at a lower strain than fne coal
refuse and cohesive foundation soils. Thus, the stability analysis should be based on strength at
compatible strains, particularly if there is a drop-of in strength with large strains. If the shearing
resistance of one material drops of afer the peak is reached, progressive failure can occur, and the
shearing resistance that can be mobilized at some parts of the failure surface may be smaller than
the peak strength. For this situation, a reliable approach is to use the residual strength rather than
the peak strength in the analysis.
For coal refuse, earth and rockfll embankments, the following critical embankment conditions should
be evaluated:
1. High pore-water pressures are present in both the embankment and foundation.
This condition occurs most ofen during or at the end of construction, particularly if
construction is rapid, the slope materials have low hydraulic conductivity, and con-
struction conditions are wet. For a coal refuse embankment, the rate of construction
usually is not fast enough to cause high pore pressures in the foundation materials.
An exception is when a thick layer of saturated clay underlies the embankment. For
this case pore pressures during construction should be estimated, and piezometers
should be installed to facilitate maintaining pore pressures within acceptable limits
during construction. The rate of construction can also be an issue if an upstream
construction pushout is constructed rapidly. Stability checks may be required both
during construction and at the end of construction when an embankment is con-
structed over setled fne refuse using the upstream method.
2. Steady seepage has developed within the embankment and may have saturated a
large part of the downstream slope. This condition occurs most ofen afer long-term
operation of an impounding embankment at full storage level, particularly if the
slope materials had a high hydraulic conductivity. For compacted embankments,
placement of refuse is usually slow enough that excess pore pressures will ade-
quately dissipate. For situations where wet materials are placed in thick lifs, (e.g.,
flter cake or combined refuse in upstream embankment zones) excess pore pres-
sures can develop, although generally the rate of construction is slow enough that
the excess pore pressures will dissipate. Some slurry impoundments are designed to
< PREVIOUS VIEW
6-176
Chapter 6
MAY 2009
store the runof from the design storm and release it relatively slowly. In such cases,
the water level may rise above the level of the slurry delta and be in direct contact
with relatively permeable upstream slope material. If storm water is impounded
against the upstream slope long enough for steady-state pore-water pressures to
develop, this condition could represent a critical stability scenario.
3. The impoundment water level drops very quickly afer steady seepage has devel-
oped within the embankment. This condition is generally referred to as rapid draw-
down and can be critical to the stability of the upstream slope of the embankment.
USBR (1987a) provides general guidance for water-detention and storage dams
considering susceptibility of earth fll materials (based on USCS classifcation) to
rapid drawdown loading (drawdown rate of 6 inches or more per day following
prolonged storage at high reservoir levels). For most coal refuse embankments that
impound slurry and runof water, rapid drawdown is either not possible or is not
an issue because the embankment material is generally free draining. For embank-
ments constructed of low-hydraulic-conductivity material (less than 10
-4
cm/sec)
and designed to store storm runof for subsequent release through a decant pipe, the
potential efects of rapid drawdown should be considered. Another situation where
rapid drawdown may need to be considered is during remining of an impoundment
for recovery of additional coal from the fne coal refuse.
4. The embankment is subjected to earthquake loading during embankment con-
struction, operation, or following abandonment. Issues related to the analysis and
design of embankments that are subjected to earthquake loading are discussed in
Chapter 7.
Analyses for the frst three critical stability conditions listed above must refect the rate of construc-
tion and pore-pressure conditions. The following analyses are typically employed:
Total-Stress Analysis is used in situations where the pore-water pressure (u) that
would act on the potential failure surface at failure is unknown and cannot be reli-
ably estimated. The embankment stability is analyzed in terms of total stress (i.e.,
the stress between the individual soil grains plus the pore pressure). This method of
stability analysis is generally considered most appropriate for evaluating relatively
short-term loading conditions such as end-of-construction and rapid drawdown of
the impoundment.
Efective-Stress Analysis is used in cases where the pore-water pressure (u) that
would act on the potential failure surface at failure is known or can be reliably
estimated. The embankment stability is analyzed in terms of efective stress (i.e., the
total stress minus the pore pressure). This method of stability analysis is generally
considered most appropriate for evaluating long-term conditions afer the transient
efects related to construction and seepage have ended.
Selection of the appropriate conditions for analysis requires knowledge of soil behavior under
drained and undrained conditions and evaluation of the conditions that will control drainage in the
feld. Shear strengths, water and pore pressures, and unit weights for slope-stability analyses are
summarized in Table 6.54.
A useful guide for determining whether total- or efective-stress methods of analysis are applicable
relates to whether the soils comprising the foundations and refuse embankment are free draining
or impermeable. Free-draining soils are those able to drain completely during the construction or
loading period. Impermeable soils are those that cannot drain completely during the construction or
< PREVIOUS VIEW
6-177
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
loading period. Duncan (1996) recommends using the dimensionless time factor T from consolida-
tion theory to estimate the degree of drainage that can occur during construction or loading using
the relationship:
T = c
v
t / D
2
(6-32)
where:
c
v
= coefcient of consolidation (length
2
/time)
t = time
D = drainage path (length)
If T > 3, the material can be treated as drained. If T < 0.01, the material can be treated as undrained.
If 0.01 < T < 3, both drained and undrained conditions should be evaluated. Duncan (1996) suggests
that, if the data needed to calculate T are not available, soils with hydraulic conductivity k > 100 feet/
year can be considered drained and soils with k < 0.1 foot/year can be considered undrained.
Undrained conditions should be analyzed in terms of total stress in order to avoid having to rely on
estimated, and sometimes unreliable, pore pressures for undrained loading conditions. Undrained
strength can be determined using in-situ tests (e.g., vane shear), UU triaxial tests, or CU tests in con-
junction with a strength normalizing procedure such as the SHANSEP (stress history and normalized
soil engineering parameters) procedure (Ladd and Foot, 1994). For multi-stage construction (e.g.,
upstream pushouts), the undrained strength can be estimated using CU triaxial test results together
with values of consolidation pressure determined from consolidation analyses (Ladd, 1991).
Drained conditions can be analyzed in terms of efective stresses using c and from drained triaxial
or direct shear tests or from CU tests. Direct-shear or CU tests are more ofen used when testing clays,
TABLE 6.54 SHEAR STRENGTHS, WATER PRESSURES, AND UNIT WEIGHTS
FOR SLOPE-STABILITY ANALYSES
Soil Type Parameter
Condition
End-of-Construction Multi-stage Loading
(1)
Long-Term
All soils
External Water
Pressures
Include Include Include
All soils Unit weights Total Total Total
Free-draining Shear Strength
Effective stress c
and
Effective stress c
and
Effective stress c
and
Free-draining
Internal Pore
Pressures
u from steady-state
seepage analyses
u from steady state
seepage analyses
u from steady state
seepage analyses
Impermeable Shear Strength
Total stress c and
from in-situ or UU or
CU lab tests
Total stress c and
from in-situ or UU or
CU lab tests
Effective stress c
and
Impermeable
Internal Pore
Pressures
No internal pore
pressures, set u = 0
in computer input
No internal pore
pressures, set u = 0
in computer input
u from steady state
seepage analyses
Note: 1. Multi-stage loading includes rapid drawdown, staged construction, and any other condition where a
period of consolidation under one set of loads is followed by a change in load under undrained conditions.
(DUNCAN, 1996)
< PREVIOUS VIEW
6-178
Chapter 6
MAY 2009
because the time required for testing is shorter than for conducting CD tests. Values of c and from
CU tests have been found to be nearly identical to values obtained from drained tests. Values of for
natural deposits of cohesionless soils are usually estimated using correlations with SPT or CPT data.
In performing embankment stability analyses, the strength behavior of coal refuse must be carefully
evaluated. There is evidence that the failure strength envelope of coarse refuse has considerable cur-
vature in the stress range associated with high embankments due to crushing of the refuse particles.
Thus, the cohesion and friction angle of the coal refuse may vary depending upon location in the
embankment. Application of a bi-linear failure model may be appropriate for such cases.
TABLE 6.55 GUIDELINE FRICTION VALUES AND EFFICIENCIES FOR VARIOUS
GEOSYNTHETIC AND SOIL COMBINATIONS
Soil-to-Geomembrane Friction Angle
Geomembrane
Soil Types
Concrete Sand
( = 30)
Ottawa Sand
( = 28)
Mica Schist Sand
( = 26)
PVC
Rough 27 (0.88)
(1)
25 (0.96)
Smooth 25 (0.81) 21 (0.79)
CSPE 25 (0.81) 21 (0.72) 23 (0.87)
HDPE
18 (0.56) 18 (0.61) 17 (0.63)
Geomembrane-to-Geotextile Friction Angle
Geotextile
Geomembrane
PVC
CSPE HDPE
Rough Smooth
Nonwoven, Needle-Punched
23 21 15 8
Nonwoven, Melt-Bonded
20 18 21 11
Woven, Monoflament
11 10 9 6
Woven, Slit Film
28 24 13 10
Soil-to-Geotextile Friction Angle
Geotextile
Soil Types
Concrete Sand
( = 30)
Ottawa Sand
( = 28)
Mica Schist Sand
( = 26)
Nonwoven, Needle-Punched
30 (1.00) 26 (0.92) 25 (0.96)
Nonwoven, Melt-Bonded
26 (0.84)
Woven, Monoflament
26 (0.84)
Woven, Slit Film
24 (0.77) 24 (0.84) 23 (0.87)
Note: 1. Effciency values in parentheses are based on the relationship E = tan
/
tan
(ADAPTED FROM KOERNER, 2006)
< PREVIOUS VIEW
6-179
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
For sites where impoundments are lined with geomembranes or geosynthetic clay liners (GCLs) to
control seepage, the potential for slope failure between the liner and subgrade and between the liner
and soil cover placed over the liner may need to be evaluated. Table 6.55 provides guideline friction
angles that may be appropriate for preliminary design for various interfaces. However, fnal design
should be based on more refned published data and manufacturers information for the specifc geo-
synthetic materials under consideration, as well as laboratory interface testing between on-site soils
and selected geosynthetic materials to simulate site-specifc conditions.
When GCLs are used on slopes, their friction properties are important. Sodium bentonite, which is
ofen used in GCLs, is a clay with a saturated, drained residual internal angle of friction of approxi-
mately 6 to 9. However, signifcantly greater friction angles may be appropriate in GCLs that are
needle-punched or stitched. Manufacturers should be consulted for design data, as these data may
be product specifc.
An important part of slope stability analysis is determining the slip surface with the lowest FS. Most
computer programs that use an assumed circular failure surface systematically change the position
of the center of the circle and the length of the radius to fnd the most critical (lowest FS) circle. For
more complex geometries typical of most real-world situations, local minima may exist. Therefore,
multiple searches should be conducted using multiple starting points and search strategies so that
the overall minimum value of FS is determined. The results of slope stability analyses should be care-
fully examined to verify that the upstream and downstream limits of the search are not so restric-
tive as to exclude potentially critical failure surfaces. Locating a critical noncircular surface is more
complex, and a variety of approaches have been developed. Methods such as random generation
of kinematically admissible slip surfaces, coupled dynamic programming minimization techniques,
and optimization have been used successfully to model slopes that do not have extremely compli-
cated geometries. Regardless of the computational procedure used, tests of reasonableness should be
applied to the results, and multiple searches should be performed to be certain that the critical slip
surface has been located. Failure surfaces through weaker embankment or foundation layers should
always be considered.
Most slope stability problems can be modeled in two dimensions because the geometry of a slope is
typically relatively constant along its length. However, some slopes are: (1) curved in plan or con-
tain corners (e.g., some diked embankments), (2) subjected to loads of limited extent at the top, or
(3) constrained by physical boundaries such as a dam in a narrow-walled valley. For these situa-
tions, consideration may be given to conducting a three-dimensional (3D) limit-equilibrium analysis.
Duncan (1996) reports that the factor of safety for 3D analysis is greater than the factor of safety for
2D analysis (i.e., FS
3D
> FS
2D
) provided that: (1) FS
2D
is calculated for the most critical two-dimen-
sional cross section of the slope and (2) the procedure used for 2D limit-equilibrium analysis satisfes
all conditions of force and moment equilibrium. If 2D analyses of refuse embankment stability meet
these criteria, a 3D analysis is not generally warranted.
Some limit-equilibrium computer programs include features that permit a probabilistic analysis of
slope stability. Some common characteristics of these programs include:
Simulation techniques (e.g., Monte Carlo) that allow the program to repeatedly
sample values from probability distributions of the uncertain variables
Modeling of input parameters as random variables (e.g., material properties,
phreatic surface location, seismic load coefcient)
Defning the probability density function of the random variables in terms of
statistical distributions commonly used in geotechnics (e.g., normal, exponential,
lognormal)
< PREVIOUS VIEW
6-180
Chapter 6
MAY 2009
Using truncated distributions to defne maximum and/or minimum values
Defning correlation coefcients between correlated values (e.g., c and )
Presentation of results in a variety of forms (e.g., histograms, cumulative plots,
scater plots)
Thus, probabilistic slope stability analysis accounts for variability and uncertainty associated with
traditional limit-equilibrium methods. Figure 6.55 illustrates the results of a stability analysis of a
cohesive soil slope using Spencers method. The fgure shows a histogram of frequency distribution
of the FS computed for 5,000 random analyses. The red bars at the lef of the distribution represent
analyses where FS is less than 1. From the histogram, the mean FS = 1.072, and the maximum and
minimum FS are 1.298 and 0.860, respectively. However, Figure 6.55 illustrates another advantage of
probabilistic analysis. The ratio of the 658 analyses where FS is less than 1 to the 5,000 total analyses
is the probability of failure p
f
. For this analysis, p
f
= 13.2 percent, which provides a measure of the
potential for failure separate from the factor of safety. For this case, the analysis shows that a low
average FS results in a high probability of failure.
Note that no value of p
f
is recommended for design of embankment dams at the time of publica-
tion of this Manual, although proposed risk evaluation criteria and guidelines for signifcant- and
high-hazard-potential dams are available (Von Thun, 1996). El-Ramly et al. (2003) present a proba-
bilistic stability analysis of a tailings dike along with a general spread sheet model, and they note
that computed values of pf for existing dams demonstrating satisfactory performance may not meet
recommended values. However, a probabilistic analysis is useful in understanding the contribution
of parameters afecting stability and in comparing conditions and confgurations for establishing reli-
able design. Probabilistic acceptance criteria will become established as more analyses are conducted
and results published for both failed and satisfactorily performing slopes.
FIGURE 6.55 HISTOGRAM OF FACTOR OF SAFETY DISTRIBUTION
0.9 1.0 1.1 1.2
0
1
2
3
4
5
6
7
SPENCER FACTOR OF SAFETY
R
E
L
A
T
I
V
E
F
R
E
Q
U
E
N
C
Y
FIGURE 6.55 HISTOGRAM OF FACTOR OF SAFETY DISTRIBUTION
FIGURE 6.55 HISTOGRAM OF FACTOR OF SAFETY DISTRIBUTION
< PREVIOUS VIEW
6-181
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
Additional information relative to probabilistic methods for the stability analysis of slopes is pro-
vided in Hoek (2000) and Baecher and Christian (2003).
6.6.4.2.2 Stability/Deformation Analysis Using Finite Element Methods
Slope stability analyses are traditionally performed using classical techniques such as the method of
slices. These approaches are based on the assumption of rigid-plastic stress-strain behavior with no
deformation occurring prior to failure. However, if elastic deformations occur prior to failure, the
fnite element method can be used to solve the elastic diferential equations, with failure defned by
placing a limit on stresses using the Mohr-Coulomb (typical) or other failure criterion. The stress-
strain behavior of the embankment material should be modeled using the simplest representation
possible that is appropriate for the problem analyzed (Duncan, 1996). However, while simple linear
elastic, multi-linear elastic or hyperbolic models may be appropriate for analyzing stress states well
prior to failure, more complex stress-strain models (e.g., elasto-plastic and elasto-viscoplastic) are
required to analyze slope behavior near failure.
The elasto-plastic and elasto-viscoplastic FE approach to slope stability analysis ofers the following
advantages over traditional methods (Grifths and Lane, 1999):
No a priori assumptions are needed relative to the shape or location of the failure
surface. Failure occurs naturally through zones in the soil mass where the shear
strength is unable to support the gravity-induced shear stresses.
Because there is no concept of slices in the fnite element approach, there is no need
for assumptions about slice side forces, and the fnite element method preserves
global and local equilibrium until failure is reached.
The fnite element method can indicate progressive failure up to and including overall
shear failure.
Slope stability analysis using the fnite element method requires the following steps:
Gravity loads are applied to the slope.
An elastic analysis is performed to compute stresses.
Stresses in each element are compared with the Mohr-Coulomb failure criterion.
If stresses exceed the Mohr-Coulomb criterion, they are redistributed to neighboring
elements that still have reserve strength.
Slope failure occurs if stress redistribution cannot be accomplished to satisfy the
Mohr-Coulomb criterion and global equilibrium. Failure is indicated by signifcantly
increased nodal displacements.
Finite element analyses are not commonly used for the design of refuse embankments, but
have application for the analysis of embankments where unexpectedly large deformations are
observed or where unusually sof foundations may lead to unacceptable plastic deformations or
diferential setlements. Additionally, fnite element analyses may prove to be a useful tool for
evaluation of the efects of deformation on embankment stress and for the evaluation of mine
subsidence efects.
6.6.4.3 Acceptable Factors of Safety
Selection of an acceptable factor of safety for the analysis or design of an embankment slope depends
on the degree of uncertainty in calculating FS and the hazards or consequences should a slope fail.
The calculation of the factor of safety for a coal refuse disposal facility embankment involves evalua-
tion of many factors, including:
< PREVIOUS VIEW
6-182
Chapter 6
MAY 2009
Uncertainties about the type and extent of sampling and testing to determine
material strengths
The variation of materials in the embankment and impoundment
Uncertainties of embankment geometry
The type of embankment (e.g., new or existing, impounding or non-impounding,
constructed by upstream method or downstream method)
The level of uncertainty concerning the location of the phreatic level and related
excess pore-water pressures
Embankment location (e.g., in a V-shaped valley, high seismic region)
The accuracy of the stability analysis method(s) used
Potential for future loads
Potential for future changes in disposal operations and practices
The hazards or consequences associated with refuse embankment failure include:
Potential for loss of life and/or property damage on site
Potential for loss of life and/or property damage of site
Economic cost to mining operations if disposal operations are lost
Economic cost associated with restoring or safely abandoning disposal operations
Economic cost of restoring environmental damage
The magnitude of the factor of safety to be used for coal refuse disposal facility design should be
determined by the designer considering state and federal regulatory requirements and the complete-
ness and accuracy of designers knowledge of the uncertainties and conditions identifed previously.
If these uncertainties and conditions are well defned, or if conservative assumptions have been made,
a lower FS may be acceptable. Conversely, when many assumptions are made relative to forces and
material strengths, a higher FS is appropriate.
In selecting an acceptable FS for the stability of embankment slopes for coal refuse disposal facility
design, the practices of major engineering organizations, both governmental and private, should
be considered. Most earth dams in the U.S. are designed based upon extensive laboratory test
information and clearly identifed loads and geometry. For these structures, a FS 1.5 is generally
considered to be acceptable for permanent or sustained loading conditions. For special circum-
stances, such as sof embankment foundations, a higher FS is sometimes adopted to limit founda-
tion or embankment deformations. For temporary loading conditions during construction, a lower
FS is ofen acceptable. For transient loads, such as earthquakes, a FS of 1.2 is generally accept-
able depending on the design earthquake magnitude, as discussed in Chapter 7. Minimizing the
volume of material used for construction is not usually a design goal for refuse embankments as it
is for earthen dams; thus, it may be possible to achieve somewhat higher factors of safety for refuse
embankments without introducing special materials or operations, and the designer may be able
to accommodate concerns about uncertainty of engineering properties or loads.
Table 6.56 presents recommended minimum factors of safety for the design of coal refuse embank-
ments based on values adopted by the cited federal agencies (USSD, 2007). These values are
generally consistent with factors of safety cited in the MSHA Impoundment Inspection and Plan
Review Handbook (MSHA, 2007) and have been adopted by many state regulatory agencies.
The values of FS provided in Table 6.56 apply if the following conditions are met:
< PREVIOUS VIEW
6-183
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
TABLE 6.56 RECOMMENDED MINIMUM FACTORS OF SAFETY FOR DESIGN
OF COAL REFUSE EMBANKMENTS
Condition Design Basis
Factor of
Safety
Source
(1)
Long-term stability
analysis with maximum
storage pool
Design based on shear strength measured in laboratory and/
or feld testing program refecting long-term site develop ment
with steady-state seepage and maximum storage (operational)
pool.
1.5
USACE,
USBR,
NRCS,
FERC
Long-term stability
analysis with maximum
surcharge pool
Design based on shear strength measured in laboratory and/
or feld testing program refecting long-term site develop ment
with steady-state seepage and maximum surcharge (design
storm) pool.
1.4
USACE,
FERC
Intermediate-stage
static analysis with
maximum storage pool
Design based on shear strength measured in laboratory and/
or feld testing program refecting intermediate-stage critical
confgurations using long-term analysis with steady-state
seepage and maximum storage (operational) pool.
1.5
Intermediate-stage
static analysis with
maximum surcharge
pool
Design based on shear strength measured in laboratory and/
or feld testing program refecting intermediate-stage critical
confgurations using long-term analysis with steady-state
seepage and maximum surcharge (design storm) pool.
1.4
Short-term, end-of-
construction, undrained
static analysis
Design based on intermediate- or long-term confgurations
and supported by shear strength measured in laboratory and/
or feld testing program using undrained strength parameters,
as appropriate.
1.3
USACE,
USBR,
FERC, TVA
Rapid-drawdown
analysis with maximum
storage pool
Design based on intermediate or long-term confgurations and
supported by shear strength measured in laboratory and/or
feld testing program using drained or undrained analysis.
1.3
USACE,
USBR
Seismic analysis Discussed in Chapter 7.
Note: 1. From USSD (2007).
The critical failure surface has been determined from stability analyses
based on systematic searches and evaluation of defned planes of weakness.
Stability analysis parameters are, with reasonable certainty, known to be
representative of the actual conditions that will exist in the embankment.
Sufcient control will be provided during construction to verify that mate-
rials placed within the embankment conform to the standards assumed or
required by the disposal facility design.
When pore-water pressure in an embankment and its foundation is a
signifcant factor in the stability of the embankment, piezometers are
installed and monitored, and the observed data are compared to design
assumptions.
If an existing coal refuse disposal embankment is found to have a low FS, it may not be possible to
modify the site sufciently within a short time period to satisfy minimum FS criteria. In such cases, a
slight, temporary reduction in the FS may be acceptable provided that:
Monitoring of pore pressures in the embankment and movements of the embank-
ment surface is performed on a scheduled basis.
< PREVIOUS VIEW
6-184
Chapter 6
MAY 2009
A plan to improve the FS is developed and implemented (e.g., the impoundment
level is lowered, the embankment upstream face is sealed to reduce seepage, pore
pressures are measured, or other steps to reduce the impounding capacity or to
implement abandonment are initiated).
When test data for embankment and foundation materials are limited or available test data are incon-
sistent, either conservative values of shear strength and pore-water pressure should be used in the
stability analysis or an increased FS should be used in the embankment design.
6.6.4.4 Stability Control Measures
The most efective measures for stabilizing coal refuse embankments and other slopes are: (1) drain-
age control and (2) butress flls. Drainage control is probably the most frequently used and ofen the
most efective measure because slope failures are very ofen the result of increases in groundwater
level, phreatic surface level or pore pressures. Also, drainage control is ofen the least expensive and
most easily implemented of the options that are typically available. The types of drainage control
measures (Section 6.6.2.3) that can be employed include:
Surface control using ditches designed with a gradient and lining that limits infltra-
tion into the slope and conveys surface fows to a drainage outlet.
Lowering of the impoundment pool level, if feasible, and consideration of partial
liner systems placed on the upstream slope to limit seepage.
Occasional movement of the slurry discharge point so that fnes are distributed along
the upstream embankment slope, limiting seepage and pushing the free water fur-
ther away from the embankment.
Horizontal drains drilled at an upslope gradient into the downstream face of the
embankment to intercept water up to 400 feet from the face. As discussed in Section
6.6.3.3, prefabricated drains have been used to control pore pressures and improve
consolidation of impoundment fnes.
Relief wells to intercept artesian water pressure at the downstream toe of the
embankment. The wells may be furnished with pumps to convey fow to a drainage
outlet.
Trench drains extending below the toe of the downstream embankment slope or into
a bench on the slope. Flow is conveyed by gravity and typically in pipes to a drain-
age outlet.
Finger drains excavated perpendicular to and typically at a shallow depth into the
downstream face of the embankment to intercept water fow at a shallow depth.
Blanket drains placed at the surface of downstream embankment slopes where seep-
age or piping is occurring. Blanket drains, especially when incorporated into a but-
tress, also add weight to help increase the efective stress.
Structural butress and gravity berm flls constructed at the toe of an embankment also serve to
increase stability. A structural butress fll constructed with well-compacted, high-strength material
improves stability by providing both strength and weight. A gravity berm fll using uncompacted
material improves stability by providing weight to reduce shear stresses at the toe of the embank-
ment slope. The efectiveness of either type fll can be improved by placing a layer of free-draining
material between the embankment face and the fll to convey water draining from the face of the
embankment slope to a drainage outlet.
Other types of stabilization measures such as structural and ground improvement options may be
considered, but these choices typically require greater time to implement and are more expensive
and less efective than drainage or toe butress stabilization measures.
< PREVIOUS VIEW
6-185
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
6.6.4.5 Stability Measurements
Parameters that infuence the slope stability factors of safety should be monitored during and afer
each stage of construction by instrumentation to provide a basis for checking the design factor of
safety so that modifcations to the design or to construction procedures can be made, if needed.
Depending on site conditions and project requirements, embankment performance monitoring
instrumentation may include:
Surface monuments for monitoring vertical and horizontal displacements
Inclinometers for monitoring lateral displacements at depth
Piezometers for monitoring piezometric head
In addition, monitoring of density test results is important to stability because it provides an indirect
method to monitor the strength of the compacted fll. Thus, the minimum compaction requirement
serves to ensure that the shear strength of the compacted fll is consistent with the strengths obtained
from laboratory testing and used in the stability analysis.
An extensive discussion of instrumentation for coal refuse disposal facilities is provided in Chapter 13.
6.6.5 Rock Excavations
Excavations into rock at coal refuse disposal facilities are usually associated with: (1) spillway chan-
nels, (2) diversion ditches, and (3) haul roads. Less frequently, rock excavations are associated with
decant structure installation and obtaining borrow material.
The degree of assurance that an excavated rock slope will remain stable and has a sufcient factor of
safety for each of these purposes will depend on facility design requirements. This section provides a
discussion of the stability of rock excavations, methods for minimizing the potential for a failure, and
procedures for improving stability.
6.6.5.1 Conditions Requiring Stability Analysis
Major factors that must be considered when evaluating the stability of rock excavations include:
Existing ground surface slope
Overburden soil thickness, type and quality
Bedrock surface conditions
Rock type, quality and confguration (particularly the orientation of discontinuities)
Groundwater conditions
6.6.5.1.1 Existing Ground Surface Slope
The steepness of the existing ground surface is an important factor in both the cost of excavation and
the resulting stability. Figure 6.56 provides a guide for estimating both the vertical and horizontal
extent of rock excavations as a function of existing slope, cut width, and cut slope. In steep areas
where existing slopes are only marginally stable, deep excavations are very expensive because of the
volume of material to be excavated and protection requirements.
6.6.5.1.2 Overburden Soil Thickness, Type and Quality
Usually the top portion of an excavation into rock extends through overlying soils. The need for a
shallower slope and erosion protection in the overburden soil portion of the excavation must be con-
sidered in evaluating the limits of excavation and related costs.
< PREVIOUS VIEW
6-186
Chapter 6
MAY 2009
W L
H 1
S
E
1
S
C
40
30
20
10
0
10 0 10 20 30 40
H
E
I
G
H
T
,
H
(
F
E
E
T
)
HORIZONTAL EXTENT, L (FEET)
0
.
5
:
1
0
.
7
5
:
1
1
:
1
1
.
2
5
:
1
1
.
5
:
1
2
:
1
0
.
5
:
1
0
.
7
5
:
1
1
:
1
1
.
2
5
:
1
1
.
5
:
1
2
:
1
BASE WIDTH OF CUT
W
FIGURE 6.56 HORIZONTAL AND VERTICAL EXTENT OF
CUTS FOR BASE WIDTH OF 10 FEET
LEGEND
EXISTING SLOPE (S
E
)
SLOPE OF CUT (S
C
)
(1) S
E
> S
C
(3) H =
L
S
C
(2) L = W
S
C
S
E
- S
C
FOR W = 10 FEET, MULTIPLY
THE VALUES OF L AND H
OBTAINED FROM THE GRAPH
BY COEFFICIENT C = W/10.
FIGURE 6.56 HORIZONTAL AND VERTICAL EXTENT OF CUTS FOR BASE WIDTH OF 10 FEET
6.6.5.1.3 Bedrock Surface Conditions
As shown in Figure 6.57, the bedrock surface can be difcult to defne because of weathering, downhill
slope creep and loosening from stress relief. Except where rock is very massive, conditions normally
change with depth from: (1) overburden soil consisting of colluvial and residual material to (2) sof,
heavily weathered rock and then to (3) sound and competent rock. Heavily weathered, decomposed
rock should be treated as soil to a depth determined by an experienced geologist or engineer. Particu-
lar caution is required for sof rocks (e.g., shale, siltstone, claystone and mudstone) that weather very
rapidly when exposed to air, rainfall and freezing conditions.
FIGURE 6.56 HORIZONTAL AND VERTICAL EXTENT OF CUTS FOR BASE WIDTH OF 10 FEET
< PREVIOUS VIEW
6-187
Geotechnical Engineering, Material Testing, Engineering Analysis and Design
MAY 2009
SANDSTONE
SANDSTONE
SHALE
WEATHERED ROCK RESIDUAL SOIL
COLLUVIUM
GROUND SURFACE
NOTE: INTERFACES CAN BE
DIFFICULT TO DETERMINE.
FIGURE 6.57 SCHEMATIC GEOLOGICAL PROFILE OF A BEDROCK SLOPE
FIGURE 6.57 SCHEMATIC GEOLOGICAL PROFILE OF A BEDROCK SLOPE
6.6.5.1.4 Rock Type, Quality and Confguration
Generally the design of a cut into competent rock is technically simple and requires only applica-
tion of the designers judgment or general experience. However, when rock conditions are such
that analyses are required for demonstrating adequate stability, detailed evaluation is frequently
more complex than the evaluation of the stability of a soil slope. A rock mass may consist of: (1) a
layered system of individual beds of diferent type rocks with variable bed thickness and strength
and (2) discontinuities such as bedding planes, joints, fractures and faults that have much lower
strengths than the intact rock. These discontinuities control rock slope stability, but can be difcult
to defne for purposes of analysis. The types and characteristics of rock most ofen found in coal
mining areas are:
Sof Rocks Shale, Siltstone, Claystone and Mudstone Rapid weathering of these
types of rocks afects their strength, which eventually decreases to that of soil. Where
these rocks are exposed, it must be expected that weathering will occur during the
useful life of the facility. When exposed sof rock underlies massive, more competent
rock, particular care must be taken to avoid instability as the sof rock weathers and
provides less support. Also, erosion of sof rocks can be a problem, particularly when
they are thinly-bedded and exposed to surface water fows for long periods.
Excavation of sof rocks can usually be accomplished with dozers, shovels, scrapers
or backhoes. Even when competent, these rocks can be ripped without blasting using
a bulldozer.
Harder Rocks Limestone and Sandstone These types of rocks are more resistant
to weathering. Strength is usually high if defects such as fractures or interbedded
units of sofer rocks do not occur at critical locations. Defects and weak cementa-
tion can also adversely afect the erosion resistance of these rocks, which is other-
wise good to excellent.
The exploration and evaluation of rock defects is very important. The characteristics of
rock defects reported as part of an exploration program generally include type, qual-
ity, thickness, strike and dip, continuity, extent, frequency, and relative strength. The
orientation of rock defects relative to the excavation is critical. For example, Figure 6.58
illustrates two conditions where orientation of bedding planes dipping into a cut and
daylighted by the excavation could cause bedrock sliding along these planes. The
extent of a potential slide could be small (Figure 6.58a) or large (Figure 6.58b) depend-
FIGURE 6.57 SCHEMATIC GEOLOGICAL PROFILE OF A BEDROCK SLOPE
< PREVIOUS VIEW
6-188
Chapter 6
MAY 2009
FIGURE 6.58 UNFAVORABLE ORIENTATION OF BEDDING PLANES
(7-1)
Methods for evaluating CSR, CRR, MSF, K
and K
av
= average cyclic shear stress induced by the earthquake = 0.65
max
(force/length
2
)
vo
= efective vertical overburden stress (force/length
2
)
As discussed in the following text, computing
av
in a zone of material requires a design earthquake
PGA or a design-earthquake, acceleration time-history.
Three methods for computing CSR that are acceptable for various situations are:
1. CSR computation method 1 The preferred method is to compute CSR by using 1D
or 2D numerical site response sofware, such as SHAKE or QUAD4, that compute
the variation in maximum shear stress
max
with depth. The value of
av
at the depth
of interest should be taken from the site response analyses as 0.65
max
at that depth.
Using a one-dimensional (1D) or two-dimensional (2D) numerical site response
analysis requires, as input to the analysis, an acceleration time-history for the design
earthquake as well as dynamic properties (shear modulus and damping) for each
material type. This method is appropriate for all sites.
Commentary: Important material properties used in performing the analysis are the shear
modulus and damping. They can be estimated using index and classifcation data, or site-
specifc testing such as crosshole testing can be performed to measure shear wave velocity (to
obtain the maximum shear modulus G
max
). Laboratory testing including resonant-column
testing and cyclic-triaxial testing can also be utilized. The resonant-column test measures
dynamic soil properties at a lower range of shear strain than the cyclic-triaxial test. Dynamic
properties of coal refuse materials are the subject of research in Kalinski and Phillips (2008)
and Zeng and Goble (2008). Whether 1D or 2D site response analyses should be used depends
on the specifc conditions for which the triggering analysis is being performed (locations of
potential slope stability failure surfaces, variability of materials, geometric complexity, and
distance from the suspect zone to the exposed slope). For analysis of upstream failures of coal
refuse impoundments, the potential failure surfaces are ofen relatively far from the down-
stream face, such that the simplifcations associated with 1D analyses are acceptable. For
analysis of downstream failures of coal refuse impoundments with wide coarse refuse stages,
the potential failure surfaces generally pass through the coarse refuse stages and possibly
thin zones of fne refuse. So even though the ground surface is sloping, this geometry can be
reasonably modeled with a series of 1D analyses. For analysis of downstream failures of coal
refuse impoundments with narrow coarse refuse stages, the potential failure surfaces pass
through highly variable materials, and the sloping ground efects (downstream face) are more
pronounced. Therefore 2D analyses should be used to model these more complex conditions.
2. CSR computation method 2 (for shallow, level-ground sites) For conditions of
nearly level ground, uniform deposits, and depths up to 75 feet, CSR can be comput-
ed from the following equation (Youd et al., 2001):
< PREVIOUS VIEW
7-26
Chapter 7
MAY 2009
CSR = (
av
/
vo
) = 0.65 ( a
max
/g) (
vo
/
vo
) r
d
(7-3)
where:
a
max
= maximum horizontal acceleration at the ground surface
(e.g., top of embankment) (mass length/ time
2
)
g = acceleration of gravity (mass length/ time
2
)
vo
= total overburden stress (force/ length
2
)
vo
= efective vertical overburden stress (force/ length
2
)
r
d
= stress reduction coefcient (1.0 at the ground surface,
decreasing with depth) (dimensionless)
Commentary:
av
= 0.65
max
max
= peak seismic shear stress at any depth = (
vo
) (a
max
/g) (r
d
)
max
/
vo
= (a
max
/g) (r
d
)
max
is the product of the total overburden stress at the depth in question
vo
times the peak
acceleration of the overburden mass above the depth in question (a
max
/g)(r
d
). This is equiva-
lent to saying that seismic shear force equals mass times seismic acceleration or seismic shear
stress (force per square foot) equals total vertical stress (mass per square foot) times seismic
acceleration.
The coefcient r
d
accounts for the deformability of the overburden mass, which results in the
seismic shear stress at depth being lower than the product of the total overburden stress
vo
times the peak acceleration at the ground surface a
max
. In other words, r
d
represents the fact that
the peak acceleration of the overburden mass (a
max
/g) (r
d
) at depth is lower than the maximum
acceleration at the ground surface, and thus there is a reduction of the ratio of peak seismic
shear stress to total overburden stress (
max
/
vo
) with depth. Seed et al. (2003) present refne-
ments for values of r
d
for specifc cases, or Youd et al. (2001) may be used directly.
The computation of total overburden stress
vo
should not include the weight of any free water
that may exist above the ground surface. It is a relatively common error to include the weight
of free water since it is part of the total overburden. However, it cannot be accelerated by the
ground below, and thus it does not count as part of the mass being shaken.
3. CSR computation method 3 For depths greater than about 75 feet or for irregular
and sloping ground conditions (i.e., for most coal refuse embankments), the follow-
ing simplifed method can, in certain cases, be used instead of Method 1. Method 3
can be used for cases where potential failure surfaces pass through relatively uni-
form materials, and the potential failure surfaces are relatively far from steep slopes.
Examples are:
Upstream failures of coal refuse impoundments where the potential failure
surfaces pass through reasonably uniform materials (no abrupt changes from
loose or sof material to dense or hard material), and the failure surfaces are
relatively far from the steep downstream slope.
Downstream failures of coal refuse impoundments with wide coarse refuse
zones near the downstream slope, where the potential failure surfaces also
pass through reasonably uniform materials.
< PREVIOUS VIEW
7-27
Seismic Design: Stability and Deformation Analyses
MAY 2009
FIGURE 7.3 COMPARISON OF PEAK BASE AND
CREST TRANSVERSE ACCELERATIONS MEASURED AT
EARTH DAMS
Method 3 uses the same equation for CSR as Method 2. However, the stress reduction coefcient r
d
used in that equation is only appropriate for level ground. Therefore, above the depth of interest,
replace the product (a
max
r
d
) with the maximum acceleration of the overburden mass k
max
. The value
of k
max
can be estimated from published relationships of the depth-dependent variation in maximum
acceleration ratio (k
max
/a
max
versus y/h, where y is the depth below the top of the embankment and h
is the embankment height). The average of all data line from Figure 7.4 can be used.
The maximum acceleration at the ground surface (top of embankment) a
max
will normally be higher
than the maximum acceleration at the ground surface at the base of the embankment (PGA). The
value of a
max
can be estimated from the PGA of the design earthquake at the base of the embankment
using Figure 7.3.
Figures 7.3 and 7.4 were developed for conventional dams with an approximately triangular shape.
Therefore, the results for Method 3 should be compared with the results of Method 2. If the results
are signifcantly diferent, the more conservative value should be used, or an analysis using Method 1
should be performed.
PEAK TRANSVERSE BASE ACCELERATION, g
P
E
A
K
T
R
A
N
S
V
E
R
S
E
C
R
E
S
T
A
C
C
E
L
E
R
A
T
I
O
N
,
g
FIGURE 7.3 COMPARISON OF PEAK BASE AND CREST TRANSVERSE
ACCELERATIONS MEASURED AT EARTH DAMS
(HARDER, 1991)
0 0.1 0.2 0.3 0.4 0.5
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
1989 LOMA PRIETA
EARTHQUAKE
PREVIOUS EARTHQUAKES
FIGURE 7.3 COMPARISON OF PEAK BASE AND CREST TRANSVERSE
ACCELERATIONS MEASURED AT EARTH DAMS
< PREVIOUS VIEW
7-28
Chapter 7
MAY 2009
Commentary: It should be noted that the ratios of maximum acceleration at the base of the embankment to
maximum acceleration at the top of the embankment are diferent in Figures 7.3 and 7.4. That is because Figure
7.4 compares the maximum acceleration of the overburden mass above the base of the embankment to the maxi-
mum acceleration at the top of the embankment, while Figure 7.3 compares the maximum acceleration at the
base of the embankment to the maximum acceleration at the top of the embankment. In other words, Figure
7.4 considers that, because the overburden mass is not rigid, the accelerations at diferent depths within the
overburden mass will be diferent. Figure 7.4 looks at the average acceleration of the overburden mass at each
time increment and then picks the maximum value k
max
. In Figure 7.4, k
max
acts at the base of the overburden
mass, but represents the average acceleration of the mass above the base. In Figure 7.3, the peak transverse base
acceleration is the acceleration at the specifc depth corresponding to the botom of the embankment.
FIGURE 7.4 VARIATION OF MAXIMUM ACCELERATION RATIO WITH DEPTH OF SLIDING MASS
7.4.2.2.2 Evaluation of Cyclic Resistance Ratio (CRR)
The cyclic resistance ratio can be evaluated through the use of feld test data including: (1) SPT data,
(2) CPT data, and (3) shear-wave velocity (V
s
) measurements. Procedures for computing CRR are
described in Youd et al. (2001). It is good practice to obtain both SPT and CPT data. Shear-wave veloc-
ity data may also be helpful. A description of each of these approaches is provided in the following:
CRR determined from SPT Numerous plots have been published over the years
showing CRR as a function of SPT blowcount N
1,60
. The plot from Youd et al. (2001)
is shown in Figure 7.5. The plot shows zones where data from case histories suggest
FIGURE 7.4 VARIATION OF MAXIMUM ACCELERATION
RATIO WITH DEPTH OF SLIDING MASS
0.0 0.2 0.4 0.6 0.8 1.0
1.0
0.8
0.6
0.4
0.2
0.0
MAXIMUM ACCELERATION RATIO, k
max
/a
max
"SHEAR SLICE"
(RANGE FOR ALL DATA)
FINITE ELEMENT
METHOD
AVERAGE OF
ALL DATA
R
E
L
A
T
I
V
E
D
E
P
T
H
O
F
S
L
I
D
I
N
G
M
A
S
S
,
y
/
h
(MAKDISI AND SEED, 1978)
FIGURE 7.4 VARIATION OF MAXIMUM ACCELERATION RATIO WITH DEPTH OF SLIDING MASS
< PREVIOUS VIEW
7-29
Seismic Design: Stability and Deformation Analyses
MAY 2009
that signifcant pore-pressure increase will occur and zones where signifcant pore-
pressure increase is unlikely to occur. The plot includes curves developed for soils
with various percentages of fnes. The curve for less than 5 percent fnes is consid-
ered the basic penetration criterion for the simplifed method described in Youd et al.
(2001) and is thus referred to as the SPT clean sand base curve. Youd et al. (2001)
describes ways to adjust N
1,60
to a clean sand value N
1,60(CS)
. The curves are valid
only for magnitude 7.5 earthquakes. Scaling factors to adjust for diferent magnitude
earthquakes are discussed later.
Two corrections are normally made to the feld blow count N. The frst (C
N
) is made
to normalize N to an overburden pressure of 1 atmosphere (about 1 tsf or 100 kPa).
The second (C
E
) is made to normalize hammer energy to 60 percent of the energy
generated by a 140-lb weight free-falling 30 inches. Other corrections less commonly
made in practice include borehole diameter (C
B
), rod length (C
R
), and samplers with
and without liners (C
S
). Thus:
N
1,60
= N C
N
C
E
C
B
C
R
C
S
(7-4)
The correction factors are discussed in Youd et al. (2001), which notes that the correc-
tion for overburden (C
N
) becomes highly uncertain for overburden pressures greater
than 3 tsf.
The presence of gravel can interfere with the penetration of the SPT sampler and
lead to higher penetration resistances. One way to correct for the presence of gravel
in the SPT is to record the hammer blows for every inch of penetration (rather than
the standard 6 inches) as proposed by Poulos in the 1970s and described in USBR
(1989) and Seed et al. (2003).
CRR determined from CPT Robertson and Wride (1998) developed curves of CRR
as a function of dimensionless CPT tip resistance ratio q
t1N
, as shown in Figure
7.6. The fgure shows zones where data from case histories suggest that signif-
cant pore-pressure increase will occur and zones where signifcant pore-pressure
increase is unlikely to occur. The feld CPT tip resistance q
t
must be normalized to
an overburden pressure of 1 atmosphere (about 1 tsf or 100 kPa) to obtain q
t1N
and
must be corrected to an equivalent clean sand value q
t1N(CS)
. Equations and charts
for normalizing q
t
and correcting to an equivalent clean sand value are discussed in
Youd et al. (2001) and updated by Robertson (2004), which also includes a discus-
sion of using average CPT values in a layer versus using all measured values. An
additional correction for thin soil layers can be made, if applicable (Youd et al.,
2001). The references by Robertson and Wride (1998) and Youd et al. (2001) use
the uncorrected tip resistance q
c
whereas, more correctly, it should be q
t
. For CPT
in clean to silty sands, q
c
and q
t
are essentially the same and are ofen used inter-
changeably.
CPTs have the advantage that they are faster to perform and typically less expensive
per test than SPTs, especially in deep, sof or loose deposits, and CPTs provide a con-
tinuous, more reliable profle of penetration resistance. CPTs are particularly useful
for fne coal refuse. Physical samples of the material where CPTs are conducted
should be obtained for purposes of description and laboratory testing.
CRR determined from shear-wave velocity CRR criteria have been developed from
feld measurements of shear-wave velocity V
s
. Shear-wave velocities can be measured
< PREVIOUS VIEW
7-30
Chapter 7
MAY 2009
using the seismic CPT (SCPT), crosshole testing, or other downhole or surface wave
techniques. Crosshole testing is generally more accurate, but more expensive than
seismic CPT. Potential advantages of using the shear-wave velocity are: (1) it is possible
to obtain data in soils that are difcult to penetrate with CPT or SPT and (2) the shear
wave velocity is directly related to small-strain shear modulus, which is a parameter
required for estimating dynamic soil response. As discussed by Youd et al. (2001),
concerns with the use of V
s
include: (1) shear wave velocity measurements are made
at small strains, whereas pore-pressure buildup and strength loss are medium to high
strain phenomena and (2) thin, low shear-wave velocity layers may not be detected.
V
s
data, if obtained, should generally be used in conjunction with data from other
methods. The shear-wave velocity obtained from the feld should be normalized to a
reference overburden pressure of one atmosphere. A recommended chart of CRR as a
function of overburden-corrected shear wave velocity is provided in Youd et al. (2001).
FIGURE 7.5 SPT CLEAN-SAND-BASED CURVE FOR MAGNITUDE 7.5 EARTHQUAKES
WITH DATA FROM LIQUEFACTION CASE HISTORIES
CORRECTED BLOW COUNT, N1,60
C
Y
C
L
I
C
S
T
R
E
S
S
R
A
T
I
O
(
C
S
R
)
O
R
C
Y
C
L
I
C
R
E
S
I
S
T
A
N
C
E
R
A
T
I
O
(
C
R
R
)
FIGURE 7.5 SPT CLEAN-SAND BASE CURVE FOR MAGNITUDE 7.5 EARTHQUAKES
WITH DATA FROM LIQUEFACTION CASE HISTORIES
(ADAPTED FROM YOUD ET AL., 2001)
MODIFIED CHINESE CODE PROPOSAL
(CLAY CONTENT = 5%)
FINES CONTENT 5% >
LIQUEFACTION
MARGINAL
LIQUEFACTION
NO
LIQUEFACTION
PAN-AMERICA DATA
JAPANESE DATA
CHINESE DATA
30
20
30
31
27
10 13
10
10
50+
60
75
27
75
67
13
22
30
80
10
20
20
25
50 92
10
10
12
12
26
20
48
20
40
10
20
10
10
50+
80
60
27
50+
20
31
29
12
12
18
18
20
17
11
10
12
25
10
ADJUSTMENT
RECOMMENDED
BY WORKSHOP
SPT CLEAN SAND BASE CURVE
PERCENT FINES = 35 15 > 5
37
0 10 20 30 40 50
0
0.1
0.2
0.3
0.4
0.5
0.6
FIGURE 7.5 SPT CLEAN-SAND-BASED CURVE FOR MAGNITUDE 7.5 EARTHQUAKES
WITH DATA FROM LIQUEFACTION CASE HISTORIES
< PREVIOUS VIEW
7-31
Seismic Design: Stability and Deformation Analyses
MAY 2009
7.4.2.2.3 Correction Factors for Earthquake Magnitude, High Overburden, Sloping Ground
and Age of Deposit
The curves used to obtain CRR from SPT, CPT and V
s
only apply to magnitude 7.5 earthquakes.
Magnitude scaling factors (MSF) have been developed to adjust the curves to smaller or larger mag-
nitudes. The recommendations by Youd et al. (2001) represent a wide consensus among experts in
the feld and are preferred at this time unless project-specifc factors indicate that other recommenda-
tions should be used.
The simplifed method was developed based on case history data from gently sloping sites (low
static shear stress) and depths less than about 50 feet. Correction factors (K
and K
(7-5)
CORRECTED CPT TIP RESISTANCE, qc1N
C
Y
C
L
I
C
S
T
R
E
S
S
R
A
T
I
O
(
C
S
R
)
O
R
C
Y
C
L
I
C
R
E
S
I
S
T
A
N
C
E
R
A
T
I
O
(
C
R
R
)
FIELD PERFORMANCE
STARK & OLSON (1995)
SUZUKI ET AL.(1995b)
0 50 100 150 200 250 300
0
0.1
0.2
0.3
0.4
0.5
0.6
LIQUEFACTION
NO LIQUEFACTION
M = 7.5
CPT CLEAN SAND
BASE CURVE
NCEER (1996)
WORKSHOP
FIGURE 7.6 CALCULATION OF CRR FROM CPT DATA ALONG WITH EMPIRICAL
LIQUEFACTION DATA FROM COMPILED CASE HISTORIES
(YOUD ET AL., 2001)
0.25 < D
50
(mm) < 2.0
LIQ. NO LIQ.
FC (%) < 5
FIGURE 7.6 CALCULATION OF CRR FROM CPT DATA ALONG WITH EMPIRICAL
LIQUEFACTION DATA FROM COMPILED CASE HISTORIES
FIGURE 7.6 CALCULATION OF CRR FROM CPT DATA ALONG WITH EMPIRICAL
LIQUEFACTION DATA FROM COMPILED CASE HISTORIES
< PREVIOUS VIEW
7-32
Chapter 7
MAY 2009
Recommendations for K
and explain that there are no generally agreed upon recommendations. Seed et al. (2003)
recommend that the K
values proposed by Harder and Boulanger (1997) be used where the initial
vertical efective overburden pressure is less than 3 tsf. Until a consensus is reached, K
should be
taken as 1.0 unless site-specifc information (laboratory test data) indicates otherwise.
It has been noted that older deposits (both older reconstituted laboratory samples and geologically
older natural soil deposits) are more resistant to pore-pressure increase. However, verifed correction
factors have not been developed. Since it is conservative to ignore age corrections, age corrections are
not recommended at this time.
7.4.2.3 Strain-Based and Stress-Based Methods to Evaluate Triggering of Strength Loss
7.4.2.3.1 Clay-Like Material: Strain-Based Method for Triggering
Sof, clay-like materials may experience strength loss during an earthquake, but the strength loss is
generally not as severe as with sand-like materials (Thiers and Seed, 1968; Castro, 2003; Boulanger
and Idriss, 2004). This is because clay-like materials generally reach both peak undrained strength
and S
us
at much higher strains than sand-like materials, except for highly sensitive clays where the
strain to peak undrained strength can be small.
Therefore, the term triggering can be somewhat misleading for clay-like material. It is more cor-
rect to say that the post-earthquake strength for clays is a function of the shear strain that occurs
during the earthquake. The post-earthquake strength may be the peak undrained strength or it may
be a reduced strength. The post-earthquake strength could be as low as S
us
in highly sensitive clays,
but for most clay-like material the post-earthquake strength will be higher than S
us
. The reduced
strength, while higher than S
us
, may still be low enough to result in a fow slide.
There are three approaches to estimating the post-earthquake shear strength of clay-like material:
1. Conservatively assume that the post-earthquake strength is equal to S
us
, and esti-
mate S
us
from laboratory or feld testing, as discussed in Sections 7.4.3.2.2 and 7.4.3.3.
2. Estimate the post-earthquake strength as a function of earthquake-induced shear
strain, using the following procedure:
Estimate the accumulated shear strain at points along critical failure
surfaces in the embankment during the earthquake. The accumulated
shear strains can be estimated using Newmark-type analyses (Section
7.5.4).
Commentary: This step of this method requires that a deformation analysis
be performed as part of the evaluation of post-earthquake strength. A design
earthquake motion (time history of acceleration) and dynamic soil properties
(modulus and damping) are required for this step.
Obtain a relationship of post-earthquake strength as a function of total
strain based on laboratory testing or other correlations as discussed in
Section 7.4.3.2.2.
Use the computed total strains during the earthquake and the
relationship of post-earthquake strength versus strain during the
earthquake to estimate post-earthquake strength in each zone of
clay-like material.
< PREVIOUS VIEW
7-33
Seismic Design: Stability and Deformation Analyses
MAY 2009
3. Estimate the post-earthquake strength based on the earthquake-induced, cyclic shear
stress. This approach involves performing cyclic-undrained, shear-strength testing
(that models the earthquake loading) on undisturbed samples of the clay-like mate-
rial and then loading the samples monotonically to evaluate how much the cyclic
loading degraded the peak strength. This method is discussed in more detail in Sec-
tion 7.4.3.2.2.
7.4.2.3.2 Sand-Like Material: Strain-Based Method for Triggering
The method described in the following text is used for evaluating whether the design earthquake
will produce shear strains that are high enough to cause strength loss in a zone of sand-like material.
This method is only appropriate if the safety factor against triggering using the pore-pressure-based
method (CRR/CSR) is between 1.0 and 1.4. If the safety factor is higher than 1.4, then triggering may
be assumed to not occur. If the safety factor is less than 1.0, then triggering should be assumed. For
safety factors between 1.0 and 1.4, the pore-pressure-based method indicates that triggering is likely,
but this strain-based method may indicate that triggering will not occur.
As discussed in Section 7.4.1, if the design PGA is larger than 0.2g and the CSR in the loose zone is
greater than 0.15, then the triggering analysis described in the following text should not be performed,
and triggering of strength loss in the loose zone should simply be assumed. The post-earthquake
strength of the material in the loose zone should be assumed to be equal to S
us
. In evaluating potential
measures to improve the stability of a dam or embankment where the CSR in a zone exceeds 0.15, it
may be valuable to perform the triggering analysis to help assess whether alternative stabilization
schemes that would reduce the CSR might ultimately improve the dam or embankment stability.
The discussion that follows is based primarily on Castro (1994), which should be consulted for fur-
ther details. The steps in the method are:
1. Select one or more potential failure surfaces through the embankment. Estimate the
total seismically-induced shear strain along the critical failure surfaces in the em-
bankment during the earthquake. Failure surfaces that had the lowest safety factors
in the post-earthquake stability analysis performed when triggering of loose material
was assumed should be selected. The total shear strains include: (1) transient cyclic
strains, which can be estimated by a 1D or 2D site-response analysis using SHAKE or
QUAD4 and (2) accumulated strains, which can be estimated using Newmark-type
analyses (Section 7.5.4) or numerical modeling (Section 7.5.5). To be conservative, the
total seismically-induced shear strain should be computed by adding the maximum
transient cyclic strain to the total accumulated strain.
If numerical modeling is used to estimate accumulated strain, the model output will
include shear strains. If a Newmark-type analysis is used to estimate accumulated
strains, the analysis will provide displacements. To convert displacements to shear
strains, the thickness of the loose zone should be estimated, and the strains should be
assumed to be uniform across the full thickness of the zone. In other words, the shear
strain will be approximately equal to the displacement divided by the thickness of the
zone. A Newmark-type analysis can be used instead of numerical modeling only for
cases in which the loose zone is relatively thin compared to the overall failure mass.
Commentary: A design earthquake motion (time history of acceleration) is required for per-
forming the deformation analysis required for this step.
2. The triggering shear strain of sand-like material is approximately equal to the shear
strain at peak strength in undrained monotonic (non-cyclic) loading. However, for
< PREVIOUS VIEW
7-34
Chapter 7
MAY 2009
the triggering analysis, conservatively assume that the triggering shear strain is
equal to one-half the shear strain at peak strength in monotonic, strain-controlled
undrained laboratory strength tests. (Shear strain is equal to 1.5 times the axial strain
in triaxial tests.)
Perform undrained shear tests on samples consolidated anisotropically to stresses
corresponding to the static anisotropic stresses along the potential failure plane
through the embankment.
Undisturbed samples should normally not be used for this testing because of con-
cern for densifcation during sampling, handling, and laboratory consolidation to the
point that the samples are not as contractive in the laboratory as the in-situ mate-
rial. Remolded samples should generally be used, atempting to bracket (afer being
anisotropically consolidated in the lab) void ratios (more correctly relative densi-
ties) that refect feld conditions. Representative samples from the zone of interest
should be mixed to obtain a batch of soil for remolded testing. The samples that are
mixed should have similar grain size distribution. (As an example, clean sand mate-
rial should not be mixed with material containing more than about 20 percent fnes).
Remolded samples can be prepared by moist tamping in multiple layers or by wet
or dry pluviation. Pluviation may not be appropriate for silty samples because of the
potential for material segregation by particle size.
Commentary: The strain to peak strength and the triggering strain vary with the degree
of anisotropic consolidation. Therefore, to model feld conditions, anisotropic consolidation is
required for these tests. In contrast, for identical specimens consolidated to the same void ratio,
the steady-state (residual) strength S
us
measured in the laboratory will be the same for isotro-
pically- and anisotropically-consolidated specimens. Therefore, for convenience, S
us
testing is
normally performed on isotropically-consolidated specimens, as discussed in Section 7.4.3.2.3.
The recommendation to estimate the triggering shear strain as being equal to one-
half the shear strain at peak undrained strength in a monotonic test is conservative.
Being conservative at this step is appropriate because of: (1) the low shear strain
required to reach peak undrained strength indicates a high potential for progressive
failure and (2) the evaluation of seismically-induced shear strains is highly uncertain.
Instead of performing laboratory strength tests, the shear strain at peak undrained
strength could be estimated as a function of the ratio of vertical to horizontal consoli-
dation stress from published data for similar sand-like material (Castro 1994). The
triggering shear strain would then be assumed to be equal to one-half the estimated
shear strain at peak undrained strength. However, at this time, there is not adequate
published data available to rely on, so either: (1) site-specifc laboratory testing
should be performed or (2) the triggering shear strain of the sand-like material can
simply (conservatively) be assumed to be 0.25 percent. This recommended shear
strain is a lower bound for the data presented in Castro (1994) for consolidation
stress ratios typical of coal refuse impoundments.
3. If the total computed seismically-induced shear strain along the critical failure sur-
face is less than the shear strain needed to trigger strength loss, as estimated from:
(1) laboratory testing, (2) comparisons to published data on similar materials, or (3)
simply taken as 0.25 percent, then strength loss will not be triggered. Use the peak
undrained shear strength (or the drained strength, whichever is lower) in the stabil-
ity analyses. If the total seismically-induced shear strain is greater than the triggering
shear strain, use S
us
in the stability analyses.
< PREVIOUS VIEW
7-35
Seismic Design: Stability and Deformation Analyses
MAY 2009
7.4.2.3.3 Sand-Like Material: Stress-Based Method for Triggering
This method is used to evaluate whether the design earthquake will produce shear stresses in a zone
of sand-like material that are high enough to cause strength loss in that zone. The basic concept of the
method is that strength loss will be triggered by earthquake shaking if the sum of the static (gravity)
shear stresses along a potential failure surface plus the seismic shear stresses exceed the yield (peak)
undrained strength S
u
(yield).
This stress-based method for triggering is only appropriate if the safety factor (CRR/CSR) against
triggering determined from the pore-pressure-based method is between 1.0 and 1.4. If the safety
factor is greater than 1.4, then triggering can be assumed to not occur. If the safety factor is less than
1.0, then it should be assumed that triggering will occur. For safety factors between 1.0 and 1.4,
the pore-pressure-based method indicates that triggering is likely, but this stress-based method may
indicate that triggering will not occur.
As discussed in Section 7.4.1, if the design PGA is greater than 0.2g and the CSR is greater than 0.15,
then the triggering analysis described in the following text should not be performed, and strength loss
for loose sand-like material should simply be assumed. The post-earthquake strength of the material
in the loose zone should be assumed equal to S
us
. If evaluating potential measures to improve the sta-
bility of a dam or embankment where the CSR in a zone exceeds 0.15, it may be valuable to perform
the triggering analysis to help assess whether alternative stabilization schemes that would reduce the
CSR might ultimately improve the dam or embankment stability.
The methodology described in the following steps is based primarily on Olson and Stark (2003),
which should be consulted for further details. The steps are:
1. Perform a slope stability analysis to estimate the static shear stress
driving
in the loose
zone of sand-like material. This is accomplished by varying the assumed undrained
strength of the material in the zone until a safety factor of one is achieved. For denser
soils, the peak drained or undrained strength should be used, as discussed in Section
7.4.4.3.
2. Divide the critical failure surface into 10 to 15 segments.
3. Compute the weighted average
vo
along the failure surface and calculate the aver-
age static shear stress ratio
driving
/
vo
.
4. Estimate the average seismic shear stress
av, seismic
applied to each segment of the
failure surface using a 1D or 2D site response analysis as provided by SHAKE or
QUAD4. The value of
av, seismic
for each segment of the failure surface can be taken as
0.65 times
max
obtained from the site response analysis.
5. Estimate the value of S
u
(yield) /
vo
from CPT or SPT data using the proposed equa-
tions in Olson and Stark (2003). The equations give a range of ratios based on SPT or
CPT data. The median value of SPT or CPT data should be used for each zone, and
the lower bound of Olson and Starks equation should be used to compute the ratio
of S
u
(yield) /
vo
.
Commentary: The equations are based on back-calculations of yield (peak) undrained strength
from failure case histories where the failures were triggered by static, not seismic, loading.
Therefore, the back-calculated strengths represent yield (peak) undrained strengths.
6. Compute values of S
u
(yield) and
driving
for each segment along the failure surface,
based on the values of S
u
(yield) /
vo
and
driving
/
vo
(average) and the value of
vo
for
each segment.
< PREVIOUS VIEW
7-36
Chapter 7
MAY 2009
7. Compute the factor of safety against triggering for each segment as:
FS
triggering
= S
u(yield)
/ (
driving
+
av, seismic
) (7-6)
8. Assume that triggering of strength loss occurs for segments where FS
triggering
is less
than or equal to 1.0. Assume that triggering of strength loss does not occur for seg-
ments where FS
triggering
is greater than 1.0.
7.4.3 Evaluation of Post-Earthquake Strength
7.4.3.1 Correlations of SPT and CPT with S
us
of Sand-Like Material
Loose, sand-like refuse and sand-like natural soil may experience signifcant strength loss due to
earthquake shaking, if the shaking is strong enough. The reduced strength is the undrained steady-
state (residual) strength S
us
, which is ofen much lower than the peak undrained strength or the peak
drained strength. Table 7.2 presents references for several correlations of S
us
with SPT data (N
1,60
) and
CPT data (q
t1
) for sand-like material.
The correlations are based on back-calculated values of S
us
from actual fow slides and measured or
estimated values of N
1,60
or q
t1
. The correlations are generally limited to N
1,60
values (uncorrected for
fnes) less than about 12 or q
t1
less than about 75 tsf because no fow slides have been reported for soils
with higher penetration resistances. For higher values of N
1,60
and q
t1
, the drained strength should be
used because, as discussed in Section 7.4.4.2, the material is dilative.
The Seed and Harder (1990) plot (Figure 7.7) is probably the most well known of the S
us
versus SPT
or CPT correlations. The plot is an update of an earlier plot presented in Seed (1987). The 1990 plot
includes a correction to increase N
1,60
values for silty materials to equivalent clean sand N
1,60(CS)
values. The correction varies from 1 to 5 blows per foot depending on the percent fnes. However, no
basis for this correction is provided in either the 1990 or 1987 paper.
TABLE 7.2 REFERENCES FOR CORRELATIONS OF S
US
WITH SPT AND CPT DATA
Reference Correlated Parameters
Seed (1987) S
us
vs. N
1,60
Davis, Castro and Poulos (1988) S
us
vs. N
1,60
Seed and Harder (1990) S
us
vs. N
1,60
Baziar and Dobry (1995) S
us
and S
us
/
v
vs. N
1,60
Castro (1995) S
us
vs. N
1,60
Wride, McRoberts and Robertson (1999) S
us
and S
us
/
v
vs. N
1,60
Yoshimine, Robertson and Wride (1999) S
us
/
v
vs. q
c1N(CS)
Olson and Stark (2002) S
us
/
v
vs. N
1,60
and q
t1
Idriss and Boulanger (2007) S
us
and S
us
/
v
vs. N
1,60
and q
c1N(CS)
Note: S
us
= Undrained steady state (residual) strength.
N
1,60
= Standard Penetration Test (SPT) N-value, normalized to an effective overburden stress of one
atmosphere (typically 1 tsf) and normalized to a hammer effciency of 60 percent.
S
us
/
v
= Undrained steady state strength normalized to vertical effective stress.
q
c1N(CS)
= Cone Penetration Test (CPT) tip resistance normalized to a reference pressure of one atmosphere
and corrected to an equivalent value for clean sand.
q
t1
= CPT tip resistance normalized to an effective overburden stress of one atmosphere.
< PREVIOUS VIEW
7-37
Seismic Design: Stability and Deformation Analyses
MAY 2009
Castro (1995) re-evaluated several of the case histories in Seed and Harder (1990) for which
detailed data were available. He also collected re-evaluations of some of the earlier case histories
performed by Poulos (1988) and Davis et al. (1988). He referred to the re-evaluations collectively
as the GEI data. Castro then reploted the Seed and Harder data points (without a fnes correction)
and compared them to the GEI data points. The resulting plot of representative N
1,60
values and
back-fgured S
us
is shown in Figure 7.8. The original Seed and Harder (1990) lower-bound curve is
somewhat lower than the Seed and Harder lower-bound curve shown in Castro (1995) because the
original Seed and Harder curve included a fnes correction while the Seed and Harder curve shown
in Castro (1995) did not.
Correlations of S
us
/
v
versus SPT and CPT data, back-calculated from case histories, are listed in
Table 7.3. Olson and Stark (2002) re-evaluated the Seed and Harder (1990) data and added new case
histories for development of plots of S
us
/
v
versus both N
1,60
and q
t1
. Olson and Stark did not include
a fnes correction to the N
1,60
values as Seed and Harder did. Olson and Starks plot of S
us
/
v
versus
q
t1
is shown as Figure 7.9. A reasonable lower bound of the Olson and Stark data is a ratio of S
us
/
v
equal to 0.04, which is independent of SPT or CPT value.
Idriss and Boulanger (2007) re-evaluated the Seed and Harder case histories and the Olson and Stark
case histories, and included the fnes correction to the N
1,60
values as Seed and Harder did. Idriss and
Boulanger recommended design curves for both S
us
versus N
1,60(CS)
and S
us
/
v
versus N
1,60(CS)
. Idriss
and Boulangers design curve for S
us
versus N
1,60(CS)
is in between the upper and lower bounds sug-
gested by Seed and Harder in Figure 7.7.
EQUIVALENT CLEAN SAND SPT BLOWCOUNT, (N
1
,
60(CS)
)
R
E
S
I
D
U
A
L
U
N
D
R
A
I
N
E
D
S
H
E
A
R
S
T
R
E
N
G
T
H
,
S
U
S
(
P
S
F
)
FIGURE 7.7 S
us
vs. N
1,60(CS)
(ADAPTED FROM SEED AND HARDER, 1990)
0
400
800
1200
1600
2000
EARTHQUAKE-INDUCED LIQUEFACTION AND SLIDING
CASE HISTORIES WHERE SPT DATA AND RESIDUAL
STRENGTH PARAMETERS HAVE BEEN MEASURED.
EARTHQUAKE-INDUCED LIQUEFACTION AND SLIDING
CASE HISTORIES WHERE SPT DATA AND RESIDUAL
STRENGTH PARAMETERS HAVE BEEN ESTIMATED.
CONSTRUCTION-INDUCED LIQUEFACTION AND
SLIDING CASE HISTORIES.
LOWER SAN FERNANDO DAM
0 4 8 12 16 20 24
FIGURE 7.7 S
us
versus N
1,60(CS)
FIGURE 7.7 S
us
versus N
1,60(CS)
< PREVIOUS VIEW
7-38
Chapter 7
MAY 2009
N1, 60 (NO FINES CORRECTION)
FIGURE 7.8 S
us
versus N
1,60
S
U
S
(
P
S
F
)
SEED AND HARDER (1990)
DAVIS et al. (1988) and
CASTRO (1995)
NO N DATA 1. JUVENILE HALL
2. LOWER SAN FERNANDO
3. MOCHIKOSHI SLIMES
4. LA MARQUESA, UPS
5. LA MARQUESA, DS
6. FORT PECK
7. CALAVERAS
8. HEBER ROAD
9. MOCHIKOSHI DIKE
10. LA PALMA
LOWER BOUND, GEI
LOWER BOUND,
SEED AND HARDER
0
500
1000
7
9
3
3
3
8
10 4
1
5
4
1
6
5
2
2
7
6
0 5 10 15
(CASTRO, 1995)
N DATA
{
GEI
FIGURE 7.8 S
us
versus N
1,60
TABLE 7.3 CORRELATIONS OF S
US
/
V
VERSUS SPT AND CPT DATA BACK-CALCULATED
FROM CASE HISTORIES
Reference
Reported
S
us
/
v
Types of Data and Materials
Baziar and Dobry
(1995)
0.04 to 0.20
Back-calculated values from failure case histories for 9 sites with silty sand
or sandy silt material (more than 10 percent fnes). One of the 9 sites was
identifed as a tailings dam.
Yoshimine,
Robertson and
Wride (1999)
0.03 to 0.19
Back-calculated values from case histories involving multiple submarine
slides at 3 sites. Materials were natural clean sand, silty sand, and sandy
silt.
Olson and Stark
(2002)
0.05 to 0.12
(proposed
limits of ratios)
Back-calculated values from 33 failure case histories, including re-evaluation
of the Baziar and Dobry sites. Four of the 33 sites were identifed as tailings
dams as compared to dams or slopes consisting of natural soils. Four of the
33 sites, including one of the tailings dam sites, had ratios of 0.02 to 0.04,
outside their proposed boundary.
Idriss and
Boulanger (2007)
0.05 to 0.22
Back-calculated values based on select case histories published by Seed
(1987), Seed and Harder (1990), and Olson and Stark (2002) with adequate
amount of in-situ measurements and reasonably complete geometric details
(7 of the 35 case histories reviewed).
FIGURE 7.8 S
us
versus N
1,60
< PREVIOUS VIEW
7-39
Seismic Design: Stability and Deformation Analyses
MAY 2009
Idriss and Boulangers design curve for S
us
/
v
versus N
1,60(CS)
is very close to the best ft line sug-
gested by Olson and Stark in the range where N
1,60(CS)
is less than about 14. For N
1,60(CS)
values higher
than about 12 to 14, which is beyond the range of the case history data, Idriss and Boulanger suggest
two curves: one for the case where void ratio redistribution efects are negligible, and one for the
case where void ratio redistribution efects may be signifcant. The diference between these two
curves is discussed in the following commentary.
Commentary (Use of S
us
/
v
Ratios): Using the ratios of S
us
/
v
for confning pressures signifcantly higher
than the confning pressures from the actual case histories may be unconservative. Twenty-eight of Olson and
Starks 33 case histories had mean
v
of 115 kPa or less. This corresponds to mean
v
of up to 2,400 psf or, for an
assumed efective unit weight of 55 pcf, a mean depth of up to 45 feet. Using S
us
/
v
ratios implies that increasing
the confning pressure by a factor of 5, for example, would also increase S
us
by a factor of 5. But laboratory testing
has shown that the slope of the void ratio versus log S
us
line is ofen steeper than the slope of the void ratio versus
log
v
line. In other words, increasing the confning pressure by a factor of 5 ofen results in increasing S
us
by
a factor less than 5. So using the S
us
/
v
ratios for high confning pressures may be unconservative. Therefore,
recommended guidance is to consider only the lower-bound ratio of 0.04. The lower-bound ratio of 0.04
v
is con-
sidered conservative enough that it is acceptable even at high confning pressures.
The lower bound ratio of 0.04 is applicable to non-plastic materials and to materials with a liquidity index
(LI) of less than one. Materials with LI > 1 are unusual, but may have even lower values of S
us
. At this time,
the recommended lower bound value of S
us
for soils with LI > 1 is 20 psf, but no higher than obtained with a
strength ratio of 0.04, based on the judgment of the authors of this chapter.
The correlations of S
us
with N
1,60
have uncertainty at high confning pressures because the correction factors
used to correct N to N
1
to account for increasing confning pressure are highly uncertain at confning stresses
FIGURE 7.9 S
us
/
v
' vs. q
t1
NORMALIZED CPT TIP RESISTANCE, q
t1
(MPa)
BACK-CALCULATED LIQUEFIED STRENGTH RATIO AND MEASURED CPT
BACK-CALCULATED LIQUEFIED STRENGTH RATIO AND CONVERTED CPT
FROM MEASURED SPT
BACK-CALCULATED LIQUEFIED STRENGTH RATIO AND ESTIMATED CPT
ESTIMATED LIQUEFIED STRENGTH RATIO AND MEASURED, CONVERTED
OR ESTIMATED CPT
NOTE: NUMBER BESIDE SYMBOL INDICATES AVERAGE FINES CONTENT
IN PERCENT
74
7
86
50
13
1
0
100
30
40
15
16
7
20
3
3
20
32
20
55
7
7
7
6
50
36+
25
7
13
3
7
20
85
0 1 2 3 4 5 6 7 8 9 10
0
0.1
0.2
0.3
0.4
OLSON (1998)
PROPOSED
RELATIONSHIP
(ADAPTED FROM OLSON AND STARK, 2002)
S
u
s
/
v
'
FIGURE 7.9 S
us
/
v
versus q
t1
FIGURE 7.9 S
us
/
v
versus q
t1
< PREVIOUS VIEW
7-40
Chapter 7
MAY 2009
higher than about 2 tsf (Youd et al., 2001). However, the uncertainty of applying the S
us
versus N
1,60
correla-
tions at high confning stresses is less than the uncertainty of applying the correlations of SPT or CPT with
S
us
/
v
ratios.
Commentary (Void Ratio Redistribution): The phenomenon of void ratio redistribution is discussed in Section
7.4.5. It refers to the possibility that if a loose zone of sand-like material is overlain by an impervious zone, then
earthquake shaking may cause pore water to migrate toward the interface of the two materials, which could
result in the sand-like material becoming looser and therefore having a lower steady-state strength. There are
no generally accepted methods for evaluating the potential for this to occur, and there is some controversy as to
whether it actually occurs in the feld at all.
The Idriss and Boulanger (2007) design curve for S
us
/
v
versus N
1,60(CS)
splits at values of N
1,60(CS)
higher
than about 12 to 14. For the case where void ratio redistribution efects are negligible, S
us
/
v
increases rap-
idly as N
1,60(CS)
increases above 12 to 14. This is consistent with the fact that sand-like material with N
1,60(CS)
values higher than 12 to 14 tend to be dilative and not susceptible to strength loss (as discussed in more detail
in Section 7.4.4.2.1). For the case where void ratio redistribution efects may be signifcant, S
us
/
v
increases
less quickly as N
1,60(CS)
increases. As Idriss and Boulanger explain, the curve for this case is largely conceptual,
since there are no case history data at these N
1,60
values.
For the purposes of this Manual, and as discussed further in Section 7.4.5, specifc evaluations for potential
void-ratio redistribution efects are not required. As discussed in Section 7.4.4.3, the post-earthquake strength
of sand-like material with N
1,60
greater than or equal to 15 may be based on the drained strength and not on
correlations of S
us
or S
us
/
v
to N
1,60
. However, if redistribution is a concern, the corresponding Idriss and
Boulanger curve can be used.
For estimating S
us
values of sand-like materials based on correlations to SPT and CPT data, the fol-
lowing three-step procedure is recommended:
1. For each zone of material, determine a representative value of N
1,60
. In general, the
representative value should be the median value. For this step, CPT data can be con-
verted to N
1,60
using the relationships discussed in Section 6.4.3.7 (Lunne et al., 1997).
2. Use either the Lower-Bound, GEI curve or the Lower-Bound, Seed and Harder
curve from the Castro (1995) plot (Figure 7.8) to obtain S
us
as a function of repre-
sentative value of N
1,60
. Both curves can be extrapolated to values of N
1,60
as high
as 14 by extending the curves along approximately straight lines. As discussed in
Section 7.4.4.3, drained strength, rather than S
us
, can be used for N
1,60
values of 15
or higher.
3. For each zone, if the resulting value of S
us
is less than 0.04
v
, use S
us
= 0.04
v
instead.
Commentary: As discussed in Section 7.4.1, S
us
of sand-like material is very sensitive to small
changes in void ratio. CPTs and SPTs are not sensitive to small changes in void ratio, and
therefore correlations of CPT and SPT to S
us
are expected to have large scater, as shown in Fig-
ures 7.7, 7.8, and 7.9. The S
us
versus SPT or CPT correlations provide conservative estimates
of S
us
. Generally S
us
is not directly related to either N
1,60
or q
t1
. The case histories used in these
correlations are cases where S
us
was low enough that a fow slide or signifcant deformations
occurred. However, there were almost certainly other sites where N
1,60
or q
t1
were similar, but
fow slides or signifcant deformations did not occur because S
us
was higher than obtained from
the failure case histories. Cases where stability failure did not occur could provide lower-bound
estimates of S
us
(i.e., the minimum value of S
us
needed to maintain stability), but these cases are
not typically studied and would still provide only conservative estimates. Laboratory testing
methods for estimating S
us
may provide less conservative estimates of S
us
.
< PREVIOUS VIEW
7-41
Seismic Design: Stability and Deformation Analyses
MAY 2009
7.4.3.2 Laboratory Testing for Measuring Post-Earthquake Strength for Clay-Like Material
and S
us
for Sand-Like Material
7.4.3.2.1 Laboratory Testing Issues
Soil fabric has been shown in the literature to afect the peak undrained strength of sand-like mate-
rial. This ofen results in diferent peak undrained strengths for undisturbed samples versus recon-
stituted samples. However, initial soil fabric should not afect S
us
, because S
us
is measured at high
strains afer the soil fabric has become remolded. An example of this is presented in Castro, Seed,
Keller, and Seed (1992). Samples of hydraulic sand fll prepared from slurry and by moist tamping
had the same steady-state line.
There is extensive published research that shows that, for peak-undrained strength of clayey soils:
S
u
(triaxial compression) > S
u
(direct simple shear) > S
u
(triaxial extension). However, there are no
data to suggest whether or not steady-state (residual) strength S
us
varies with test type. The dif-
culty is that test methods other than triaxial compression can not generally be run to high enough
strain levels to reach S
us
before non-uniformities in the specimen become so large that the test data
lose meaning. Triaxial extension tests experience necking at relatively low strains. Direct simple
shear tests have signifcant stress non-uniformities (at all strains) because there are no vertical shear
stresses (and therefore no horizontal shear stresses) at the outside edges of the specimen. Investiga-
tors who have tried to evaluate the efect of test type on measured S
us
of sand-like material have
encountered these difculties, which make it difcult to identify S
us
on stress-strain curves from
tests other than triaxial compression.
It is reasonable that the peak strength varies with test type, because the peak strength is very much
dependent on soil structure and therefore on the method of loading. For S
us
, however, the initial
soil structure has been lost (remolded) due to the high strain, so the method of loading should not
be as signifcant. While there might be a diference in S
us
for sand-like material related to the fact
that the intermediate principal stress varies depending on the test type, this diference will be small
compared to the variation in steady-state strength that one should expect for diferent samples
from the same layer or zone of material. Expected in-situ strength variation was previously dis-
cussed in Section 7.4.1.
For most clay-like material, S
us
is such a conservative estimate of post-earthquake strength that minor
diferences that may or may not be a function of test type do not seem signifcant. In any event, the
S
us
value for clay-like soils cannot be measured using triaxial (compression or extension) or direct
simple-shear tests, because the strain to S
us
is so high. Testing of clay-like material to estimate post-
earthquake strengths that are closer to S
up
involves using cyclic loading followed by static loading,
as discussed in Section 7.4.3.2.2. For this testing procedure, diferences in strength between triaxial
compression and other test types should be small compared to the variation in strength one should
expect for diferent samples from the same layer. However, one could, if desired, use direct simple-
shear testing or correct the triaxial test strengths to equivalent direct simple-shear strengths using the
information in Article 20 of Terzaghi, Peck, and Mesri (1996).
Based on the preceding discussion, the use of triaxial compression testing to estimate S
us
for sand-
like material, and to estimate post-earthquake strength for clay-like material, as discussed in Sections
7.4.3.2.2 and 7.4.3.2.3, is reasonable.
7.4.3.2.2 Laboratory Testing of Soft Clay-like Material to Measure Post-Earthquake Strength
As discussed in Section 7.4.2.3.1, the post-earthquake strength for clay-like material is a function of
the shear strains that occur during the earthquake. The post-earthquake strength may be the peak
undrained strength S
up
or it may be a reduced strength. The post-earthquake strength could possibly
(but not likely) be as low as S
us
.
< PREVIOUS VIEW
7-42
Chapter 7
MAY 2009
A conservative estimate of post-earthquake strength for sof clay-like materials is to use S
us
, which
can be obtained by performing feld vane-shear tests and/or CPTs, as described in Section 7.4.3.3. A
less conservative estimate can be obtained by performing laboratory testing to obtain an undrained
strength that is appropriate based on considerations of accumulated strain and/or cyclic stress level.
As discussed in Section 7.4.4.3, the post-earthquake strength of stif clay-like material can be taken as
the peak-undrained strength.
If laboratory testing is performed, strain-controlled, undrained-triaxial (or perhaps undrained direct
simple-shear) tests should be performed to measure: (1) the peak undrained strength, (2) the shear
strain to peak undrained strength, and (3) the drop-of in shearing resistance with continued strain
afer peak. It should be remembered that shear strain in the triaxial test is 1.5 times the axial strain
and that peak strength refers to the peak principal stress diference. Test specimens should be con-
solidated anisotropically to stresses that model in-situ conditions.
If the shear strain at peak strength is less than about 5 percent, or if the shearing resistance drops
of signifcantly afer peak, then either: (1) measure S
us
(in the laboratory or the feld) and use it as a
conservative estimate of post-earthquake strength, (2) perform a series of laboratory tests with cyclic
loading followed by monotonic loading to obtain a less conservative estimate of post-earthquake
strength based on the strain that occurs during cyclic loading, or (3) perform a series of laboratory
tests with cyclic loading followed by monotonic loading to obtain a less conservative estimate of post-
earthquake strength based on the cyclic stress levels applied by the design earthquake. Details for
these options are discussed in the following text. Note that the number of loading cycles in the cyclic
loading sequence must be appropriate compared to the number of representative loading cycles asso-
ciated with the design earthquake magnitude. Seed and Idriss (1982) correlated the number of repre-
sentative loading cycles with earthquake magnitude (e.g., M = 6.5, N = 10; M = 7.5, N = 15; and M = 8.5,
N = 26, where M is earthquake magnitude and N is the number of representative loading cycles).
If the shear strain at peak strength is more than 5 percent, and there is litle drop-of in shearing
resistance afer peak, then options (1), (2), and (3) discussed in the preceding paragraph can be used.
However, a simplifed version of option (2) can also be used. If the seismically-induced shear strains,
as described in Step 2a, are less than one-half the shear strain to peak strength in the monotonic
(static) test, then the post-earthquake undrained strength can be assumed to be equal to the peak
undrained strength S
up
. This is based on previous testing of clay-like materials (Thiers and Seed,
1968; Castro and Christian, 1976; Castro, 2003).
The three options for evaluating the post-earthquake strength of clay-like material are:
1. Conservatively assume that the post-earthquake strength is S
us
and measure S
us
. To
measure S
us
in the feld, CPT or feld vane tests can be used, as discussed in Section
7.4.3.3. To measure S
us
in the laboratory, vane-shear tests can be used. The strain rate
should be high to ensure that the sample is being sheared undrained and that the
undrained S
us
is being measured (see discussion of feld vane-shear rates in Section
6.4.3.8). S
us
cannot normally be reached for clay-like material (except highly sensitive
clays) within the strain limitations of laboratory tests, other than possibly rotation
shear. Since it is unlikely that the post-earthquake strength will be reduced to S
us
except in highly sensitive clays, laboratory testing to measure S
us
of clay-like material
may not be needed in practice.
2. Measure the relationship between cyclic strain and post-earthquake strength. An-
isotropically-consolidated, undrained-triaxial tests (or perhaps undrained, direct
simple-shear tests, also consolidated with an initial shear stress), with cyclic load-
ing followed by monotonic loading, can be used to obtain the relationship between
accumulated shear strain during the earthquake and post-earthquake, undrained
< PREVIOUS VIEW
7-43
Seismic Design: Stability and Deformation Analyses
MAY 2009
strength. Accumulated strain is a measure of the efects of cyclic loading on the post-
earthquake strength of clay-like material. Afer low accumulated strains, the post-
earthquake strength will be close to the peak-undrained strength. Afer very high
accumulated strains, the post-earthquake strength may be close to S
us
. Instead of ac-
cumulated strain, peak strain could also be used as a measure of the efects of cyclic
loading. However, the use of accumulated strains is recommended. A procedure for
performing this type of testing is described in Castro (2003). The basic steps are:
a. Select one or more potential failure surfaces through the embankment.
Estimate the accumulated strains that develop due to earthquake shaking
using either Newmark-type analyses (Section 7.5.4) or numerical model-
ing (Section 7.5.5). If numerical modeling is used to estimate accumulated
strains, the model output will include shear strains. If a Newmark-type
analysis is used to estimate accumulated strains, the analysis will com-
pute displacements. To convert displacements to shear strains, the thick-
ness of the loose zone must be estimated, and the strains should be as-
sumed to be uniform across the full thickness of the zone. In other words,
the accumulated shear strain will equal the displacement divided by the
thickness of the zone. (A Newmark-type analysis can be used instead of
numerical modeling only for cases in which the loose zone is relatively
thin compared to the overall failure mass. CPT data are probably the best
way of delineating the thickness of the loose zone. Another way to delin-
eate the thickness of the loose zone, so that shear strain can be computed
from displacement, is to identify the range of potential failure surfaces for
which the safety factor is within perhaps 10 percent of the safety factor
of the most critical failure surface. The thickness of the range of failure
surfaces can be taken as the thickness of the loose zone.
Commentary: A design earthquake motion (time history of acceleration) and
dynamic soil properties (modulus and damping) are required for performing the
deformation analysis required for this step.
b. Perform a series of tests on undisturbed specimens consolidated aniso-
tropically to in-situ stresses. First apply cyclic stresses to each specimen,
and measure the accumulated strains. Then, without allowing for dissipa-
tion of the pore pressures generated by the cyclic loading, apply mono-
tonic loading to each specimen to measure the peak strength. Perform
tests at various cyclic stresses to bracket the accumulated strains indicated
by the deformation analysis. The cyclic loading portion of the test should
be load-controlled, and the monotonic loading portion of the test should
be strain-controlled.
c. Plot the value of peak undrained strength under monotonic loading
versus the accumulated strain during loading. There should be a trend of
constant monotonic peak strength for small accumulated strains and then
decreasing monotonic strength for higher accumulated strains. (In the
example in Castro (2003), post-earthquake strength was ploted against
maximum cyclic strain instead of accumulated strain, but the concept
is the same. The monotonic strength began to decrease for cyclic shear
strains exceeding about 7.5 percent.)
d. Select the post-earthquake strength as the value of monotonic strength
that corresponds to the estimated accumulated shear strain from the de-
< PREVIOUS VIEW
7-44
Chapter 7
MAY 2009
formation analysis. For clay-like materials, 80 to 100 percent of the peak-
undrained strength should be used, as discussed in Section 7.4.4.3.
3. Measure the efect of cyclic loading on post-earthquake strength This approach
involves performing cyclic undrained shear strength testing on undisturbed samples
of the clay-like material, which are consolidated anisotropically to stresses that model
the more critical in-situ conditions (i.e., higher anisotropic stress ratio for materials
along the critical potential failure zone). The samples should be cyclically loaded for a
conservative number (not less than 20) of cycles, while bracketing a conservative range
of cyclic stress or CSR based on the design earthquake. (Cyclic stress or CSR for the
design earthquake can be obtained from 1D or 2D site response analyses, as described
in Section 7.4.2.2.1, Method 1.) The samples should then be subject to undrained mono-
tonic loading to measure the range in post-cycling S
u
for the range of CSR.
The intent is to test clay-like samples at the higher end of the in-situ anisotropic stress ratio, to a con-
servative number of cycles (not less than 20 cycles) of loading and a conservative range of CSR, not
to obtain a unique relationship between cyclic loading and S
u
, but only to bracket a post-earthquake
S
u
(ofen higher than S
us
) that can be applied in post-earthquake limit equilibrium analyses. This
approach might be helpful in instances where undrained triaxial or direct simple-shear monotonic
tests indicate a shear strain at peak strength somewhat less than 5 percent, but the post-peak drop-of
in shearing resistance is not dramatic or particularly signifcant. It may also be helpful when such
testing indicates a shear strain at peak strength greater than 5 percent, but the post-peak behavior
and/or the perceived sensitivity of the clay-like material warrants more direct evaluation of the und-
rained strength behavior following cyclic loading.
For material that is clay-like, or that borders between sand-like and clay-like behavior, laboratory
testing of undisturbed samples of the site-specifc materials, as described above, should be consid-
ered. The laboratory testing might be monotonic testing to obtain just S
up
and strain to S
up
, or (for
critical structures) cyclic loading followed by monotonic testing to obtain the relationship between
seismically-induced shear strain and post-earthquake strength. If quality undisturbed samples of
site-specifc materials can not practically be obtained, then testing of carefully reconstituted samples
of the site-specifc materials should be considered. However, data from reconstituted samples should
only be used if they can be compared to data from at least some undisturbed samples or a thorough
base of in-situ and laboratory test data, to confrm that the reconstituted samples are representative
of and/or bracket in-situ conditions.
For proposed new facilities, where material for testing is not available, undisturbed or carefully
reconstituted samples of similar materials from other facilities can be used. Once a signifcant base of
testing of fne coal refuse is built up, then it may be possible to estimate strength reduction based on
correlations with previous testing. Provisions should be made to confrm soil properties of proposed
facilities by performing testing as the facility is constructed.
7.4.3.2.3 Laboratory Testing of Sand-like Material to Measure S
us
In loose sand-like material, if the earthquake shaking triggers strength loss, then the post-earth-
quake strength will be equal to the undrained steady state (residual) strengths S
us
. Estimates of
S
us
for sand-like material should frst be made based on correlations with SPT and/or CPT data,
as discussed in Section 7.4.3.1. If desired, laboratory testing to measure S
us
can also be performed.
Laboratory testing, while complex, ofen results in less conservative estimates of S
us
than the SPT
and CPT correlations. If laboratory testing to measure S
us
is planned, the designer should make the
parties responsible for obtaining, transporting, and testing the samples fully aware of the methods
to be followed. For example, since S
us
is highly sensitive to void ratio, high quality undisturbed
samples are required and changes in void ratio must be tracked during material sampling, trans-
< PREVIOUS VIEW
7-45
Seismic Design: Stability and Deformation Analyses
MAY 2009
portation, and testing. Failure to follow the testing guidance may bring laboratory measured values
of S
us
into question. The designer should consider discussing the approach and methods with MSHA
prior to initiating sampling and testing.
The discussion that follows is based primarily on Poulos, Castro, and France (1985), which should be
consulted for further details. In addition, Castro, Seed, Keller, and Seed (1992) discuss application of
this laboratory testing method to the 1971 slide at the Lower San Fernando Dam. Additional informa-
tion is provided in Appendices 7A, 7B, and 7C of this Manual.
At this time, an upper bound is recommended for the values of undrained steady state (residual)
strength S
us
obtained from laboratory testing of sand-like material. That upper bound is an envelope
based on the median N
1,60
value of the loose layer being tested, as indicated in Table 7.4. The purpose
of the upper bound is to make sure that unreasonably high values of S
us
based on laboratory testing
of sand-like material are not proposed.
TABLE 7.4 UPPER-BOUND VALUES OF S
US
FOR CORRESPONDING VALUES OF N
1,60
N
1,60
Maximum Allowable Value of S
us
Based on Laboratory Testing
0 200 psf
5 500 psf
10 1100 psf
14 1700 psf
Some basic concepts associated with laboratory testing for measuring S
us
of sand-like material are:
1. The undrained steady state (residual) strength S
us
is equal to the undrained
strength at high strains, afer the initial structure or fabric of the material has been
fully remolded.
2. S
us
is very sensitive to void ratio and to minor changes in grain-size distribution. As
discussed in Section 7.4.1, deposits of sand-like material have inherent variations in
both. Therefore, multiple samples from the same layer or zone are expected to have
varying values of S
us
.
3. Strain-controlled, undrained triaxial compression tests with pore-pressure mea-
surement are typically used. Experience has shown that isotropic and anisotropic
consolidation in the triaxial cell yield the same values of S
us
. Figure 9 in Castro et
al., (1992) provides a good example. Isotropic consolidation is commonly used for
simplicity.
4. At strains beyond about 20 percent in the triaxial test, non-uniformities in specimen
stress and strain become signifcant.
5. S
us
is obtained from the triaxial test, as shown in the following. Note that these equa-
tions are derived from the Mohr-Coulomb failure envelope in Appendix 7A.
q
s
= (
1s
3s
)/2, where q
s
is one-half the principal stress diference at steady state
sin
s
= q
s
/(
3s
+ q
s
), where
s
is the efective stress friction angle at steady state
S
us
= q
s
cos
s
6. For diferent samples of a given material (all with exactly the same grain-size distri-
bution), S
us
is a function only of void ratio. The relationship between void ratio and
< PREVIOUS VIEW
7-46
Chapter 7
MAY 2009
S
us
for a given soil, with S
us
ploted on a log scale, is referred to as the steady-state
line (Figure 7.10).
7. Since S
us
and
3s
are related by
s
(Appendix 7A), the steady-state line can be plot-
ted as either void ratio versus S
us
or void ratio versus
3s
. Ploting void ratio versus
3
makes it possible to plot the state of the sample before and afer consolidation in
the triaxial cell, as well as at steady state.
8. It has been shown experimentally that similar materials with the same grain shape
and mineralogy, but slightly diferent grain-size distributions, will have steady-state
lines that are parallel but not coincident. Examples are provided in Figures 3, 4, and
5 of Poulos, Castro, and France (1985).
9. High quality undisturbed samples are used to measure S
us
of in-situ material. Be-
cause of unavoidable densifcation that occurs during sampling, handling, and es-
pecially during consolidation in the triaxial cell, S
us
measured in the laboratory will
be higher than the in-situ S
us
. Correcting the value of S
us
measured in the laboratory
back to the in-situ S
us
for a given sample requires two things. First, careful measure-
ments must be made during sampling, handling, and testing to measure the total
change in void ratio from the in-situ condition to the laboratory-consolidated condi-
tion. Second, the slope of the steady-state line must be measured.
100 1,000 10,000
STEADY-STATE STRENGTH, SUS (PSF)
V
O
I
D
R
A
T
I
O
,
e
FIGURE 7.10 STEADY-STATE LINE
0.6
0.7
0.8
FIGURE 7.10 STEADY-STATE LINE
The basic procedural steps in a steady-state (residual) strength laboratory testing program for
sand-like material are provided in the following list (for each zone of interest). Detailed procedures
are presented in Appendix 7B.
1. Identify the zone of loose material based on SPT, CPT, and/or other desired feld
methods. Obtain enough disturbed or undisturbed samples from which a batch mix
can be created that will contain enough soil for at least 5 tests to measure the slope of
the steady-state line. Obtain enough high quality undisturbed samples to perform at
least 8 tests on undisturbed samples from each zone of interest to measure the in-situ
S
us
. A detailed sampling procedure that has been used with fxed-piston sampling to
record sample densifcation during sampling is provided in Appendix 7C.
FIGURE 7.10 STEADY-STATE LINE
< PREVIOUS VIEW
7-47
Seismic Design: Stability and Deformation Analyses
MAY 2009
2. Perform grain-size, hydrometer, specifc-gravity, and (if the material has any plastic-
ity) Aterberg-limits tests on material from the batch mix. Also perform grain-size
tests on the trimmings from all undisturbed samples used to measure the in-situ S
us
.
3. To measure the slope of the steady-state line, prepare at least 5 very loose uniform
specimens from the batch mix for triaxial testing. Moist tamping in 10 layers has
worked well for 3-inch-diameter by 7-inch-tall specimens. Isotropically consolidate
the specimens to varying consolidation pressures such that a range of fnal void ra-
tios is obtained. Carefully measure the volume of the sample in the triaxial cell afer
saturation but before beginning consolidation. This is typically done by applying a
small vacuum to the sample so that it will maintain its shape with no cell pressure
applied. Afer measuring the sample volume, carefully keep track of all volume
change during consolidation. Shear the samples in undrained strain-controlled com-
pression, and measure S
us
at high strains. (See testing Note 2 below for how to iden-
tify whether S
us
has been reached.) Afer the test, oven dry the entire tested sample
to measure its water content. Using the water content and the specifc gravity of the
material, compute the void ratio during shear. Then plot the void ratio during shear
versus S
us
for each test. All the tests should plot close to a straight-line ft of the data
(Figure 7.10). This is the steady-state line for the batch mix.
4. Set up, saturate, and consolidate each undisturbed sample. Measure the sample
volume before consolidation and volume changes during consolidation, as discussed
in the previous step. Shear each undisturbed sample in undrained, strain-controlled
compression and measure S
us
at high strains. Afer the test, oven dry the entire tested
sample to measure its water content and dry weight. Alternatively, measure the en-
tire wet weight, then use half the sample for water content measurement and half the
sample for grain-size and/or specifc-gravity testing. For fne coal refuse, a specifc
gravity measurement must be performed for each undisturbed test sample. Using
the water content, the dry weight, and the specifc gravity of the material compute
the void ratio of the sample during undrained shear (Appendix 7B). The measured
value of S
us
corresponds to the void ratio during undrained shear in the triaxial cell.
To estimate the in-situ value of S
us
for that sample, frst plot the result of the test on
a diagram of S
us
versus void ratio. Then draw a line upward and to the lef from that
data point, parallel to the steady-state line measured for the batch mix. Where the
line drawn reaches the in-situ void ratio, select the value of in-situ S
us
(Figure 7.11).
Refer to Appendix 7B for detailed procedures for measuring the as-tested and in-situ
void ratios of each sample.
5. The eight or more tests on undisturbed samples will result in a range of estimated
in-situ values of S
us
. Select a design value of S
us
that is lower than two-thirds of the
estimated in-situ values and higher than one third of the values. The appropriate to-
tal number of tests that should be performed for each zone depends primarily on the
variability of that zone. An example is shown in Figure 7.12. As shown in the fgure,
one should not expect the as-tested S
us
values from the undisturbed samples to plot
on a straight line of void ratio versus log S
us
, because slight variability in grain-size
distribution from sample to sample may result in signifcant variability in the vertical
position, but not the slope of the correlation of S
us
versus void ratio.
Some notes on testing include the following:
1. Accuracy in measuring void ratio is critical to proper measurement of S
us
. Appen-
dices 7B and 7C discuss the required precision of various measurements in the feld
and laboratory. A given error in void ratio will result in a higher error in S
us
for mate-
< PREVIOUS VIEW
7-48
Chapter 7
MAY 2009
rials with relatively fat steady-state lines. In other words, materials with low slopes
of the steady-state line (low values of e/logS
us
), are more sensitive to errors in the
measurement of void ratio.
The range of slopes of steady-state lines (e/log S
us
) for coal refuse reported for
seven sites in West Virginia and one site in Kentucky was 0.090 to 0.140 (personal
communication, GEI, 2007). For the fatest slope of 0.090 (the most sensitive case),
errors in void ratio measurement would lead to errors in S
us
, as shown in Table 7.5
As discussed in Appendix 7B, void ratios at the various steps of sampling and test-
ing should be computed to the nearest 0.001. The goal is that the computed in-situ
and as-tested void ratios should be correct to within 0.010. The resulting values of S
us
should then be correct to within 30 percent, as indicated in Table 7.5.
2. When looking at a stress-strain curve to identify whether the steady state has been
reached, curves for contractive samples are more defnitive than curves for dilative
samples. Contractive samples will typically exhibit peak strength at small strains
followed by a drop-of in strength and a fatening out at the steady state within the
strain limits of the triaxial test. Dilative samples, on the other hand, typically show a
gradually increasing strength and may not reach steady state within the strain limits
of the test. Also, dilative samples are more likely to develop failure planes during
shear, meaning that steady state is reached only within the thin failure zone where
the void ratio is unknown. Therefore, contractive samples are preferred.
For an undisturbed sample with a given void ratio, consolidating the sample to a
higher efective confning stress before undrained shear is likely to make the sample
more contractive during undrained shear, as explained further in the commentary
that follows. Contractive samples tend to have stress-strain curves that are more
defnitive for interpretation of S
us
, as explained in the previous paragraph. Therefore,
it is ofen good practice to consolidate the sample to a high confning pressure before
undrained shear. The disadvantage is that consolidating the sample to a high con-
fning pressure also decreases the void ratio of the sample in the test and requires a
bigger correction to obtain the in-situ value of S
us
.
FIGURE 7.11 METHOD FOR CORRECTING LABORATORY S
us
TO IN-SITU S
us
100 1,000 10,000 100,000
0.5
0.6
0.7
0.8
0.9
1.0
STEADY-STATE SHEAR STRENGTH, SUS
(PSF)
V
O
I
D
R
A
T
I
O
,
e
FIGURE 7.11 METHOD FOR CORRECTING LABORATORY S
US
TO IN-SITU S
US
IN-SITU VOID RATIO
OF "UNDISTURBED"
SPECIMEN
B
A
VOID RATIO OF
"UNDISTURBED"
SPECIMEN DURING
SHEAR
SUS OF "UNDISTURBED"
SPECIMEN
ESTIMATED
IN-SITU SUS
STEADY-STATE
LINE FOR RECONSTITUTED
SPECIMENS FROM BATCH MIX
FIGURE 7.11 METHOD FOR CORRECTING LABORATORY S
us
TO IN-SITU S
us
< PREVIOUS VIEW
7-49
Seismic Design: Stability and Deformation Analyses
MAY 2009
It is not necessary to consolidate the sample to stresses that model in situ conditions.
This is because the steady-state strength is the strength at high strains and is not
afected by stress history. Therefore, isotropic consolidation is normally used.
Commentary: Consolidated samples that plot well to the right of the steady-state line will be
contractive, as shown in Figure 6 of Poulos et al. (1985). Experience with laboratory testing
tells us that the e log
3
consolidation curve is usually slightly fater than the steady-state
line (ploted as e versus log
3
per item 7 on page 7-46). Therefore, increasing the consolidation
stress tends to move the sample further to the right compared to the steady-state line, and the
sample tends to become more contractive.
At the steady state, both the principal stress diference and
3
should be constant
with strain.
100 1,000 10,000
STEADY-STATE STRENGTH, S
US
SUS MEASURED AFTER
LABORATORY CONSOLIDATION
SUS AFTER CORRECTION
TO IN-SITU VOID RATIO
STEADY STATE LINE
FROM BATCH MIX
V
O
I
D
R
A
T
I
O
,
e
FIGURE 7.12 INTERPRETATION OF LABORATORY S
US
TESTING
0.5
0.6
0.7
0.8
0.9
FIGURE 7.12 INTERPRETATION OF LABORATORY S
us
TESTING
TABLE 7.5 ERROR IN VOID RATIO AND CORRESPONDING PERCENT ERROR IN S
US
Error in void
ratio (e)
logS
us
(1)
S
us
(measured) /
S
us
(actual)
(2)
Error in S
us
(%)
0.005 0.0555 1.14 14
0.010 0.1111 1.29 29
-0.005 -0.0555 0.88 12
-0.010 -0.1111 0.77 23
Note: 1. logS
us
= e/0.090 = [log S
us
(measured)] - [log S
us
(actual)] = log [S
us
(measured)/S
us
(actual)]
2. S
us
(measured) / S
us
(actual) = 10
(logSus)
FIGURE 7.12 INTERPRETATION OF LABORATORY S
us
TESTING
< PREVIOUS VIEW
7-50
Chapter 7
MAY 2009
3. It is helpful to plot the compression curves (e - log
3
) along with the values at
steady state. Especially for the samples tested from the batch mix to measure the
steady-state line, these plots will help determine what consolidation stresses are
needed to achieve contractive samples at desired void ratios. (Figure 7.13)
4. Published data for coal refuse, based on laboratory testing as described above, in-
clude: (1) a reported range of 0.06 to 0.27 for S
us
/
v
and (2) a reported slope of the
steady-state line e/log S
us
in the range of 0.11 to 0.13 based on laboratory testing
of 34 undisturbed samples (PI in the range of 0 to 12) of fne coal refuse from fve
West Virginia sites (Genes et al., 2000). Also, reported slopes of the steady-state line
e/log S
us
for one site in West Virginia and one site in Kentucky were found to be
0.09 and 0.14 (GEI, 2007).
Because the preceding parameters are based on limited data and a wide range of ratios, the data
should be used carefully. Also, as discussed in Section 7.4.3.1, using the ratios of S
us
/
v
for confning
pressures signifcantly higher than the confning pressures from the actual samples may be uncon-
servative. However, if values of S
us
are measured using laboratory testing of site-specifc materials or
comparable deposits at comparable confning pressures, then the site-specifc ratios may be valuable
for adjusting the measured S
us
values for other portions of the embankment, or for future construc-
tion stages (Genes et al., 2000).
Over time, if a database of strength ratios for fne coal refuse is developed, it should become possible
to use these ratios with more confdence.
Estimation of strengths using strength ratios requires an estimate of the consolidation pressure. One
should not make the assumption a priori that the soils are normally consolidated without obtain-
ing confrming data and monitoring. In areas where fne coal refuse will serve as the foundation
for embankment construction, the design plan should include provisions for monitoring pore pres-
sures and, if signifcant upstream construction is planned, controlling the rate of construction to
mitigate excess pore pressure development and/or reducing the likelihood that signifcant zones of
fne refuse will be under consolidated as construction proceeds. When evaluating existing embank-
ments, piezometer data should be obtained to allow estimation of efective consolidation stresses.
Also, pore-pressure-dissipation tests performed with the piezocone at various depths are an efective
CONSOLIDATION IN
TRIAXIAL CELL
UNDRAINED SHEAR
STEADY-STATE LINE
V
O
I
D
R
A
T
I
O
,
e
FIGURE 7.13 CONSOLIDATION CURVE AND STEADY-STATE LINE
0.9
0.8
0.7
0.6
0.5
100 1000 10,000 100,000
MINOR EFFECTIVE PRINCIPAL STRESS,
3
FIGURE 7.13 CONSOLIDATION CURVE AND STEADY-STATE LINE
FIGURE 7.13 CONSOLIDATION CURVE AND STEADY-STATE LINE
< PREVIOUS VIEW
7-51
Seismic Design: Stability and Deformation Analyses
MAY 2009
way of obtaining the pore-pressure profle at a given location. Consolidation tests on undisturbed
samples of clayey layers can also be used to determine the current consolidation stresses on the
clayey layer and adjacent more sandy layers.
7.4.3.3 Field testing to Measure S
us
of Soft Clay-Like Material
Field vane-shear testing can be used to estimate S
us
of clay-like material. For the feld vane measure-
ments to be valid, they must be performed to high enough strains to reach S
us
and must be performed
quickly to be confdent that the material is being sheared undrained. Recommended vane shear test-
ing procedures are provided in Chapter 6, Section 6.4.3.8. It is a good idea to pair feld vane-shear
tests with CPTs, to confrm that the feld vane is being performed in a layer of clay-like material.
CPT sleeve friction measurements can also be used for estimating S
us
of clay-like material (Lunne et
al., 1997), although comparison feld vane tests are recommended because the CPT sleeve friction
is less reliable in sensitive clays. CPTs should be performed with pore-pressure measurements to
confrm that the pore-pressure response indicates undrained shear. Lack of elevated pore-pressure
response may indicate that the material is behaving as drained rather than undrained.
Commentary: For most clay-like materials the post-earthquake strength will be higher than S
us
. Therefore,
using S
us
as an estimate of post-earthquake strength may be overly conservative. Laboratory methods of esti-
mating post-earthquake strength of clays, as discussed in Section 7.4.3.2.2, may provide less conservative esti-
mates of post-earthquake strength.
As a lower bound, clay-like material with a liquidity index of less than 1.0 can be considered to have
a post-earthquake strength of 0.04
v
. If the liquidity index is greater than or equal to one, a lower
bound post-earthquake strength of 20 psf, but no higher than 0.04
v
, may be used. These values
are guidance based on the judgment of the authors. Clay-like materials with a liquidity index less
than 1.0 are unlikely to lose strength all the way to S
us
due to earthquake shaking. A reduction in
undrained strength for these materials from a pre-earthquake peak strength of 0.2
v
(a value within
the representative range for natural clays) to a post-earthquake strength of 0.04
v
is considered con-
servative. Clay-like material with a liquidity index of 1.0 or higher may act like a quick clay with
a very low remolded or post-earthquake strength. Therefore, a lower bound of just 20 psf, but no
higher than 0.04
v
, is considered appropriate.
7.4.4 Analysis Steps
Engineers use various methods for evaluating triggering and post-earthquake strength. Engineers
also perform these evaluations and the related limit-equilibrium, slope-stability analyses in varying
sequences. The following steps represent a generalized approach that can be adjusted for individual
projects. Reference should be made to the fow chart in Figures 7.1a, 7.1b and 7.1c.
7.4.4.1 Step 1 - Defne Embankment Geometry
Defne the geometry of the embankment and of the material zones within the embankment. Identify
zones of clay-like versus sand-like materials.
7.4.4.2 Step 2 - Screen for Potential Strength Loss
Review the subsurface conditions at the embankment to evaluate whether any zones have the poten-
tial for strength loss due to earthquake shaking.
7.4.4.2.1 Sand-Like Material
For this initial screening step, saturated to nearly-saturated, sand-like materials with N
1,60
values less
than 15, or q
t1
values less than 75 tsf, should be considered potentially susceptible to strength loss.
< PREVIOUS VIEW
7-52
Chapter 7
MAY 2009
Commentary: The screening criteria that sand-like materials with N
1,60
values greater than 15 or q
t1
values
greater than 75 tsf are not susceptible to strength loss are based on case-history data from large earthquakes
indicating that fow slides have only occurred where N
1,60
values were 13 or lower (Seed and Harder, 1990;
Castro, 1995; Wride et al., 1999) and q
c1
values were less than 65 tsf (Olson and Stark, 2002). Olson and Stark
(2003) present curves showing dilative versus contractive behavior for sands. The boundary for dilative versus
contractive in terms of N
1,60
values varies from 10 to 15. The boundary in terms of q
c1
values varies from 65 to
85. Also, Seed (1979) presents data showing that sands with relative densities higher than about 50 percent or
N
1
values higher than about 15 have limited strain potential and therefore may experience cyclic mobility but
not strength loss, even though pore pressures may have reached 100 percent.
Many experts agree that a fnes correction might be reasonable in this screening analysis for sand-like material.
That is, a silty sand may be less likely to be contractive than a clean sand with the same N-value. However, the
data to support this are sparse and vague. Therefore, a fnes correction should not be applied here unless addi-
tional data is published in the future that can justify it.
It is interesting to note that the curves presented in Youd et al. (2001) indicate that liquefaction may occur
during large earthquakes in sand-like material having N
1,60
values as high as 30, and q
c1N
values as high as
150. This is because, in Youd et al. (2001), liquefaction refers to increases in pore pressure but not necessarily
to strength loss. As stated in the introduction to Youd et al. (2001): In moderately dense to dense materials,
liquefaction leads to transient sofening and increased cyclic shear strains, but a tendency to dilate during shear
inhibits major strength loss and large ground deformations. (The term increased cyclic shear strains is
referred to herein as cyclic mobility.) In other words, moderately dense to dense sand-like materials (N
1,60
values
higher than 15 or q
t1
values higher than 75 tsf) may experience high excess pore pressures and cyclic mobility,
but they will not experience strength loss.
The criterion that sand-like materials with N
1,60
values greater than 15, or q
t1
values greater than 75 tsf, be con-
sidered not susceptible to strength loss is a conservative criterion. Many materials with N
1,60
values less than
15, or q
t1
values less than 75 tsf, will be dilative and not susceptible to strength loss. Therefore, if almost all N
1,60
values exceed 13 and the mean N
1,60
value exceeds 15 within a zone, or almost all q
t1
values exceed 65 tsf and the
mean q
t1
value exceeds 75 tsf within a zone, then the zone can be considered not susceptible to strength loss.
Conservative criteria and conservative engineering judgments are appropriate and intended at this step for
three reasons: (1) litle if any of the data supporting these criteria are based on coal refuse materials (although,
considering that coal refuse particles are more compressible than natural soil particles, one might expect these
criteria to be even more conservative, because the compressibility of the coal refuse may mean that SPT and
CPT values are lower for coal refuse than for natural soils at the same void ratio), (2) there is some uncertainty
in the overburden corrections used to convert measured feld values to N
1,60
and q
t1
values, and (3) no further
analysis is recommended for materials that satisfy these criteria.
Thorough site-specifc feld and laboratory testing should ultimately yield data other than N
1,60
and q
t1
to beter
evaluate whether sand-like materials at a site are or are not susceptible to strength loss. If in-situ and laboratory
test data are available during the initial screening phase, it should also be used in classifying sand-like zones as
susceptible or not susceptible to strength loss.
7.4.4.2.2 Clay-Like Material
For this initial screening step, clay-like material may be screened for zones with potential for sig-
nifcant strength loss based on CPT data, as shown in Figure 7.14 (Robertson, 2008). This fgure is a
slightly more conservative version of a similar fgure presented in Robertson and Wride (1998). In this
fgure, Q is equivalent to q
c1N
and F is the friction ratio. Computation of both Q and F is discussed in
Youd et al. (2001). CPT data in Zone B indicate clay-like material for which signifcant strength loss is
unlikely. CPT data in Zone C indicate potentially highly sensitive clay-like material for which signif-
cant strength loss can occur (Zones A1 and A2 indicate sand-like material). If CPT data are not avail-
< PREVIOUS VIEW
7-53
Seismic Design: Stability and Deformation Analyses
MAY 2009
able, then SPT data can be used along with Aterberg limits to confrm that the material is clay-like.
N > 6 corresponds to Zone B and N < 6 corresponds to zone C. For this screening step, the N-values
of the clay-like material should be corrected for hammer efciency per Youd et al. (2001), but should
not be corrected for overburden pressure. Overburden pressure corrections are applicable only to
sand-like material, not clay-like material. Undrained strength data, if available, can also be used.
Peak undrained strength of 1500 psf or higher corresponds to Zone B, and peak undrained strength
less than 1,500 psf corresponds to zone C.
Commentary: These screening criteria for clay-like material are recommended guidance. No generally accepted
screening criteria for susceptibility to strength loss were found in the literature. Selection of these criteria is
based on the authors judgment and experience, and considering published data from sites where liquefaction
(strength loss or excess pore pressure) was observed in soils with signifcant fnes content.
N
O
R
M
A
L
I
Z
E
D
C
P
T
T
I
P
R
E
S
I
S
T
A
N
C
E
,
Q
FIGURE 7.14 USE OF CPT DATA TO SCREEN CLAY-LIKE MATERIAL
FOR SUSCEPTIBILITY TO STRENGTH LOSS
0.1 1 10
1
10
100
1000
I
c
= 2.6
A
2
2
3
C
7
6
A
1
B
5
4
N
O
R
M
A
L
L
Y
C
O
N
S
O
L
I
D
A
T
E
D
I
N
C
R
E
A
S
I
N
G
O
C
R
,
A
G
E
,
C
E
M
E
N
T
A
T
I
O
N
I
N
C
R
E
A
S
I
N
G
O
C
R
A
N
D
A
G
E
IN
C
R
E
A
S
IN
G
S
E
N
S
IT
IV
IT
Y
NORMALIZED FRICTION RATIO, x 100%
NOTE: 1. VALUES OF Q AND F ARE COMPUTED FROM CPT DATA AT THE DEPTHS OF INTEREST,
AS DESCRIBED IN YOUD ET AL. (2001).
2. CPT DATA THAT PLOT IN ZONES A1 AND A2 INDICATE MATERIAL THAT IS NOT CONSIDERED
CLAY-LIKE, SO THIS SCREENING METHOD IS NOT APPLICABLE.
3. CPT DATA THAT PLOT IN ZONE B INDICATE CLAY-LIKE MATERIAL THAT IS NOT SUSCEPTIBLE
TO STRENGTH LOSS.
4. CPT DATA THAT PLOT IN ZONE C INDICATE CLAY-LIKE MATERIAL THAT MAY BE SUSCEPTIBLE
TO STRENGTH LOSS.
5. NUMBERED ZONES SEPARATED BY DASHED BLUE LINES ARE FOR REFERENCE ONLY AND
ARE A GUIDE TO SOIL TYPES, AS DESCRIBED IN YOUD ET AL. (2001). FOR EXAMPLE, DATA IN
ZONE 3 INDICATE SILTY CLAY TO CLAY WHILE DATA IN ZONE 6 INDICATE CLEAN SAND TO
SILTY SAND. THE LINE SEPARATING SAND-LIKE FROM CLAY-LIKE MATERIAL CORRESPONDS
TO A SOIL BEHAVIOR TYPE INDEX (IC) OF 2.6, AS ALSO DESCRIBED IN YOUD ET AL. (2001).
f
s
q
t
-
v0
FIGURE 7.14 USE OF CPT DATA TO SCREEN CLAY-LIKE MATERIAL
(ADAPTED FROM ROBERTSON, 2008; 2008 NRC CANADA OR ITS LICENSORS; REPRODUCED WITH PERMISSION.)
FIGURE 7.14 USE OF CPT DATA TO SCREEN CLAY-LIKE MATERIAL
< PREVIOUS VIEW
7-54
Chapter 7
MAY 2009
In applying these criteria, CPT data are considered the most reliable, laboratory strength data the next most
reliable, and SPT data the least reliable.
7.4.4.2.3 Screening Results
If no zones within the embankment or foundation are susceptible to strength loss, then the seis-
mic stability of the embankment is acceptable (assuming, of course, that the static stability has been
analyzed and is acceptable, including cases where undrained strengths are considered). No further
analyses of seismic stability are needed. The next step is to evaluate seismic deformations.
If one or more zones within the embankment or foundation are susceptible to strength loss, then fur-
ther analyses should be performed as discussed in Step 3, which follows.
7.4.4.3 Step 3 - Defne Post-Earthquake Strengths for Limit Equilibrium Stability Analyses
Stability analyses are typically performed using 2D limit-equilibrium, slope-stability sofware for
one or more potentially critical cross sections of the embankment. The stability analyses should be
static analyses using post-earthquake strengths. Analyses should be performed for both potential
upstream and potential downstream failures.
Upstream failures may or may not pose a risk of uncontrolled release, depending on whether the
location of the critical surface leaves adequate freeboard and crest width in place. Guidelines for
deciding whether potential upstream failures pose a safety hazard are discussed in Sections 6.6.4.1
and 6.6.4.3.
Values of post-earthquake strength for various zones of the embankment should be selected as dis-
cussed in the following:
Dense sand-like materials (N
1,60
> 15 and q
t1
> 75 tsf) such as compacted coarse refuse
and dense sand-like natural soils tend to be dilative when they are sheared. That is, the
undrained strength tends to be higher than the drained strength. Also, these materi-
als do not experience strength loss due to earthquake shaking. For post-earthquake
stability analysis, one cannot be certain whether the material will act as if it is drained
or undrained, and the negative pore pressures required to mobilize a higher strength
may not develop because they cause cavitation. Therefore, it is reasonable and conser-
vative to use the drained strength for these materials, as discussed in Chapter 6.
Stif clay-like materials (SPT N > 6 and CPT data in Zone B) tend to have high
shear strain up to the peak undrained strength and limited drop-of in shearing
resistance afer the peak, so they should not experience signifcant strength loss
due to earthquake shaking. Unlike dense sand-like materials, stif clay-like materi-
als should act as undrained during the most critical earthquake and post-earth-
quake period. For these materials, 80 to 100 percent of the peak undrained strength
should be used.
Commentary: Available data indicate that the cyclic stresses and strains caused by earth-
quakes are unlikely to cause signifcant strength reduction in stif clays. However, some practi-
tioners have commonly used 80 percent of the peak undrained strength for clays that are subject
to signifcant seismic loading, presumably to account for possible strength degradation and
other factors. Therefore, although it is perhaps overly conservative, the 80-percent value is rec-
ommended in areas of moderate to high seismic hazard potential. Use of 100 percent of the peak
undrained strength is reasonable in areas of low seismic hazard potential and when applying
a factor of safety of 1.5 in initial post-earthquake stability evaluations to classify the structure
and foundation (Figure 7.1a, Boxes 5 and 6).
< PREVIOUS VIEW
7-55
Seismic Design: Stability and Deformation Analyses
MAY 2009
For loose saturated sand-like material, as in most fne coal refuse impoundments,
strength should be selected based on the results of triggering analyses, and S
us
es-
timates should be made in accordance with the previous sections of this chapter. If
triggering of strength loss is assumed or computed to occur, use S
us
. If triggering is
assumed or computed to not occur, use the peak undrained strength, but no higher
than the drained strength. The S
up
values for sand-like materials can be obtained
from published data such as Castro (2003) and Olson and Stark (2003). Alternatively,
the S
up
values can be obtained from the testing on remolded samples performed as
part of the testing program described in Section 7.4.3.2.3. Tests on remolded samples
with values of steady-state strength and confning pressure similar to in-situ samples
can be considered representative of in-situ conditions.
Drained strengths for these materials can be used if they are clearly determined to be
unsaturated (saturation ratio less than 80 percent), and it can be demonstrated that
they will remain unsaturated. (When the degree of saturation is higher than 80 per-
cent, strength loss is still possible for loose, sand-like material.) Materials above the
phreatic surface should not necessarily be assumed to be unsaturated. Materials above
the phreatic surface may be saturated or close to saturated, particularly if the water
level was previously higher than the current phreatic surface.
For sof or sensitive clay-like material (SPT N < 6 or CPT data in Zone C), strength
should be selected based on the results of strength-loss estimates made in accordance
with Sections 7.4.2.3.1, 7.4.3.2.2, and 7.4.3.3. If triggering (signifcant strain during the
earthquake) is assumed or computed to occur, use a value lower than S
up
, but prob-
ably not as low as S
us
. (If the zones of this material are relatively small and probably
not signifcant to the seismic stability, then S
us
can be used as the post-earthquake
strength even though it is probably overly conservative. For highly sensitive clays,
S
us
may be the appropriate post-earthquake strength.) If triggering (signifcant strain
during the earthquake) is computed to not occur, use 80 percent of S
up
. S
up
for clay-
like material can be obtained from CPT data, feld vane shear testing, or laboratory
testing. S
up
for borderline material can be obtained from CPT data (if the push is
undrained based on pore-pressure response) or from laboratory testing or possibly
from feld vane-shear tests performed at high strain rates.
Coal refuse impoundments tend to be stratifed. Variations in post-earthquake undrained strengths
may be more signifcant than variations in drained strengths. Therefore, post-earthquake strengths
should be selected conservatively. For zones of mixed coarse and fne refuse (such as the lower por-
tion of an upstream stage), use strength properties weighted toward the properties of the overlying
coarse refuse zone or the underlying fne refuse zone, based on the prevalence and extent of each
zone, the index properties of each zone, and a realistic estimation of the upstream embankment
cross section (based on subsurface conditions and past experience with displacement of fnes during
upstream construction pushouts).
Commentary: Strain compatibility is not an issue for these stability analyses, because the strengths used all
represent reasonably conservative estimates of the undrained strength that the soil might have at high strains
(5 percent to 10 percent). As discussed previously, the recommended strengths are:
Loose sand-like material Undrained steady-state (residual) strength. This is the lowest
resistance that the material can have at any strain beyond a few percent.
Dense sand-like material Drained strength. The undrained stress strain curve of a dense,
dilative, sand-like material increases steadily with strain with no drop-of at high strain. For
these materials, the drained strength is a conservative estimate of the undrained resistance at
any strain beyond a few percent.
< PREVIOUS VIEW
7-56
Chapter 7
MAY 2009
Sof clay-like material (SPT N < 6 or CPT data in Zone C) Either: (1) the undrained steady
state (residual) strength as a very conservative estimate of the available strength, (2) the
post-earthquake peak strength estimated from cyclic tests followed by monotonic (static) tests,
or (3) 80 percent of the peak-undrained strength. But option 2 or 3 can only be used if the
clay-like material has a broad peak to the stress-strain curve, with no rapid drop-of in resis-
tance afer peak, which is the case for most clay-like material. So, the strength selected for use
should be appropriate for a wide range of strain.
Stif clay-like material (SPT N > 6 and CPT data in Zone B) 80 to 100 percent of undrained
peak strength. The undrained stress-strain curve of stif clay-like material reaches nearly its
peak strength at a few percent strain and then levels of or increases slowly with strain with
no drop-of at high strain. So, again, the strength used is appropriate for a wide range of
strain. As discussed above, the 80-percent value is recommended in areas of moderate to high
seismic-hazard potential. Use of 100 percent of the peak-undrained strength is reasonable in
areas of low seismic-hazard potential, and when applying a factor of safety of 1.5 in initial
post-earthquake stability evaluations to classify the structure and foundation (Figure 7.1a,
Boxes 5 and 6).
7.4.4.4 Step 4 - Perform Initial Stability Analysis
First perform an initial, conservative, post-earthquake stability analysis. This initial stability analysis
is performed with the assumption that the earthquake shaking is strong enough that soils potentially
susceptible to strength loss do in fact experience strength loss.
Use S
us
for sand-like material in zones of the embankment that, based on the screening analysis in
Section 7.4.4.2, have low N-values and CPT values and may therefore be susceptible to strength loss.
For this analysis, S
us
for sand-like materials can be conservatively estimated from correlations with
SPT and/or CPT data (Section 7.4.3.1). Also, the post-earthquake strength of clay-like material consid-
ered not susceptible to signifcant strength loss (Zone B on the CPT chart or N > 6) can be estimated
as 80 to 100 percent of S
up
, as discussed in Section 7.4.4.3. The post-earthquake strength of clays
considered potentially susceptible to signifcant strength loss should be taken as an estimated value
of S
us
. S
up
and S
us
for clays can be estimated from CPT data, feld vane shear data, or laboratory data
(Chapter 6 and Section 7.4.3.3).
The slope stability analysis is a total-stress analysis using undrained strength, which is a function
of the pre-earthquake consolidation stress. The location of the phreatic surface afects the pre-earth-
quake consolidation stresses and thus the undrained strength. The estimated location of the phreatic
surface at the time of the earthquake should be used in the stability analysis. Elevated pore pressures
caused by earthquake shaking need not be considered.
If these initial stability analyses (for various potential upstream and downstream failure surfaces)
indicate a minimum safety factor of at least 1.2, then seismic stability is considered acceptable, and
the next step is to perform a deformation analysis.
If the minimum safety factor is below 1.2, then the embankment geometry should be modifed to
make the factor of safety acceptable, or else more sophisticated (and less conservative) evaluations
of post-earthquake shear strength and triggering, as subsequently discussed, may be appropriate.
If the safety factors are very low, then it may be clear at this initial stage that some modifcation to
the embankment design is needed. It is ofen useful to vary the assumed values of S
us
in the zones
potentially susceptible to strength loss to fnd the values of S
us
that would be needed to obtain a
safety factor against instability of 1.2, and then judge if the materials in question could possess such
strengths before undertaking more detailed evaluations of S
us
.
< PREVIOUS VIEW
7-57
Seismic Design: Stability and Deformation Analyses
MAY 2009
Commentary: A safety factor of 1.2 is recommended for this analysis because, even though the shear strengths
selected for this analysis are fairly conservative, there may be signifcant uncertainties in the geometry of the
embankment and the various zones within the embankment and foundation.
7.4.4.5 Optional Step 5 - Perform More Detailed Evaluation of S
us
for Sand-Like Material
Obtain representative high quality undisturbed samples of sand-like material in zones that may
experience strength loss (Section 7.4.4.2). Perform a laboratory steady-state shear strength testing
program involving monotonic undrained tests on both undisturbed and reconstituted specimens
(Section 7.4.3.2.3) to obtain beter estimates of S
us
.
Re-run the limit equilibrium stability analyses using the beter estimates of S
us
. If the minimum safety
factor is at least 1.2, then the seismic stability of the embankment is acceptable, and seismic deforma-
tions should be evaluated next.
Commentary: The preceding discussion refers to laboratory testing to measure S
us
for sand-like material.
Performing this testing before undertaking triggering analyses should be considered, because the measured
values of S
us
might result in satisfactory seismic stability, and thus triggering would not be an issue. Also, the
S
us
testing will give information on stress-strain behavior, including strain at peak-undrained strength, which
may be used in the triggering analysis. However, this step involves complex sampling and testing, and it can
be deferred until afer a triggering analysis is performed. This step can also be skipped entirely, especially if the
post-earthquake strength of sand-like materials was not a signifcant factor in the initial stability analyses.
7.4.4.6 Optional Step 6 - Perform Triggering Analysis
If the minimum safety factor is still less than 1.2 afer a more detailed evaluation of S
us
, then for loose
sand-like material, consider performing a triggering analysis, as discussed in Section 7.4.2, to evalu-
ate whether the earthquake is large enough to trigger strength loss in the critical zones.
For sand-like materials, triggering analyses should only be performed if the design earthquake is
relatively small. If the design earthquake produces a peak ground acceleration at the embankment
site of more than 0.2g and causes a cyclic stress ratio (CSR) within the zone of interest of more than
0.15, then triggering of strength loss in sand-like material potentially susceptible to strength loss
should simply be assumed.
If triggering analyses are performed for sand-like materials, the pore-pressure-based method, as dis-
cussed in Section 7.4.2.1 (Youd et al., 2001) should be performed frst. If the safety factor against trig-
gering using the pore-pressure-based method (CRR/CSR) is higher than 1.4, then triggering can be
assumed to not occur. If the safety factor is less than 1.0, then triggering of strength loss should be
assumed. For safety factors of between 1.0 and 1.4 calculated using the pore-pressure-based method,
triggering of strength loss is possible. Either assume that triggering of strength loss will occur, or per-
form more rigorous triggering analyses (strain-based or stress-based methods as discussed in Section
7.4.2.3) to make a fnal evaluation of whether or not triggering occurs.
For clay-like material, a strain-based triggering analysis (possibly involving laboratory testing with
cyclic loading followed by monotonic loading as discussed in Section 7.4.3.2.2) can be performed to
evaluate whether the earthquake shaking is strong enough to trigger a decrease in S
up
and to deter-
mine what the decrease will be.
If strength loss is not triggered in one or more of the zones that had previously been assumed to expe-
rience strength loss, then re-run the stability analysis using the appropriate, higher strength in those
zones as discussed in Step 3, Section 7.4.4.3. If the safety factor is at least 1.2, then the seismic stability
is acceptable, and seismic deformations should be evaluated next.
< PREVIOUS VIEW
7-58
Chapter 7
MAY 2009
If the safety factors are still less than 1.2, then the embankment is not seismically stable. Additional
investigations should be performed to beter characterize the geometry and materials in the embank-
ment or the embankment must be redesigned or modifed.
7.4.5 Comments on Methods for Evaluating Triggering and Strength
All of the methods described in this chapter require special expertise, and should only be applied by
geotechnical engineers with experience in seismic stability and deformation analyses.
Laboratory steady-state (residual) strength testing programs provide site-specifc values of S
us
that
are ofen less conservative than the values obtained from empirical correlations with SPT and CPT
data, but these testing programs require sophisticated undisturbed sampling and laboratory testing.
Strain-based triggering analyses require sophisticated laboratory testing and sophisticated analyses
of shear strains induced by the design earthquake for evaluation of whether the earthquake shaking is
strong enough to cause strength loss down to S
us
. Thus, steady-state laboratory programs and strain-
based triggering analyses are normally performed only for large embankments where the additional
testing and analysis costs are relatively small compared to the cost of redesigning or modifying the
embankment.
If contractive zones of material are confned by overlying layers of material with low hydraulic con-
ductivity, then drainage of excess pore water generated by earthquake shaking can be impeded.
There is some indication from centrifuge model tests that this may result in migration of pore water
toward the overlying low-hydraulic-conductivity layer, with consequent redistribution of void ratio
and loosening of the material near the interface with the low-hydraulic-conductivity layer. This in
turn reduces the post-earthquake strength of the material near the interface. This mechanism was
suggested by Whitman in the 1980s. Recent work on the subject includes Malvick et al. (2006) and
Naesgaard et al. (2005). There are currently (2008) no generally accepted methods for analyzing this
potential mechanism, and therefore it need not be directly considered in stability analyses at this time.
A method for including the potential efects of void ratio redistribution on post-earthquake strength
has been proposed by Idriss and Boulanger (2007). As discussed in Section 7.4.3.1, this method can be
used for sand-like material with N
1,60
> 15 if one has a concern for the possibility of void-ratio redis-
tribution (contractive zones overlain by lower-hydraulic-conductivity materials that would impede
dissipation of excess pore-water pressure). In such a case drainage provisions (such as wick drains,
as discussed in Section 6.6.3.3) could be incorporated. The need for such drainage provisions for dis-
sipation of pore pressure in fne coal refuse at a slurry impoundment could be assessed based on
pore-pressure monitoring during upstream construction.
Commentary: Pore-water migration and void-ratio redistribution (which may or may not actually occur
outside laboratory centrifuge tests) are not considered in either laboratory testing or SPT/CPT methods for
estimating post-earthquake strength. The laboratory testing relates to pre-earthquake void ratios. If there is
pore-water migration, the relationship of void ratio to strength obtained from the laboratory testing would
allow one to quantify the efects of a given amount of void ratio change (assuming that one day we will have
methods to estimate to what extent, if any, pore-water migration takes place). SPT and CPT data also refect
pre-earthquake conditions. One cannot expect that SPT or CPT data will be able to predict whether or not,
and to what extent, pore-water migration might take place, and what impact the pore-water migration might
have on post-earthquake strength.
At the present time (2008), it is debatable whether pore-water migration did or did not occur for the
case histories used to develop the relationships between SPT/CPT data and post-earthquake strength. For
example, extensive investigations of the Lower San Fernando Dam slide indicate that the characteristics of
the slide can be explained without assuming pore-water migration (Castro, Seed, Keller, and Seed, 1992).
< PREVIOUS VIEW
7-59
Seismic Design: Stability and Deformation Analyses
MAY 2009
7.5 SEISMIC DEFORMATION ANALYSES
7.5.1 General Discussion
The procedures described in Sections 7.5.2 through 7.5.6 are appropriate for sites where seismic stabil-
ity has been shown to be adequate based on analyses discussed in the previous sections of this chapter.
Even if the seismic stability is adequate, a dam or embankment will deform during earthquake shak-
ing because of the development of accumulated strains along a potential sliding surface due to the
superposition of seismic and static (driving) shear stresses. Two mechanisms for developing accumu-
lated strains should be considered:
Yielding mechanism When seismic accelerations in an embankment are high
enough, total shear stresses (static plus cyclic) along a potential failure surface
may tend to exceed the available shear strength. The soil along the potential failure
surface yields during short earthquake time increments when the total shear stress
tends to exceed the available shear strength. Another way of saying this is that the
soil along the potential failure surface yields during the short time increments of the
earthquake when accelerations are higher than the acceleration that can be transmit-
ted by the soil. This mechanism can occur in both sand-like and clay-like material.
The resulting deformations can be evaluated using Newmark-type analyses.
Ratcheting mechanism As shown in Figure 7.15, cyclic stress-strain curves are
sofer in loading than in unloading. When there is a static shear stress, repeated
cyclic loading will result in accumulated strains (deformations) in the direction of the
static shear stress, even if the available shear strength is not exceeded. These ratchet-
ing accumulated strains (deformations) are generally signifcant only if cyclic mobil-
FIGURE 7.15 CYCLIC STRESS-STRAIN CURVE DEMONSTRATING
RATCHETING MECHANISM
0.6
0.4
0.2
0.0
-0.2
SHEAR STRAIN (%)
S
H
E
A
R
S
T
R
E
S
S
N
O
R
M
A
L
I
Z
E
D
T
O
C
O
N
S
O
L
I
D
A
T
I
O
N
P
R
E
S
S
U
R
E
(SEED ET AL., 2003)
-2 0 2 4 6 8 10 12
FIGURE 7.15 CYCLIC STRESS-STRAIN CURVE DEMONSTRATING RACHETING MECHANISM
FIGURE 7.15 CYCLIC STRESS-STRAIN CURVE DEMONSTRATING RACHETING MECHANISM
< PREVIOUS VIEW
7-60
Chapter 7
MAY 2009
ity occurs. If the cyclic loading involves stress reversal (i.e., the cyclic shear stress
exceeds the static consolidation shear stress), and if enough cycles of loading occur
that pore pressures approach 100 percent, then the cyclic stress-strain curve will
develop an S-shape, and the cyclic strains will become signifcant. This type of cyclic
stress-strain behavior is referred to as cyclic mobility. If there is no static shear stress,
then cyclic mobility will not cause signifcant accumulated strain. Similarly, if there is
a static shear stress, but cyclic mobility does not occur, then accumulated strains will
generally be insignifcant. To evaluate whether a material will develop cyclic mobil-
ity due to earthquake shaking, the Youd et al. (2001) procedure should be used to
evaluate whether excess pore pressures will approach 100 percent. For the purpose
of deformation analyses, a safety factor (CRR/CSR) of 1.0 or less should be consid-
ered to result in cyclic mobility.
For both the yielding and ratcheting mechanisms, the deformations are driven by the earthquake
shaking. Therefore, the magnitude of the deformations is related to the intensity of the earth-
quake. This is in contrast to seismic instability, where the instability failure may be triggered by
the earthquake shaking, but the instability failure is driven by the static (gravity) weight of the
embankment.
Deformations may involve the embankment crest and both the upstream and downstream embank-
ment slopes. Deformations along the upstream slope may or may not pose a risk of uncontrolled
release, depending on whether the location and magnitude of the potential movement leaves ade-
quate freeboard and crest width in place. Guidelines for deciding whether potential upstream slope
movements pose a safety hazard are discussed in Sections 6.6.4.1 and 6.6.4.3. If these guidelines are
met, detailed deformation analyses may not be required (i.e., if deformation screening or simpli-
fed analysis indicates acceptable performance for potential sliding masses that include the crest or
provide overall dam freeboard and integrity, then detailed analyses may not be needed for other
potential sliding masses that do not afect the integrity of the dam). While this point may also apply
to potential movements on the downstream slope, the deformation, strain and stability of the remain-
ing embankment should be evaluated considering both undrained and drained strength conditions
and the potential for progressive failure.
Commentary: If seismic stability and deformation screening or simplifed analysis demonstrate that a sub-
stantive portion of the embankment and crest would be preserved and would be sufcient to preclude a release
from the impoundment (although other portions of the embankment might be compromised), then detailed
deformation analysis to assess the signifcance of other zones that might be subject to cyclic mobility may not be
warranted. This judgment should be based on the site-specifc cross-section and material properties with care-
ful consideration of: (1) the presence, extent and depth of zones of potential cyclic mobility and (2) mitigating
factors that would limit large deformations or otherwise provide support for the adjoining embankment (e.g.,
the confguration of the upstream slope and impoundment might restrict the amount of displacement that could
be expected, thus providing some remaining confnement and support to preserve a substantial portion of the
foundation and embankment).
Limit-equilibrium slope stability analyses, employing post-earthquake strengths, might be useful in identify-
ing the foundation-embankment zone most prone to signifcant deformation and in judging what portion of the
foundation and embankment would remain to preclude a release from the impoundment. For broader dams and
broader upstream embankment stages, limit-equilibrium analyses with post-earthquake strengths may be ade-
quate to justify that sufcient embankment breadth, crest width and freeboard would remain such that detailed
deformation analyses are not warranted.However, for narrower dams and narrower upstream embankment
stages founded on signifcant zones subject to potential cyclic mobility, the initiation and progression of exces-
sive deformation and breaching of the crest may be a concern, and detailed deformation analyses are likely to be
unavoidable. The distinction between broader and narrower dams and embankments is not defnitive, but
< PREVIOUS VIEW
7-61
Seismic Design: Stability and Deformation Analyses
MAY 2009
professional judgments can be supported by the margin of safety suggested by the static and seismic (post-earth-
quake) stability analyses, considering potential failure surfaces within portions of the embankment-foundation
that are anticipated to be preserved (e.g., factors of safety much greater than the minimum recommended values
of 1.5 and 1.2, respectively) and supported by a limited extent of potential failure refected by surfaces with low
factor of safety (e.g., near the recommended minimum). Generally, to justify such judgment, potential failure
surfaces with a low factor of safety should be limited in extent relative to the entire embankment breadth.
7.5.2 Preliminary Screening
If the site is in a low seismic-hazard area, which is defned in Section 7.7.3.7 as an area where the
National Seismic Hazard Map of the United States (Frankel et al., 2002) indicates that the PGA with
a return period in 2,500 years is less than or equal to 0.10g, and if the seismically-induced, pore-pres-
sure buildup does not approach 100 percent (which could lead to cyclic mobility), then deformations
will be small enough that no further evaluation of deformations is required. As noted in Section 7.5.1,
all deformation evaluations, including this preliminary screening, are applicable only if the embank-
ment has frst been determined to be seismically stable. Therefore, for preliminary screening, the fol-
lowing steps can be followed, as indicated in the fow chart in Figure 7.1c:
1. Is this a structure-foundation system for which no deformation analysis at all is
required (e.g., structures of the type indicated in Figure 7.1a, Box 1)? If the answer is
Yes, then no seismic deformation analysis is required. If the answer is No, then
continue to Step 2.
2. If the site is in an area of low seismic hazard potential (as defned in Section 7.7.3.7),
then continue to Step 3. If the site is in an area of higher seismic hazard area, then skip
to Step 5.
3. In sand-like materials, evaluate the potential for pore-pressure increase using the
methods discussed in Section 7.4.2.2 (Youd et al., 2001). If the safety factor against
100 percent pore-pressure increase in any zone (CRR/CSR) is greater than 1.0, then
cyclic mobility will not develop in that zone. For clay-like materials, confrm the
safety factor from the static stability analysis using post-earthquake strengths. If the
safety factor is greater than 1.2, then signifcant deformations in zones of clay-like
material are unlikely.
4. If the analyses in Step 3 do not indicate cyclic mobility in sand-like material and the
analyses in Step 3 indicate high safety factors for static stability in clay-like material,
and if the site is in a low-seismic-hazard-potential area, then deformations will be
small enough that no further evaluation of deformations is required.
Commentary: These criteria are intended to be conservative and represent recommended guid-
ance based on experience with deformation analyses.
5. If the criteria in Step 3 are not met, or if the site is not in an area of low seismic haz-
ard potential, then permanent deformations should be evaluated, as described in
Sections 7.5.3 through 7.5.6, which follow.
7.5.3 Pseudo-Static Procedure for Screening
A relatively simple pseudo-static procedure (Hynes-Grifn and Franklin, 1984) can be used for
screening of seismically-induced deformations. This procedure is applicable only if all of the follow-
ing conditions are met:
Design earthquake of magnitude less than 8 (M < 8).
No signifcant zones susceptible to strength loss, as determined during the seismic
< PREVIOUS VIEW
7-62
Chapter 7
MAY 2009
stability analyses (Sections 7.4.2 or 7.4.4.2).
Small displacements (less than 3 feet) are not signifcant to the performance of the
dam or embankment.
For this procedure, limit-equilibrium slope-stability analyses are performed with a pseudo-static
seismic coefcient of one-half the PGA determined at the base of the embankment. For clay-like
material, 80 percent of the peak undrained shear strength should be used. For sand-like material, 80
percent of the peak undrained shear strength, but no higher than 80 percent of the drained strength
should be used. If the factor of safety is greater than 1.0 (based upon a thorough scope of geotechnical
investigation or a high confdence level in material characterizations) or greater than 1.2 (based upon
a moderate geotechnical investigation or less certain material characterization), then the deforma-
tions can be assumed to be less than 3 feet.
Commentary: Hynes-Grifn and Franklin (1984) suggested that a safety factor of 1.0 is adequate for assum-
ing deformations would be less than 3 feet. The 1.2 safety factor has been added for cases where there is limited
confdence in the site characterization. The method was developed based solely on earthquake records from the
1971 San Fernando earthquake. However, it is considered acceptable to apply the method to eastern as well as
western sites. This screening analysis applies to any height of embankment. The height is taken into account
indirectly by the pseudo-static limit-equilibrium analysis.
7.5.4 Newmark-Type Analysis (No Cyclic Mobility)
This method of analysis addresses the yielding mechanism of strain accumulation discussed in Sec-
tion 7.5.1. In these Newmark-type analyses (Newmark, 1965), deformations are analyzed by consid-
ering movements of a sliding mass over a sliding surface caused by yielding of the soil at its available
strength. The basic assumption is that movement (or yielding) occurs when the sum of the static plus
the seismic shear stresses along the sliding surface reaches the value of the available shear strength.
Thus, the acceleration that can be transmited to the sliding mass is limited to a value referred to as
the yield acceleration.
A Newmark-type analysis is not appropriate for very large deformations because it is based on the
assumption that the embankment geometry is the same before and afer the earthquake.
The basic steps are as follows:
1. For a given sliding mass, perform a pseudo-static stability analysis to determine
the horizontal acceleration that produces a safety factor of one. This is the yield ac-
celeration k
y
, and it has the units of gravity. Critical sliding masses may be selected
by searching for the most critical mass (critical sliding surface) while varying the
pseudo-static horizontal acceleration until a safety factor of one is achieved. The es-
timated location of the phreatic surface at the time of the earthquake should be used.
Elevated pore pressures caused by earthquake shaking need not be considered. The
material strengths for the pseudo-static stability analyses should be the available un-
drained strengths determined as indicated in Sections 7.4.1 through 7.4.3 and 7.4.4.3.
In general, the following strengths should be used:
Loose sand-like material Undrained peak strength S
up
or undrained steady-
state (residual) strength S
us
depending on results of triggering analysis
Dense sand-like material Peak undrained strength S
up
but no higher
than peak drained strength S
dp
Sof to medium clay-like material 80 percent of S
up
Stif clay-like material 80 percent of S
up
or 80 percent of the post-earth-
quake strength determined as per Section 7.4.3
< PREVIOUS VIEW
7-63
Seismic Design: Stability and Deformation Analyses
MAY 2009
Commentary: The use of 80 percent of S
up
instead of 100 percent of S
up
for stif
clay-like materials in Newmark deformation analyses is based on a recommenda-
tion in Makdisi and Seed (1978). The recommendation accounts for the fact that
some deformation accumulates when the peak stresses exceed about 80 percent of
S
up
. The Newmark analysis assumes that there is no deformation until the selected
strength is reached. The 80-percent value results in some computed deformation
when stresses exceed 80 percent of peak.
2. Defne a time history of average acceleration of the sliding mass k(t) during the de-
sign earthquake. Values of k(t) are in units of gravity. The maximum value of k(t) is
referred to as k(max). One- or two-dimensional computer programs such as SHAKE,
FLUSH, or QUAD4 can be used to compute k(t) by modeling the propagation of the
earthquake motion from bedrock to the zone of interest. The k(t) time history is not
the same as the acceleration time history output by SHAKE (or other programs) at the
elevation of the base of the sliding mass. Rather, to obtain k(t), one should take the
time history of seismic shear stress output by the computer program at the base of the
sliding mass and divide it by the total overburden pressure at that elevation (accelera-
tion equals force/mass or stress/pressure). In this way, k(t) will represent the average
acceleration of the sliding mass and not the acceleration of individual layers or ele-
ments at the base of the sliding mass. Note that the total overburden pressure must
not include the overburden stress caused by any free water above the ground level.
Commentary: In SHAKE, or other programs, potential yielding of the soil is ignored, so
the computed accelerations, k(t) can be higher than the yield acceleration. The analysis is
said to be decoupled because it separates the computations of accelerations and the compu-
tations of deformations due to yielding, when in fact they occur simultaneously. However,
comparisons with more rigorous coupled methods indicate that generally the uncoupled
method is satisfactory. Wartman et al. (2004) conclude that uncoupled analyses are sat-
isfactory for cases where the predominant frequency of the motion is at least 1.3 times the
natural frequency of the soil mass, which is generally the case for earthquake shaking of
dams and embankments.
3. During time intervals that k(t) exceeds k
y
, displacements are initiated. Displacements
continue even afer k(t) becomes less than k
y
, until the velocity of the mass becomes
zero. Compute the total displacement of the sliding mass by double-integrating
the portion of the k(t) curve above and below k
y
for time intervals where velocity is
greater than zero (Figure 7.16).
4. If performing one-dimensional analyses at diferent points along the potential failure
surface to compute k(t), then the resulting displacements must be combined in order
to estimate the displacement of the overall potential failure wedge. A method for
combining the computed displacements is presented in Appendix 7D.
The deformations computed using a Newmark-type analyses are appropriate only if there is no sig-
nifcant sofening (loss of stifness) of the soil caused by cyclic mobility, as described in the second
bullet point (ratcheting mechanism) in Section 7.5.1. This is because the Newmark-type analyses are
based on the assumption that no displacements are initiated during time intervals when k(t) is less
than k
y
, but signifcant displacements could occur during those time intervals if cyclic mobility and
resulting accumulated strains are occurring.
Commentary: Newmark-type analyses should generally not be used if cyclic mobility is expected (i.e., if
CRR/CSR 1.0). Newmark-type analyses are based upon the assumption that there is no deformation when
the sum of static plus seismic shear stresses is less than the available shear strength. But if cyclic mobility
occurs, then signifcant deformations may occur due to the ratcheting mechanism described in Section 7.5.1
< PREVIOUS VIEW
7-64
Chapter 7
MAY 2009
even when the sum of the shear stresses is less than the available strength. Therefore, if cyclic mobility occurs,
the Newmark-type analyses will under-predict deformations. In areas of low seismic hazard potential, the
earthquake shaking may not be high enough to cause cyclic mobility. For small earthquakes, CRR/CSR may
be greater than 1, such that Newmark-type analyses may be valid even for cases of upstream construction
with loose sand-like material.
A simplifed procedure for evaluating embankment deformations using the Newmark method
was developed by Makdisi and Seed (1978). The method provides estimates of embankment
deformations if the following information is known or a representative range of values can be
established:
The natural period of the embankment
The maximum crest acceleration due to earthquake shaking
The earthquake magnitude
The method was developed for embankments in the range of 100 to 200 feet high with a natural
period ranging between 0.26 seconds and 5.22 seconds, based on several western earthquake records
with magnitudes of about 6.5 to 8.25. A simplifed formula for estimating the natural period of the
embankment (T = 2.6 H/V
s
where H is the height of the embankment and V
s
is the representative
shear-wave velocity) is provided in Makdisi and Seed (1977). This simplifed formula for the natural
period is applicable to embankments with a cross section that is approximately triangular and there-
fore may not be applicable to non-trapezoidal coal refuse embankments. For embankment confgura-
tions that deviate from a triangular cross section, a range in natural period may be preferable to an
approximate single value. Because of the limited range of conditions upon which it was based, the
Makdisi and Seed method is not considered appropriate for evaluation of signifcant- or high-haz-
FIGURE 7.16 DOUBLE INTEGRATION TO OBTAIN DISPLACEMENT
FIGURE 7.16 DOUBLE INTEGRATION TO OBTAIN DISPLACEMENT
D
I
S
P
L
A
C
E
M
E
N
T
V
E
L
O
C
I
T
Y
A
C
C
E
L
E
R
A
T
I
O
N
t
a
t
b
t
c
A
A
A
B
B
B
t
a
t
c
t
b
t
c
t
a
t
b
FIGURE 7.16 DOUBLE INTEGRATION TO OBTAIN DISPLACEMENT
< PREVIOUS VIEW
7-65
Seismic Design: Stability and Deformation Analyses
MAY 2009
ard-potential coal refuse impoundments unless the embankment under consideration approximates
a triangular confguration or more conservative confguration (with respect to seismically-induced
deformation) and is founded on materials that are not susceptible to cyclic mobility.
7.5.5 Numerical Modeling with No Cyclic Mobility
If stress reversal and cyclic mobility are not likely (i.e., if CRR/CSR > 1.0, as discussed in Section 7.5.1),
then Newmark analyses may be used. An alternative to Newmark analyses for this case is computer
modeling based on fnite-element or fnite-diference analyses.
With fnite-element modeling, the cross section of the structure is divided into a grid of discrete
(fnite) elements that connect to each other at node points. The elements are assigned material prop-
erties such as unit weight, Poissons ratio, and stress-strain properties. Loads (forces) are applied at
the node points, and displacements at the node points are computed by solving the stifness matrix
for the assembly of fnite elements. Strains and stresses within elements are computed based on the
displacements at the node points and the material properties.
With fnite-diference modeling, the cross section of the structure is divided into a grid of lumped
masses that interact with each other in accordance with constitutive equations based on the mate-
rial properties. The displacements resulting from forces applied to the lumped masses in the grid
are computed by solving the equations of motion (force equals mass times acceleration) at sequen-
tial time steps.
One of the key issues is assigning appropriate material properties (especially stress-strain proper-
ties) to the elements. Various non-linear soil models have been developed that, when used properly,
can reasonably model material behavior when cyclic mobility does not occur. The soil models can
be complex, and using them without fully understanding them can lead to errors. The results of
the analyses are ofen sensitive to relatively minor variations in the parameters used to defne the
stress-strain properties. Checking the sensitivity by varying the parameters is an important part of
the analysis. It is also important to check the model output to be sure that the patern of stresses and
strains across the cross section of the structure appears reasonable.
7.5.6 Numerical Modeling Considering Cyclic Mobility
If stress reversal and cyclic mobility are likely, (i.e., if CRR/CSR < 1.0) then the evaluation of deforma-
tions becomes highly uncertain, and Newmark analyses (Section 7.5.4) are not appropriate. Finite-
element or fnite-diference numerical models can be used, but the stress-strain models must have
special provisions to account for cyclic mobility. These models are particularly complex because they
model the S-shaped, stress-strain curve associated with cyclic mobility (very low stifness at low
to moderate strains followed by dilation and increasing stifness at high strains). An example of
such a stress-strain model is presented in Byrne et al. (2004). It is particularly important to check
the output of these analyses for sensitivity to modeling parameters and for reasonableness. The U.S.
Army Corps of Engineers demonstrated a procedure using commercially available fnite-diference
sofware (FLAC) and a modifed version of the UBSAND model (Byrne et al., 2004) to conduct a
deformation analysis for a dam in California, including comparison of the results with somewhat
more simplifed methods (Perlea et al., 2009).
Cyclic mobility is most likely to be a concern for cases of upstream construction in areas of medium
to high seismic risk. Deformation analyses of complex embankment-foundation confgurations and
situations where signifcant zones of the structure are subject to cyclic mobility require experience
with numerical analysis techniques (e.g., fnite element and fnite diference) and soil behavior con-
stitutive models. An experienced user of the selected modeling sofware for deformation analyses
should be involved in the modeling and/or validation of the results.
< PREVIOUS VIEW
7-66
Chapter 7
MAY 2009
7.5.7 Acceptable Deformations
Once estimates of deformation have been made, the question arises as to what values of deformation
are considered acceptable and what values are considered excessive. For embankments that do not
retain liquids, fne coal refuse, or similar materials, deformations are not normally a problem so long
as the embankment is stable. For dams or embankments that do retain liquids, fne coal refuse, or
similar materials, deformations should be small enough that:
Freeboard is not signifcantly compromised.
Cracking through the dam, which afects the integrity of the dam for retaining liq-
uids or might allow a release of material, does not occur.
Appurtenant structures that afect dam safety are not severely damaged.
Afected structures can be repaired before they might deteriorate further and pose a
greater hazard.
The following are generally considered acceptable:
Downstream analyses Computed displacements or deformations less than 25 per-
cent of the available freeboard and less than 3 feet.
Upstream analyses Computed displacements or deformations less than 25 percent
of the available freeboard and less than 6 feet.
7.6 EMBANKMENT MODIFICATIONS FOR IMPROVING STABILITY OR RESISTANCE TO
DEFORMATIONS
Measures to improve the seismic stability and the resistance to deformations of new and existing
facilities can be grouped into the following general categories:
Changes to embankment geometry (e.g., fater slopes, butressing, wider coarse
refuse zone)
Drainage control measures (including both internal and surface drainage)
Soil/refuse improvement or reinforcement
Increased freeboard
These measures are discussed in Chapter 6.
7.7 SEISMIC HAZARD ASSESSMENT (SEISMICITY)
The purpose of a seismic hazard assessment is to estimate the appropriate design earthquake(s) for a
site in terms of earthquake magnitude (M) and peak ground acceleration (PGA) and, when needed,
to defne representative time histories of ground motions.
This discussion of recommended procedures for deriving the seismic hazard is intended to be appli-
cable in the U.S. where coal is mined. Figure 7.17 shows the various regions where anthracite, bitu-
minous and lignite coals are mined. Although the greatest amount of coal production is from the
western U.S., the type of coal processing in that region does not typically require the construction
of tailings dams for coal refuse. Accordingly, the procedures defned to derive seismic hazard place
emphasis on eastern U.S. conditions, which are diferent from the western U.S. in many respects, as
discussed in Section 7.7.1.
Anthracite coal is mined in eastern Pennsylvania. Bituminous coal deposits in the eastern U.S. are
found in the Appalachian, Illinois, Michigan, and the Western Interior Basins, with some bituminous
coal also found within the Gulf Coast lignite deposits (Figure 7.17). These sources, primarily anthra-
< PREVIOUS VIEW
7-67
Seismic Design: Stability and Deformation Analyses
MAY 2009
cite in eastern Pennsylvania and bituminous in the Appalachian and Illinois Basins, are where coal
needs to be processed and where coal refuse impoundments are required. Within these large areas,
the seismic hazard is not uniform and can range from being nearly insignifcant to an important
design consideration.
The U.S. Geological Survey (USGS) has prepared national seismic hazard maps based on probabi-
listic analyses that defne the variation of seismic hazard throughout the conterminous U.S., Alaska,
Hawaii, Puerto Rico and the U.S. Virgin Islands, among other regions. Figures 7.18 and 7.19 present
USGS seismic hazard maps respectively for the PGA with a 10 percent and 2 percent probability of
exceedance in 50 years (~ 500 year and ~ 2,500-year return periods, respectively) in the contermi-
nous U.S. Surface or near-surface active faults are the main sources of seismic hazard in the western
U.S., whereas the eastern U.S. coal felds are afected primarily by three less-well-defned seismic
sources: the New Madrid Seismic Zone, the Eastern Tennessee Seismic Zone, and the Charleston,
South Carolina Seismic Zone (Figure 7.20). The coal felds with the most signifcant seismic hazard
LEGEND
ANTHRACITE/OTHER USES
ANTHRACITE/POTENTIALLY MINABLE
LIGNITE/OTHER USES
LIGNITE/POTENTIALLY MINABLE
LOW VOLATILITY BITUMINOUS/OTHER USES
LOW VOLATILITY BITUMINOUS/POTENTIALLY MINABLE
MEDIUM AND HIGH VOLATILITY BITUMINOUS/OTHER USES
MEDIUM AND HIGH VOLATILITY BITUMINOUS/POTENTIALLY MINABLE
SUB-BITUMINOUS/OTHER USES
SUB-BITUMINOUS/POTENTIALLY MINABLE
FIGURE 7.17 LOCATION OF U.S. COAL RESOURCES
NOTE: IMAGE WAS OBTAINED FROM THE
USGS NATIONAL ATLAS WEB SITE
AND WAS COMPILED BY THE USGS
EASTERN ENERGY RESOURCES
TEAM (EERT).
FIGURE 7.17 LOCATION OF U.S. COAL RESOURCES
FIGURE 7.17 LOCATION OF U.S. COAL RESOURCES
< PREVIOUS VIEW
7-68
Chapter 7
MAY 2009
in the eastern U.S. are those in the southern half of the Illinois Basin. Portions of the coal felds in the
southern Appalachians have moderate hazard. The Michigan and Western Interior Basins have very
low hazard.
Seismic evaluation of coal refuse facilities and embankment dams requires the estimation of at least
two ground motion parameters (earthquake magnitude and PGA). Another factor that needs to be
considered in association with these two parameters is frequency content. Earthquake magnitude is
critical because of the general relationship of magnitude and duration, as well as the strength of ground
motion and related parameters such as frequency content, atenuation characteristics, etc. In essence,
the larger the earthquake magnitude, the longer the earthquake, which implies that the soil will have
to resist a greater number of ground-motion cycles of greater magnitude. Magnitude is not the only
consideration in evaluating earthquake duration, as variations in fault mechanism, distance from the
source, and local geologic conditions also contribute to duration, but acceptable practice is to consider
LEGEND
20
15
10
9
8
7
6
5
4
3
2
1
0
FIGURE 7.18 SEISMIC HAZARD MAP OF THE U.S. WITH OVERLAY
OF MAIN BITUMINOUS COAL FIELDS
NOTE: 1. IMAGE WAS OBTAINED FROM THE
USGS NATIONAL ATLAS WEB SITE
AND WAS COMPILED BY THE USGS
GEOLOGIC HAZARDS TEAM.
2. ANTHRACITE AND BITUMINOUS
COAL FIELDS ARE DENOTED BY
RED BORDERS.
PEAK HORIZONTAL GROUND ACCELERATION
IN % g WITH A 10% PROBABILITY OF
EXCEEDANCE IN A 50-YEAR PERIOD
FIGURE 7.18 SEISMIC HAZARD MAP OF THE U.S. WITH OVERLAY OF MAIN
BITUMINOUS COAL FIELDS
FIGURE 7.18 SEISMIC HAZARD MAP OF THE U.S. WITH OVERLAY OF MAIN
BITUMINOUS COAL FIELDS
< PREVIOUS VIEW
7-69
Seismic Design: Stability and Deformation Analyses
MAY 2009
duration as a function of magnitude (Youd et al., 2001). PGA is the most commonly used parameter to
defne the strength of ground motion, but its signifcance also depends on frequency content. The actual
seismic hazard will depend on the combination of duration and ground motion strength. With respect
to dams and embankments, a single cycle of high PGA at a high frequency (in the range of 0.25 to 0.45g
at a predominant frequency of 25 Hz) will likely not be as signifcant as a larger number of cycles of
lower acceleration (in the range of 0.10 to 0.15g at 1 Hz). The reason for this is that for a dam to be sig-
nifcantly excited by earthquake ground motion, some portion of the ground motion must be close to
the predominant frequency of the dam. A simple method for estimating the predominant frequency of
an embankment based on height and shear-wave velocity of the embankment materials is provided by
Makdisi and Seed (1977). For a typical coal refuse dam ranging in height from 100 to 400 feet and with FIGURE 7.19 ILLUSTRATION OF MODERATE- TO HIGH-HAZARD REGION
LEGEND
300
200
160
120
80
60
50
40
30
20
18
18
16
14
12
10
8
6
4
2
0
PEAK HORIZONTAL GROUND ACCELERATION
IN % g WITH 2% PROBABILITY OF EXCEEDANCE
IN A 50-YEAR PERIOD (2475-YEAR RETURN PERIOD).
NOTE: 1. IMAGE REDRAWN FROM 2002 HAZARD
MAP FOR A 2% PROBABILITY OF EXCEED-
ANCE IN 50 YEARS FROM THE USGS
"NATIONAL ATLAS" WEB SITE.
ZONE WHERE COMPLETE DSHA OR
PSHA IS REQUIRED FOR EVALUATION
OF SEISMIC HAZARD. BY INFERENCE,
A SIMPLIFIED EVALUATION OF SEISMIC
HAZARD IS PERMISSIBLE IN ALL OTHER
AREAS.
2. ANTHRACITE AND BITUMINOUS
COAL FIELDS ARE DENOTED BY
RED BORDERS.
FIGURE 7.19 ILLUSTRATION OF MODERATE- TO HIGH-HAZARD REGION
FIGURE 7.19 ILLUSTRATION OF MODERATE- TO HIGH-HAZARD REGION
< PREVIOUS VIEW
7-70
Chapter 7
MAY 2009
average shear wave velocities ranging from 500 to 1,600 f/sec, the predominant frequency for the frst
mode of vibration could range from about 0.5 to 6 Hz. Accordingly, coal refuse dams are more likely
to be afected by low-frequency ground motion, implying that more than one earthquake source may
need to be considered in order to fully characterize the seismic hazard at a given site.
FEMA (2005b) defnes the following design earthquakes:
Maximum Credible Earthquake (MCE) The MCE is the largest earthquake magni-
tude that could occur along a recognized fault or within a particular seismotectonic
province or source area under the current tectonic framework.
Maximum Design Earthquake (MDE) or Safety Evaluation Earthquake (SEE) This
is the earthquake that produces the maximum level of ground motion for which
a structure is to be designed or evaluated. The MDE or SEE can be set equal to
the MCE or to a design earthquake less than the MCE, depending on the circum-
LEGEND
AGE OF QUATERNARY FAULTING
HISTORIC
HOLOCENE (< 15,000 YEARS AGO)
LATE QUATERNARY (< 130,000 YEARS AGO)
OBSERVED EVIDENCE OF PALEOLIQUEFACTION
MID-LATE QUATERNARY (< 750,000 YEARS AGO)
QUATERNARY (< 1.6 MILLION YEARS AGO)
POSSIBLY OLDER THAN QUATERNARY
NOTE: 1. IMAGE WAS OBTAINED FROM THE
USGS NATIONAL ATLAS WEB SITE
AND WAS COMPILED BY THE USGS
GEOLOGIC HAZARDS TEAM.
2. BITUMINOUS COAL FIELDS ARE
DENOTED BY RED BORDERS.
FIGURE 7.20 QUATERNARY FAULTS AND SIGNIFICANT SEISMIC SOURCES
WITH OVERLAY OF MAIN BITUMINOUS COAL FIELDS
CHARLESTON SEISMIC ZONE -
TWO INTERPRETATIONS GIVEN
EQUAL PROBABILITY OF BEING
CORRECT IN THE 2002 UPDATE
FOR THE NATIONAL SEISMIC
HAZARD MAP.
PEMBROKE FAULTS/GILES
COUNTY SEISMIC ZONE -
THE LARGEST APPALACHIAN
EARTHQUAKE IS IN THIS
ZONE, BUT THE FAULTING
APPEARS TO BE PRE-
HOLOCENE
WABASH VALLEY SEISMIC ZONE -
IDENTIFIED BASED ON PALEOLIQUE-
FACTION EVENTS INTERPRETED
TO NOT BE RELATED TO THE NEW
MADRID SEISMIC ZONE THAT
OCCURRED ~ 6,100 YEARS BP.
EASTERN TENNESSEE
SEISMIC ZONE - SOURCE
USED FOR THE 1996 NATIONAL
SEISMIC HAZARD MAP.
NEW MADRID SEISMIC ZONE -
THREE INTERPRETATIONS OF
A POSSIBLE FAULT LOCATION
TO ACCOUNT FOR HISTORICAL
SEISMICITY. THE MIDDLE FAULT
LOCATION IS CONSIDERED TO
BE TWICE AS LIKELY AS THE
OUTER TWO IN THE 2002 UPDATE
FOR THE NATIONAL SEISMIC
HAZARD MAP.
FIGURE 7.20 QUATERNARY FAULTS AND SIGNIFICANT SEISMIC SOURCES WITH
OVERLAY OF MAIN BITUMINOUS COAL FIELDS
FIGURE 7.20 QUATERNARY FAULTS AND SIGNIFICANT SEISMIC SOURCES WITH
OVERLAY OF MAIN BITUMINOUS COAL FIELDS
< PREVIOUS VIEW
7-71
Seismic Design: Stability and Deformation Analyses
MAY 2009
stances. Factors to consider in establishing the size of MDE or SEE are the hazard
potential classifcation of the dam (FEMA 2004a), criticality of the project function
(water supply, recreation, food control, protection of the environment, etc.), and
the turnaround time to restore the facility to operability. In general, the associated
performance requirement for the MDE or SEE is that the project performs without
catastrophic failure (such as uncontrolled release of a reservoir) although signifcant
damage or economic loss may be experienced. If the dam contains a critical water
supply reservoir, the expected damage should be limited to an extent that allows the
project to be restored to operation within an acceptable timeframe.
Operating Basis Earthquake (OBE) The OBE is an earthquake that produces ground
motions at the site that can reasonably be expected to occur within the service life
of the project. The associated performance requirement is that the project functions
with litle or no damage and without interruption of function. The purpose of the
OBE is to protect against economic losses from damage or loss of service. Therefore,
the return period for the OBE may be based on economic considerations.
Coal refuse impoundments and other mining dams ofen have signifcant or high hazard potential,
and recommendations for selection of the MDE and seismic hazard assessment are presented in Table
7.6. To protect site operational personnel, the public and the environment, it is necessary to design
coal refuse impoundments having high hazard potential so that catastrophic failure does not occur.
This is normally accomplished by considering the MCE when developing the design earthquake.
The MCE might be represented by more than one earthquake event (i.e., from near-feld and far-
feld sources) for which the structure should fail catastrophically. For coal refuse impoundments and
dams having signifcant hazard potential, some federal dam safety agencies are giving consideration
to events with return periods of the order of 2,500 years for design earthquakes. Further guidance
on return-period criteria for signifcant- and high-hazard-potential dams is anticipated in the future
from federal dam safety agencies.
Although the OBE does not typically govern the safe design of a coal refuse impoundment, it is still
recommended that an OBE be defned on the basis of its probability of occurrence during the life
of the dam. A 500-year (or more precisely 475-year) return period is consistent with the basis for
defning an OBE in some conventional building codes and corresponds to one of the hazard maps
published by the USGS (corresponding to a 10-percent probability of non-exceedance event for a
50-year lifetime), as shown on Figure 7.18. The OBE earthquake parameters should be used to check
the reasonableness of the design MCE. It is expected that the MCE will be a remote event with a PGA
substantially higher than the OBE.
7.7.1 Analytical Procedures
Two approaches, probabilistic seismic hazard analysis (PSHA) and deterministic seismic hazard
analysis (DSHA), are commonly applied to estimation of earthquake ground motions at a site. Fun-
damentally, only a DSHA can be used to estimate the Maximum Credible Earthquake (MCE). In
PSHA, an MDE is associated with a return period far beyond the life span of the structure and is
associated with a very low probability of occurrence.
7.7.1.1 Deterministic Seismic Hazard Analysis (DSHA)
The basic steps in a DSHA are as follows:
1. Develop a seismotectonic model. Review and summarize the basic geology and
tectonics of a region surrounding the site (approximately 322-km radius) with par-
ticular atention to specifc seismic sources, both area sources (provinces) and linear
< PREVIOUS VIEW
7-72
Chapter 7
MAY 2009
sources (faults). Note that a 332-km (200-mile) potential radius of infuence has been
chosen in order to account for the susceptibility of fne tailings to liquefaction efects
from large, distant earthquakes.
2. Conduct a review of the seismic history of the sites region; locate the epicenters of all sig-
nifcant earthquakes and plot these earthquakes on a map that shows magnitude values.
3. Relate these epicenters to the mapped sources defned in Step 1.
4. Based on the results of Step 3, postulate a group of conceivable MCEs by selecting
the most severe earthquake within or along each source that could be larger than the
largest historical earthquake and move each earthquake to the point on the respec-
tive source nearest to the site.
5. Derive the ground motion at the site in terms of PGA on the basis of regionally ap-
plicable atenuation functions (i.e., relationships that allow the estimation of ground
motion parameters at the site as a function of earthquake magnitude, source-to-site
distance, and soil conditions at the site).
6. From the various PGAs obtained in Step 5 for the group of conceivable MCEs from
Step 4 (applying the minimum distances from source-to-site), determine the most
severe ground motion and classify this motion as the MCE. In the case of dams and
embankments, the most severe earthquake event is not simply a function of the
site-specifc PGA, so it might not be apparent that a single event controls. A distant,
TABLE 7.6 MDE AND SEISMIC HAZARD ASSESSMENT RECOMMENDATIONS
Dam Hazard
Potential
Classifcation
Maximum Design
Earthquake
(MDE)
Seismic Hazard Assessment
Low-Seismic-Hazard-Potential
Area
(1)
Moderate- to High-Seismic-
Hazard-Potential Area
(1)
High Hazard
Potential
MCE
(2)
Simplifed Minimum EQ and
Simplifed Design Ground Motion
(3)
or
Site-specifc DSHA
(2)
supported by PSHA
(4)
Site-specifc DSHA
(2)
supported by PSHA
(4)
Signifcant Hazard
Potential
MDE ~ 2500 return
period
(5)
USGS EQ Hazard Maps
or
Site-specifc PSHA
USGS EQ Hazard Maps
or
Site-specifc PSHA
Note: 1. Low-seismic-hazard-potential areas, as defned in this Manual, are distinguished using the USGS
Earthquake Hazard Maps for a return period of 2,500 years with a PGA 0.1g and M 5.5. All other
areas are considered moderate- to high-seismic-hazard-potential areas.
2. The Maximum Credible Earthquake (MCE) derived on the basis of a DSHA, as described in Section
7.7.1.1, is recommended as the MDE for high-hazard-potential dams based on FEMA (2005b). If a
PSHA is used to establish the MDE for a high-hazard-potential facility, the designer should be able
to demonstrate that the probability of occurrence is very remote (the actual return period may be a
function of the seismic hazard area and tectonic conditions).
3. See Section 7.7.3.7 for Simplifed Design Ground Motion and Simplifed Minimum Earthquake (EQ)
description.
4. The results of a Probabilistic Seismic Hazard Assessment (PSHA) can aid in demonstrating that
the probability of catastrophic failure is very remote and in deciding whether mean, or mean-plus-
one standard deviation or greater estimates of ground motion, would be justifed for the MDE from
the deterministic ground motion analysis to achieve an acceptably low probability of exceedance
(FEMA, 2005b).
5. For a signifcant-hazard-potential facility, the designer should be able to demonstrate that the prob-
ability of catastrophic failure is remote. The actual return period may be a function of the seismic-
hazard area and tectonic conditions (e.g., consideration of a return period of the order of 2,500
years may be reasonable for low-seismic-hazard areas).
< PREVIOUS VIEW
7-73
Seismic Design: Stability and Deformation Analyses
MAY 2009
large-magnitude earthquake might not be associated with a PGA as high as for a
smaller magnitude, nearer event, but such a distant earthquake could pose a greater
threat because of its greater energy content and potential to induce more cycles of
signifcant vibrations around the natural damped frequency of vibration of the struc-
ture-foundation system. Consequently, an examination of representative ground
motions for more than one event might be necessary (Step 7).
7. Associate the MCE ground motion(s) to time history(ies) of ground motion consis-
tent with the magnitude of the source and, with lesser emphasis, the calculated PGA
from that source.
The ground conditions for which the selected atenuation functions in Step 5 are applicable require
consideration when determining where on or in the ground the derived ground motion should be
applied in site-specifc dynamic analyses of the structure-foundation system. For example, some
atenuation functions are only applicable to site conditions characterized by frm rock and thin,
very stif soil. Hence, the derived ground motion at a site with less stif or thicker soil overburden
should be applied at the top of rock, rather than at the ground surface. Depending on the site condi-
tions, this derivation of ground motion at the top of relatively sof ground could be accomplished
using the SHAKE program (Schnabel et al., 1972) or other similar approach as further discussed in
Section 7.4.2.2.1 and 7.5.
7.7.1.2 Probabilistic Seismic Hazard Analysis (PSHA)
A PSHA requires the same basic input as the DSHA in terms of the defnition of seismotectonic models
and atenuation functions to characterize site ground motion. However, the results from PSHA and
DSHA are fundamentally diferent. PSHA addresses the chance of a given level of ground motion
being exceeded from all possible earthquakes.
The methodology used for the PSHA is well established in the literature (Cornell, 1968; McGuire,
1976; McGuire, 1978). PSHA is the methodology used by the USGS to derive the seismic hazard of the
United States, and details of the procedures followed are described by Frankel et al. (1996). Calculation
of the seismic hazard requires specifcation of the same inputs as required for the DSHA, except with
the additional requirement that the uncertainties associated with each input need to be quantifed:
Source geometry the geographic description of the seismic sources in the region of
infuence around the site. A seismic source is a portion of the earth associated with
a tectonic fault or (if individual faults cannot be identifed) with an area/province of
homogeneous seismicity. Source geometry determines the probability distribution of
distance R from the earthquake to the site f
R
(r).
Seismicity the rate of occurrence and the magnitude distribution f
M
(m) of earth-
quakes within each seismic source.
Atenuation functions the defnition of atenuation functions is the same as with the
DSHA, except that it is important to defne the uncertainty of each possible function.
One consideration related to uncertainties with each input is how to weight the events within a
seismic source and weight the infuence of a source. The USGS report, Documentation for the 2002
Update of the National Seismic Hazard Maps, prepared by Frankel et al. (2002), explains the data,
evaluations, judgments and assumptions behind their PSHAs for diferent regions.
Mathematically, the calculation of whether the annual probability Y > y the ground-motion char-
acteristic parameter Y , expressed in terms of peak ground acceleration PGA or spectral acceleration
S
a
for various frequencies of vibration, exceeds level y involves the summation, over seismic sources
that may pose a threat to the site, of the following relationship:
< PREVIOUS VIEW
7-74
Chapter 7
MAY 2009
N
r
max
m
u
Y > y =
i {
P[Y>y
|
m,r]
f
M,R
(m,r) dm dr
}
i
(7-7)
i=1
r
min m
l
where:
N = number of the nearby seismic sources, including both line sources (i.e., active
faults) and area sources
i
= annual mean rate of occurrence of earthquakes generated by source i with
magnitudes between m
l
and m
u
r
min
= minimum source-to-site distance for source i
r
max
= maximum source-to-site distance for source i
m
l
= lower-bound values of magnitude for source i
m
u
= upper-bound values of magnitude for source i
and:
P[Y>y
|
m,r] is the conditional probability that the site may be struck by a ground
motion with parameter Y > y generated by an earthquake of magnitude M = m with
epicenter in source i at distance R = r from the site.
f
M,R
(m,r) is the joint probability density function of magnitude M and distance R for
source i
Hence the expression within brackets in Equation 7-7 is the probability that at the site an earthquake
with parameter Y > y will occur because a seismic event of magnitude m (m
l
m m
u
) originated
anywhere in source i.
Ground motion at the site is modelled by an atenuation function, which is a mathematical equa-
tion that defnes the relationship between the ground-motion parameter Y, earthquake magnitude,
source-to-site distance and local soil condition at the site. The atenuation function is used to compute
the conditional probability in the integrand of Equation 7-7. Note that the USGS hazard maps are for
a frm-rock site condition, where the shear-wave velocity averaged over the top 30 meters (V
s30
) is
760 meters per second (boundary of NEHRP site classes B and C). On a practical level, this means
that the USGS has not tried to model site-specifc conditions. If there are site-specifc conditions that
need to be accounted for, an extra step in the analysis is required.
If a scalar parameter such as peak ground acceleration (PGA) is chosen, the repetition of these calcu-
lations for diferent values of the threshold level y allows the construction of the seismic hazard curve
for the site (a plot of annual probability of exceedance, or return period, versus the parameter level).
If the parameter of the ground motion is expressed in terms of spectral acceleration, S
a
(f, ), a family
of seismic hazard curves for the site can be evaluated for various values of f and ( f is the frequency
and is the percentage of the critical damping). These results based on spectral ordinates can also be
rearranged in order to obtain Uniform Hazard Spectra (UHS) (McGuire, 1974), which represent the
response spectra whose ordinates are associated with the same probabilities of exceedance at the site.
These methods that make use of spectral quantities include those that employ only a high-frequency
parameter such as PGA.
7.7.1.3 DSHA versus PSHA
Regardless of whether seismic hazard is defned by a DSHA or PSHA methodology, the results still
have some degree of uncertainty. The advantage of PSHA is that it can incorporate a range of uncer-
< PREVIOUS VIEW
7-75
Seismic Design: Stability and Deformation Analyses
MAY 2009
tainties inherent in the seismotectonic model, occurrence frequency, and ground-motion atenuation
relationships, that is not possible with a DSHA approach. However, PSHA has some signifcant limi-
tations, especially in terms of selecting ground motion parameters that can be considered representa-
tive of a maximum credible event (MCE). A discussion of the limitations of the PSHA approach for
defning the seismic input for critical facilities is presented by Krinitzsky (1995).
Selection of the design ground motion parameters from a PSHA is complex because there are infnite
points (choices) on the hazard curve. It is not practical to pick a point from a hazard curve and be
able to relate it to the ground motion representing a reasonable worst case. Furthermore, the ground
motion derived from PSHA does not have a clear physical meaning (Wang et al., 2003), because
the total hazard (total annual frequency of exceedance) at a site is the sum of the individual haz-
ards (annual frequency of exceedance) and is not associated with any individual earthquake, but
potentially many earthquakes. For example, on the 2002 USGS national seismic hazard maps, the
total seismic hazard in Chicago, Illinois was derived from a series of earthquakes with magnitudes
ranging from 5.0 to 8.0 at distances ranging from 0 to 500 km (Harmsen et al., 1999). Therefore, it can
be difcult to identify from the results of a PSHA a specifc earthquake that represents a worst case
scenario.
Although the earthquake magnitude and PGA are not physically related, the USGS Earthquake Haz-
ards Program now provides probability mapping for earthquake magnitude (Moment Magnitude M
within 50 km of a site as discussed in Section 7.7.2.5) as well as PGA, which allows for defnition of a
local earthquake event by M and PGA for a selected return period or probability. (The USGS provides
maps of PGA for return periods up to a 2,475-year event, although the PGA for a 4,975-year return
period can be computed on the basis of available site-specifc deaggregations.) The use of the USGS
hazard mapping to arrive at these parameters does not provide any specifc insight to the controlling
source and earthquake event, or a representative ground motion, but should provide a reasonable set
of design parameters for lower seismic hazard sites for a given return period or probability, as each
variable would be statistically represented by the maximum mean value. Then actual ground motion
records for locations within 50 km of past earthquakes of magnitude M not less than the design M
can be adopted in design, in lieu of atempting to identify the controlling source earthquakes and to
derive atenuated ground motions from each of those events.
The primary advantage of the DSHA approach in terms of defning the seismic hazard for a structure
like a dam or embankment is that it deduces a particular seismic scenario, consisting of the postu-
lated occurrence of an earthquake of a specifed size at a specifed location upon which a ground-
motion-hazard evaluation is based. A DSHA approach provides ground-motion parameters from
those earthquakes that have the most signifcant impacts. This is particularly advantageous for seis-
mic design and analysis of dams and embankments.
As noted previously, the design basis for high-hazard-potential dams and impounding embankments
is generally the MCE. Fundamentally, probability is not a consideration for the MCE, except as veri-
fcation that the MCE is an extremely remote event. Previous designs for some coal refuse impound-
ments in some areas of the northern Appalachian and Illinois Basins have associated a 10,000-year
return period with a MCE event, absent a detailed site-specifc study or the availability of a detailed
study from a similarly critical nearby structure/facility. This return period is at the upper end of typi-
cal recommendations (USCOLD, 1999). Table 7.6 presents recommendations for the seismic hazard
assessment (and design earthquake), and the subsequent sections of this chapter describe develop-
ment of the design ground motion based on a DSHA, which yields an earthquake magnitude, peak
horizontal ground acceleration, and associated time histories/response spectra. The procedures pre-
sented herein include verifcation that the design earthquake is an extremely remote event, through
comparison with published return periods or results of a PSHA. If a designer elects to base the design
for a high-hazard-potential dam on a PSHA, the return period should be established such that the
< PREVIOUS VIEW
7-76
Chapter 7
MAY 2009
MDE can be equated to the controlling MCE, as cited in FEMA (2005b). Further PSHA guidance on
return-period criteria for signifcant- and high-hazard-potential dams is anticipated in the future
from federal dam safety agencies.
7.7.2 Seismotectonic Modeling
Most seismotectonic studies and publications focus on the western U.S. where there is a relative
abundance of data and observations. The seismotectonic modeling of the central and eastern U.S.
requires that the diferences between the eastern and western U.S. be appreciated. The following
characteristics of eastern U.S. earthquakes represent signifcant diferences when compared to the
western U.S.
More than an order of magnitude lower seismicity rates (longer recurrence periods
for the same magnitudes).
General lack of surface faulting, such that it is difcult to defne source models
Slower atenuation of ground motions with distance, which implies larger areas of
damage for the same earthquake magnitude.
Higher high-frequency content of seismic ground motions to larger distances.
Relatively higher site amplifcation where sof soil is over rock (considered to be
more signifcant in glaciated areas where the contact with highly competent rock is
abrupt; Youd et al., 2001).
Greater uncertainty in quantitative hazard assessments because the historic seismic-
ity record (300 years) is too short compared to the recurrence periods of major dam-
aging events.
Few sets of eastern strong-motion data exist. They marginally constrain the atenu-
ation of shaking with distance, as well as the dependence of local ground motion on
magnitude, distance and depth of the earthquake.
Higher frequency content of eastern earthquakes may lead to a greater number of
cycles for the same magnitude.
In spite of the diferences between eastern and western U.S. seismic events and the limitations of
the eastern U.S. database, these factors can be accounted for in a seismic hazard analysis. The high-
frequency motions associated with eastern earthquakes are generally limited to near-feld rock sites.
These motions tend to atenuate rapidly when they propagate through soil, which efectively reduces
the amount of energy and the number of pertinent ground motion cycles that could contribute to the
hazards of seismically-induced strength loss or ground deformation. Therefore, as noted by Youd et
al. (2001), the duration diferences between eastern and western soil sites are not likely to be signif-
cant when evaluating strength loss and ground deformations.
Compared with a plate boundary environment like California, an intra-plate environment like the
central and eastern U.S. is difcult to characterize. Although there have been many recent advances
in the understanding of earthquake occurrence based mainly on discoveries of paleoliquefaction phe-
nomena, many gaps still exist in the knowledge of why and how ofen earthquakes occur in the cen-
tral and eastern U.S., as summarized by the USGS Geologic Hazards Team (Crone and Wheeler, 2000)
that supports the National Earthquake Hazard Reduction Program (NEHRP).
Defning a seismotectonic model for the central and eastern U.S. requires considerable judgment and
is a subject of ongoing research. It is anticipated that it will be necessary to review the status of ongo-
ing research whenever new seismic hazard assessments are required. This Manual includes some of
the concepts and sources of information suitable for the derivation of seismic hazard at the time of
publication, as summarized in the following subsections.
< PREVIOUS VIEW
7-77
Seismic Design: Stability and Deformation Analyses
MAY 2009
7.7.2.1 Regional Tectonic Framework
One of the major accomplishments with respect to the defnition of seismic hazard in the central and
eastern U.S. over the past 20 years is the defnition of potential seismic sources on the basis of tectonic
evidence. The zones where there has been Quaternary fault movement are summarized in Figure
7.20. Fault sources are the dominant means for deriving seismic hazard in the western U.S. Fault
sources are poorly defned in the central and eastern U.S., but recent work has identifed a tectonic
framework for earthquake occurrence, as summarized in Crone and Wheeler (2000). Recent data for
the New Madrid Seismic Zone (NMSZ) is presented in Tutle et al. (2002).
In terms of conducting a deterministic analysis, there is no agreement regarding the actual defnition
of the main seismic sources afecting the central and eastern U.S. Nevertheless, it is practical to use
the sources that are currently defned by the USGS Geologic Hazards Team. The zones where Qua-
ternary tectonic movements have been documented are predominantly in the New Madrid Seismic
Zone, and the modeling of this zone has been considered to be as fault zones of varying probability,
as shown in Figure 7.20. This zone has signifcant infuence to the southern Illinois Basin, as docu-
mented in the probabilistic hazard map shown in Figure 7.18 for a 500-year return period and in
Figure 7.19 for a 2,475-year return period.
Another relatively recent discovery is the presence of paleoliquefaction in southern Illinois and Indi-
ana (Figure 7.20), which was identifed by Crone and Wheeler (2000) as the Wabash Seismic Zone.
These widespread paleoliquefaction efects have been documented and interpreted to be associated
with a large earthquake (M~ 7.5) possibly centered in the Wabash Valley area between southern Illi-
nois and Indiana that occurred about 6,100 years before the present (BP). Specifc faults that pro-
duced the paleoliquefaction features have not yet been identifed.
Within the Appalachian Basin coal felds, the most signifcant source is the Eastern Tennessee Seis-
mic Zone, a belt of seismicity in northeastern Alabama, northwestern Georgia and much of eastern
Tennessee (Figure 7.20). The largest historical shock was magnitude 4.6, and occurred in 1973. No
evidence for larger prehistoric shocks has been discovered, yet the microearthquake data suggest
coherent stress accumulation within a large volume (Chapman et al., 2002). The largest earthquake
within the Appalachian system had a Modifed Mercali Intensity (MMI) equal to VIII (M = 5.9) and
occurred on May 31, 1897 in Giles County, Virginia in the general area where evidence of Quaternary
faulting has been uncovered. This faulting is referred to as the Pembroke Fault zone by Crone and
Wheeler (2000). There is no surface expression of these faults and it is uncertain if they are due to
tectonic movements or from solution collapse.
The third signifcant earthquake source in the eastern U.S. is the Charleston, South Carolina Seis-
mic Zone. Paleoliquefaction data indicate that for characteristic large earthquakes in the Charleston,
South Carolina region, the recurrence interval is approximately 550 years, as presented by Talwani
and Schaefer (2001). The USGS (Frankel et al., 2002) have assigned two source zones for this area
with an equal probability of their validity. Although a signifcant seismic source in the eastern U.S.,
the efects of this zone do not signifcantly afect the Appalachian coal felds.
The defnition of seismic source zones in the central and eastern U.S. based strictly on tectonics is still
very much an evolving science. In conducting a deterministic analysis it will be necessary to carefully
evaluate the most recent work of the USGS and other organizations, especially universities, involved
with ongoing research into the identifcation of seismogenic tectonic structures. For example, the
Mid-America Earthquake Center (MAE Center) is a good source of information for the central and
eastern U.S. seismic hazard. The MAE Center, headquartered at the University of Illinois at Urbana-
Champaign, is a consortium of nine core institutions funded by the NSF and each core university, as
well as through joint collaborative projects with industry and other afliations.
< PREVIOUS VIEW
7-78
Chapter 7
MAY 2009
7.7.2.2 Historical Seismicity
The seismic hazard derived for locations in the central and eastern U.S. needs to be at least as
severe as has been historically recorded. The National Earthquake Information Center of the U.S.
Geological Survey and the Engineering Research and Development Center of the U.S. Army Corps
of Engineers compile source and magnitude information for earthquake events. These sources
also provide access to isoseismal maps of historical events and available strong motion records,
although few are available for the central and eastern U.S. The seismic hazard evaluation process
should consider the historical maximum Modifed Mercalli Intensity experienced at the site, and
the design ground motion should not be less than that corresponding to the maximum MMI. The
correlation of MMI to PGA is not an exact science, but correlations developed for California such
as developed by Wald et al. (1999) and adapted for use in the central U.S. by Atkinson and Kaka
(2006) are worth considering.
7.7.2.3 Selection of Seismic Sources
To defne the source zones for deriving seismic hazard, the distribution of earthquakes needs to be
evaluated in the context of available information regarding Quaternary tectonic activity. In the west-
ern U.S., careful evaluation of surface and near-surface faults should be conducted in association
with a review of historical seismicity, which is ofen of limited quality. It may be necessary to conduct
site-specifc studies of faults whose rupture could control the seismic design in order to determine
their degree of activity and their physical characteristics, as summarized in detail by Slemmons and
DePolo (1986). Such studies are seldom practical in the central and eastern U.S.
In many cases in the central and eastern U.S., earthquake distribution by itself is the basis for defn-
ing a seismic source. For example, the Eastern Tennessee Seismic Zone is defned primarily based on
the distribution of instrumentally determined epicenters. The overall active zone in the New Madrid
area has been similarly defned.
The seismic sources most critical to the central and eastern U.S. coal felds are those identifed in
Figure 7.20. The fault sources that are signifcant to the defnition of seismic hazard in the western
U.S. coalfelds are too numerous to describe in this Manual, and the reader is referred to lists and
discussions provided by the USGS or state geological surveys. The interpretation presented in
Figure 7.20, which represents the current best interpretation of the USGS, and forms the basis for
the National Seismic Hazard Maps (Frankel et al., 2002), examples of which are shown in Figures
7.18 and 7.19 , illustrates the current uncertainty with respect to the main sources afecting central
and eastern U.S. coal felds. It should be emphasized that the presence of earthquakes is not an
absolute criterion for delineating a seismic source zone. For example, the Wabash Valley Seismic
Zone shown in Figure 7.20 was developed based on the identifcation of paleoliquefaction phe-
nomena. Other than for special areas of relatively high seismic hazard, the approach of the USGS
to defning seismic hazard has been to construct hazard maps directly from the historic seismicity
data following the procedure in Frankel (1995). The number of events greater than the minimum
magnitude are counted on a grid with spacing of 0.1 in latitude and longitude. Accordingly, to
develop the seismic hazard following a DSHA, it is necessary to develop source models more fully
than the main sources defned by the USGS.
In addition to the USGS, there has been considerable efort by other organizations to delineate seismic
sources in the central and eastern U.S. In particular, from 1981 to 1989, Lawrence Livermore National
Laboratory (LLNL) developed a PSHA methodology for the eastern United States (Bernreuter et al,
1989), followed in 1993 by improvements in the handling of the uncertainties (USNRC, 1993). Difer-
ences between these results and those of a utilities-sponsored study (Electric Power Research Insti-
tute, 1989) led to the formation of the Senior Seismic Hazard Analysis Commitee (SSHAC) to identify
the sources of diferences and give guidance on how to perform a state-of the-art PSHA (USNRC,
< PREVIOUS VIEW
7-79
Seismic Design: Stability and Deformation Analyses
MAY 2009
1997). This work continues to be reviewed and updated (Savy et al., 2002). These documents provide
diferent expert opinions regarding appropriate source models for the central and eastern U.S., and
they are important sources for detailed information regarding the seismotectonics of the central and
eastern U.S. Nevertheless, it is cautioned that the USGS modeling is not one of the source models
considered, and that some of the experts involved in development of the seismic source models for
the LLNL studies hold widely divergent interpretations, as documented in the review of the seismic
hazard at USDOE sites (USDOE, 1996).
Another source of seismotectonic modeling information is the U.S. Army Corps of Engineers (USACE).
The USACE seismic design manual, Response Spectra and Seismic Analysis for Concrete Hydraulic Struc-
tures (USACE, 1999), provides a detailed seismotectonic interpretation of the central U.S. in terms of
a source model based on a comprehensive review of seismotectonic data, and it is a notable reference
for the defnition of seismic source zones in this area.
7.7.2.4 Maximum Magnitude
In either a deterministic or probabilistic analysis, it is necessary to identify the strongest earthquake
that has occurred within each source and then determine if there is a reason to consider that a larger
earthquake could reasonably be expected to occur. Substantial judgment is required for this assess-
ment. For example, it may be reasonable to assume that the 1811-1812 earthquakes in the New Madrid
Seismic Zone are a maximum event. However, the known historical seismicity in southern Illinois
does not encompass the largest earthquake that could be produced in that area, given the widespread
paleoliquefaction phenomena that have been mapped.
In an intraplate environment, the evaluation of maximum magnitude is especially problematic
because the source mechanisms are poorly defned. Where source characteristics are well defned,
such as along a mapped fault, common practice is to use relationships developed between earthquake
magnitude and total fault length, rupture area, maximum displacement per event, fault slip rate and
seismic moment (Slemmons and Depolo, 1986; Wells and Coppersmith, 1994 or several others). With
some exceptions such as the Meers Fault in Oklahoma, these relationships can be applied only with
considerable judgment in the central and eastern U.S.
DSHA is intended to defne the worst-case ground motion and should therefore be based on the least
favorable combination of earthquake source characteristics and location, and the strongest ground
motion that could be generated by this scenario. In practice, DSHA generally uses the logarithmic
mean or mean-plus-one-standard-deviation level of ground motion from predictive equations (Krin-
itzsky, 2002). If a mean plus a standard deviation is used, the ground motion will be conservative 84
percent of the time. A more detailed discussion of the difculties in defning maximum magnitude is
summarized by Bommer et al. (2004). Because of the uncertainties in the defnition of maximum mag-
nitude, the best approach is one of targeting a consensus opinion, such as is provided by the USGS
Geologic Hazards Team (Frankel et al., 2002).
Another potential source of information regarding maximum earthquake magnitude is to consider
probability of occurrence. In addition to the PGA hazard maps, the USGS (Frankel et al., 2002) also
provides maps of the probability of occurrence of earthquakes of various magnitudes within specifc
distances of a given location as a function of return period. An example of a map depicting the prob-
abilities (in terms of return periods) that an earthquake with M > 5.5 will occur within a distance of
50 km of diferent locations within the central and eastern U.S. is presented in Figure 7.21. As this
map was prepared by considering earthquake recurrence with a return period up to 10,000 years, the
results show that large earthquakes are only reasonably conceivable in limited areas of the central
and eastern U.S. and, conversely, that earthquakes of M > 5.5 are a remote possibility in most areas
of the central and eastern U.S.
< PREVIOUS VIEW
7-80
Chapter 7
MAY 2009
FIGURE 7.21 RETURN PERIOD FOR M > 5.5 EVENT AT
A DISTANCE LESS THAN 50 KM
LEGEND
RETURN PERIOD IN YEARS FOR AN
EARTHQUAKE OF M > 5.5 AT A DISTANCE
OF LESS THAN 50 KM.
9999
7500
4000
3000
2000
1000
500
300
NOTE: 1. IMAGE REDRAWN FROM 2002 MAP
BY THE USGS OF RETURN PERIOD
FOR AN M > 5.5 EVENT AT A DISTANCE
OF LESS THAN 50 KM.
2. ANTHRACITE AND BITUMINOUS
COAL FIELDS ARE DENOTED BY
RED BORDERS.
FIGURE 7.21 RETURN PERIOD FOR M > 5.5 EVENT AT A DISTANCE LESS THAN 50 KM
FIGURE 7.21 RETURN PERIOD FOR M > 5.5 EVENT AT A DISTANCE LESS THAN 50 KM
< PREVIOUS VIEW
7-81
Seismic Design: Stability and Deformation Analyses
MAY 2009
7.7.2.5 Earthquake Recurrence
The evaluation of earthquake recurrence for each source considered is a requirement of the PSHA.
Recurrence is of interest to a DSHA if the distribution of earthquakes within a source as determined
from the historical seismicity and geologic data as appropriate is suggestive that a maximum magni-
tude other than the historic maximum should be considered.
The determination of earthquake recurrence requires several steps:
1. Identifcation of both historical and instrumentally-located earthquakes within each
postulated source.
2. Conversion of the available data into a common magnitude base. There are several
magnitude scales, but for about the past 25 years it has been common practice to
defne magnitude in terms of seismic moment, as defned by Kanamori (1977) and
Hanks and Kanamori (1979). Moment magnitude is the scale most commonly used
for engineering applications and is the scale preferred for calculation of liquefac-
tion resistance. The relationship between moment magnitude M and other com-
monly used magnitude scales can be reviewed as a chart published by Youd et
al. (2001). In the central and eastern U.S., the most commonly derived magnitude
scale is m
blg
, which was developed by Nutli (1973). Frankel et al. (1996) provide
the following conversion to moment magnitude:
M = 3.45 0.473 m
blg
+ 0.145 (m
blg
)
2
(7-8)
3. Evaluate the completeness of the earthquake catalog afer the removal of statistically
dependent events afershocks and foreshocks. Determination of catalog complete-
ness is a function of location (e.g., earthquake history is beter defned for a longer
period in Charleston, South Carolina than in central Montana). Procedures for evalu-
ating completeness are presented by Veneziano and Van Dyck (1985).
4. Formulate earthquake recurrence for each source. The most common means of defn-
ing recurrence is the Richter formula (Richter, 1958):
log N = a b m (7-9)
where:
m = earthquake magnitude
N = number of earthquakes of magnitude m or greater per year per unit area
a, b = constant coefcients defning a linear relationship between log N and m
This simple exponential recurrence model has been found to have good applicability for area
sources, but there are many other ways to defne the probability of earthquake recurrence, particu-
larly when specifc faults are considered (Schwartz and Coppersmith, 1986). It should also be noted
that the simple Richter formula can be non-linear for the highest magnitudes and that caution
should be used when estimating the recurrence of large magnitude events on the basis of limited
historical data.
7.7.3 Design Ground Motion
The basic seismic data needed for determination of an MDE, including the MCE, is a time history(ies)
representative of the source magnitude of the earthquake and a duration (number of cycles of strong
ground motion) consistent with the source magnitude. The derivation of an MCE for a high-hazard-
< PREVIOUS VIEW
7-82
Chapter 7
MAY 2009
potential dam involves a seismic-hazard assessment, and this Manual describes procedures for per-
forming a seismic-hazard assessment using a DSHA with comparison to published probabilistic
hazard assessments, such as those available from the USGS or the USNRC. The MCE should be shown
to be an extremely remote event. FEMA (2005b) indicates that a PSHA can aid in deciding whether
mean, or mean-plus-one-standard-deviation, or even greater estimates of ground motion are justi-
fed in a deterministic ground motion analysis for the MDE to achieve an acceptably low probability
of exceedance. Until published guidance from federal dam safety agencies is available, the designer
should consider that for high-hazard-potential facilities in areas of medium to high seismic hazard,
the primary role of a PSHA is to serve as technical support in the selection of an MCE consistent with
FEMA (2005b).
The OBE, as well as the MDE for signifcant-hazard-potential dams, can be defned on the basis of a
probabilistic analysis, which could simply be obtained from published results of the USGS. As noted
in the introduction to this chapter, a probabilistic analysis considering a 500-year return period can
serve as the basis for defning an OBE, and a return period of the order of 2,500 years may be reason-
able for establishing the MDE for a signifcant-hazard-potential dam.
Another means for evaluation of the results of a DSHA is to review the design basis for nuclear
power plants, which have all been designed on the basis of a DSHA. The location of nuclear power
plants with respect to U.S. coal felds is presented in Figure 7.22. Table 7.7 provides the Safe Shut-
down Earthquake (SSE) peak horizontal ground accelerations (PGA) that formed the basis for seismic
design for each nuclear plant. The SSE PGA values for these nuclear plants should be considered as
minimum values for a MCE. The conservatism associated with the seismic input for a nuclear power
plant has proven to be the overall spectral shape of their seismic input, rather than the PGA. Further-
more, earthquake magnitude values associated with the SSE are generally not defned, because the
practice for defning seismic hazard in the 1960s and 1970s, when these nuclear power plants were
licensed, was to base the design on Modifed Mercalli intensity (MMI) relationships.
As noted in Section 7.7.2, much of the areas mined for coal in the central and eastern U.S. have
a low seismic hazard. For these areas a simplifed approach is recommended for defning a mini-
mum standard for seismic design. Section 7.7.3.7 of this Manual proposes standard seismic inputs for
design in areas of low seismic hazard, and these inputs are believed to be reasonably conservative for
coal refuse impoundments. If the designer judges that less conservative inputs than recommended
herein are appropriate, it is acceptable to perform a complete hazard analysis to justify less conserva-
tive project-specifc and/or site-specifc seismic design inputs. The following sections present more
detailed information regarding the defnition of design ground motion for an MCE for a high-hazard-
potential dam.
7.7.3.1 Selection of Design Earthquakes
The MCE for each seismotectonic structure or source area within the region examined needs to be
defned by a common parameter, preferably by moment magnitude M, but in a manner that is com-
patible with the atenuation function used to derive site ground motion. It is recommended that the
maximum historical site MMI be evaluated as an additional estimate to site ground motion, but it is
recommended that the historical MMI be converted to magnitude based on scientifc evaluations of
the ground efects. Such studies are usually available for the most signifcant historical events. For
example, using a new method for evaluating magnitude by directly inverting observations of intensi-
ties, Bakun and Hopper (2004) determined a moment magnitude M of 7.4 (7.0 to 7.7 at a 95 percent con-
fdence level) for the largest New Madrid earthquake in the 1811-12 sequence. Scientifc publications for
evaluation of the magnitude of historical earthquakes are generally available.
Where the earthquake history is incomplete or where there is no geologic evidence regarding past
earthquake activity, judgment is required in assigning a maximum magnitude to each source. Where
< PREVIOUS VIEW
7-83
Seismic Design: Stability and Deformation Analyses
REV. AUG. 2010
fault sources are inferred to be present, fault movement within the range of 35,000 to 100,000 years
BP is considered recent enough to warrant an active or capable classifcation, and they should be
modeled as sources according to their fault dimensions and histories. The assignment of a maximum
magnitude to each source should consider the concepts presented in Section 7.7.2.4.
The MCEs identifed for each source potentially afecting the site are candidates for one or more con-
trolling MCEs at the site. It is also important to look at a variety of earthquakes that have a long duration
and are rich in lower frequency contents, but do not necessarily cause the highest peak acceleration at
the coal refuse site. For coal refuse impoundments, this longer-duration earthquake may be the control-
ling event if it triggers strength loss of the embankment/foundation materials. Seismic input for a coal
refuse impoundment facility may be defned by both near-feld and far-feld events. For purposes of
clarifcation, a near-feld event is an earthquake that is postulated to have an epicenter at or very near a
site of interest and a far-feld event is one where the earthquake is not expected to occur near the site.
7.7.3.2 Ground Motion Attenuation
The diference in ground-motion atenuation between the western and eastern U.S. is signifcant and
needs to be accounted for in the hazard analysis. This is a topic that has been extensively researched
and there are a number of atenuation functions that have been developed for eastern U.S. ground
motion. Examples include Toro et al. (1997), Frankel et al. (1996), Atkinson and Boore (1995, 2006),
Somerville et al. (2001) and Campbell (2003). The atenuation relationships were developed based on
FIGURE 7.22 LOCATION OF NUCLEAR POWER PLANTS
WITH RESPECT TO U.S. COAL FIELDS
NOTE: 1. THE SAFE SHUTDOWN EARTHQUAKE (SSE)
DESIGN PEAK HORIZONTAL GROUND ACCELER-
ATIONS FOR THESE NUCLEAR POWER PLANTS
ARE PROVIDED IN TABLE 7.7.
TURKEY POINT 3,4
ST. LUCIE 1,2
CRYSTAL RIVER
SAN
ONOFRE 2,3
DIABLO
CANYON 1,2
PALO VERDE 1,2,3
COLUMBIA
CALVERT CLIFF 1,2
GINNA
FITZPATRICK
VERMONT
YANKEE
PILGRIM
SEABROOK
NINE MILE
POINT 1,2
OYSTER CREEK
MILLSTONE
2,3
HARRIS
BRUNSWICK 1,2
VOGTLE
INDIAN
POINT 2,3
SOUTH
TEXAS 1,2
COMMANCHE
PEAK 1,2
FARLEY 1,2
McGUIRE 1,2
HOPE CREEK
SALEM 1,2
RIVER BEND
ARKANSAS
NUCLEAR ONE
1,2
GRAND
GULF 1
MONTICELLO
PRAIRIE
ISLAND 1,2
SLURRY 1,2
NORTH
ANNA 1,2
SUMMER 1
FORT CALHOUN 1
COOPER 1
WOLF
CREEK
DUANE
ARNOLD
KEWAUNEE
POINT
BEACH 1,2
LIMERICK 1,2
OCONEE
1,2,3
ROBINSON 2
CATAWBA 1,2
BROWNS
FERRY
SEQUOYAH 1,2
WATTS BAR 1
HATCH 1,2
CLINTON 1
CALLAWAY 1
PERRY 1
FERMI 2
DAVIS
BESSE 1 DC COOK
1,2
PALISADES
BYRON
1,2
QUAD
CITIES 1,2
LASALLE 1,2
DRESDEN
2,3
BRAIDWOOD 1,2
BEAVER
VALLEY 1,2
SUSQUEHANNA 1,2
PEACH BOTTOM 2,3
3 MILE ISLAND 1
2. ANTHRACITE AND BITUMINOUS COAL FIELDS
ARE DENOTED BY RED BORDERS.
FIGURE 7.22 LOCATION OF NUCLEAR POWER PLANTS WITH RESPECT TO U.S. COAL FIELDS
FIGURE 7.22 LOCATION OF NUCLEAR POWER PLANTS WITH RESPECT TO U.S. COAL FIELDS
< PREVIOUS VIEW
7-84
Chapter 7
MAY 2009
numerical modeling and sparse strong-motion records from small earthquakes, and while subject to
uncertainty, they are considered to be appropriate.
7.7.3.3 Selection of Peak Ground Motion Parameters
The MCE is defned as the most severe ground motion calculated from the selected ground motion
atenuation relationship(s) from the various potential MCEs identifed from the possible sources.
The credible severe combination(s) of magnitude (preferably moment magnitude) and distance will
defne the basis for the MCE. The ground motion associated with the credible severe combinations of
magnitude and distance is normally defned in terms of peak ground acceleration (PGA).
In many cases the most severe PGA is associated with a relatively small nearby earthquake, especially in
the areas of relatively low seismic hazard. Where there is moderate to high risk, just deriving the cred-
ible severe PGA from a small local earthquake is not sufcient, and it is necessary to supplement this
MCE with a credible severe earthquake from a relatively distant large magnitude source where the PGA
might be lower than a small local event, but will be associated with more cycles of ground motion.
7.7.3.4 Selection of Acceleration Time Histories
A time-history representation of the seismic input is generally presented in terms of the ground
accelerations (accelerograms) where the variation of ground motion is ploted as a function of time.
Actual strong motion recordings of earthquakes are time histories of the earthquake ground motion.
Ultimately, the goal of selecting a time history for use in design is to be able to match the MDE ground
motion with a PGA and number of cycles of ground motion defned for design. There are two ways
to obtain time histories for engineering design and analysis: (1) actual ground motion records or (2)
synthetic ground motions.
Actual ground motion records can be obtained from the USGS or COSMOS (Consortium of Orga-
nizations for Strong-Motion Observation Systems) web sites. If actual strong motion recordings are
used as seismic input, they need to be selected to match as closely as possible the source magnitude
and expected faulting mechanism, recognizing that faulting mechanisms may not be known if the
site is in the central or eastern U.S. Although distance is a consideration in the selection of the ground
motion, it is more signifcant that the earthquake simulate the PGA derived for the design MDE(s)
and that the duration/number of cycles of ground motion be consistent with the number of cycles
derived with the simplifed approach for liquefaction analysis. In addition, care should be taken to
select records that are well represented in the frequency range of interest to an embankment dam
(approximately 0.5 to 1.5 Hz). In order to efectively bound the range of possible ground motions,
several hypothetical design ground motions, or a synthetic time history rich in all possible frequen-
cies of interest, should be used.
As noted in Section 7.7.3.2, most of the strong motion records available as potential design time
histories in the range of a central and eastern U.S. MCE will be from areas with diferent atenua-
tion characteristics (e.g., the western U.S.). There are very few records from the central and eastern
U.S., especially for moderate to strong earthquakes of magnitude M greater than 5.0. Nevertheless,
as noted in Section 7.7.2, in spite of the diferences associated with western U.S. records and limita-
tions to the eastern U.S. database, these diferences can be accounted for in a seismic hazard analy-
sis because, within soils where strength loss occurs, it is expected that western and eastern ground
motions will be similar (Youd et al., 2001). If actual strong ground motion records are used as the
basis for defning the MCE, it is recommended that at least three records be selected for the analysis
and that justifcation provided for their selection. If the site is in an area of medium to high seismic
hazard, more than three time histories should be used.
Although real time histories are the preferred means for defning ground motion, an alternative
approach for defning a design time history that is ofen used in the central and eastern U.S. is to derive
< PREVIOUS VIEW
7-85
Seismic Design: Stability and Deformation Analyses
MAY 2009
TABLE 7.7 SAFE SHUTDOWN EARTHQUAKE (SSE) PEAK HORIZONTAL GROUND
ACCELERATIONS FOR U.S. NUCLEAR POWER PLANTS
Plant SSE PGA (g) Plant SSE PGA (g)
Arkansas Nuclear 1, 2 0.21 Millstone 2, 3 0.17
Beaver Valley 1, 2 0.13 Monticello 0.13
Braidwood 1, 2 0.19 Nine Mile Point, 1, 2 0.15
Browns Ferry 1, 2, 3 0.21 North Anna, 1, 2 0.13
Brunswick 1 0.17 Oconee 1, 2, 3 0.11
Byron 1, 2 0.19 Oyster Creek 0.27
Callaway 0.19 Palisades 0.19
Calvert Cliffs 1, 2 0.15 Palo Verde 1, 2, 3 0.19
Catawba 1, 2 0.17 Peach Bottom, 2, 3 0.15
Clinton 0.21 Perry 1 0.15
Columbia Generating Station 0.25 Pilgrim 1 0.15
Comanche Peak 1, 2 0.13 Point Beach 1, 2 0.13
Cooper 0.21 Prairie Island 1, 2 0.13
Crystal River 3 0.11 Quad Cities 1, 2 0.27
D.C. Cook 1, 2 0.21 River Bend 1 0.11
Davis-Besse 0.15 Robinson 2 0.19
Diablo Canyon 1, 2 0.80 Saint Lucie 1, 2 0.11
Dresden 2, 3 0.21 Salem 1, 2 0.21
Farley 1, 2 0.11 San Onofre 2, 3 0.63
Fermi 2 0.15 Seabrook 1 0.25
FitzPatrick 0.15 Sequoyah 1, 2 0.17
Fort Calhoun 0.19 South Texas 1, 2 0.11
Ginna 0.19 Summer 0.15
Grand Gulf 1 0.15 Surry 1, 2 0.15
Harris 1 0.15 Susquehanna 1, 2 0.11
Hatch 1, 2 0.15 Three Mile Island 1 0.15
Hope Creek 1 0.19 Turkey Point 3, 4 0.15
Indian Point 2, 3 0.15 Vermont Yankee 0.19
Kewaunee 0.13 Vogtle 1, 2 0.19
La Salle 1, 2 0.21 Waterford 3 0.11
Limerick 1, 2 0.15 Watts Bar 1 0.19
McGuire 1, 2 0.15 Wolf Creek 1 0.13
(ADAPTED FROM SOBEL, 1994)
synthetic ground motions. The most widely used methods to generate synthetic ground motion are
the stochastic point- and fnite-source models (Hanks and McGuire, 1981; Toro and McGuire, 1987)
and the stochastic fnite-fault model, which can simulate some of the near-source efects (Atkinson
and Silva, 1997; Beresnev and Atkinson, 1997). With the improvement of computers, it has become
possible to use more sophisticated numerical methods for simulating strong ground motion based on
empirical or theoretical source functions and two- and three-dimensional wave propagation theory,
which has been successfully used in ground-motion simulations for many earthquakes (Somerville
et al., 2001; Saikia and Somerville, 1997). Zeng et al. (1994) also developed a composite source model
that has been recommended for use in generating synthetic ground motions for seismic design and
analysis of highway bridges in Kentucky.
< PREVIOUS VIEW
7-86
Chapter 7
MAY 2009
7.7.3.5 Applicability of Design Ground Motion
The end result of a seismic hazard assessment is to derive the ground motion that will afect the base
of the impounding structure. Most ground-motion-atenuation relationships derive motion for bed-
rock or stif (frm) soils. Where the site foundation consists of less stif/dense or deeper soil deposits,
or otherwise deviate from the site conditions for which the atenuation relationships are applicable,
the derived ground motion should be applied at the top of rock and the ground motion (at the ground
surface) estimated from local site response analyses based on computer simulation with programs
such as SHAKE, QUAD4M, or DESRA (Schnabel et al., 1972; Finn et al., 1977). For these analyses it is
necessary to use a time history input as discussed in Section 7.5.
7.7.3.6 Application of Seismic Parameters in Design Process
Figure 7.23 presents a fow diagram for seismic hazard assessment to determine the design earth-
quake inputs (M, PGA, ground motion). At a minimum, the design M and PGA are required, as
these data are used in the Youd, et al. (2001) pore-pressure-based triggering analysis, to establish a
minimum number of cycles of loading for cyclic shear strength testing, in very simplifed dynamic
response analyses, and for other general purposes. More rigorous dynamic response, seismic stabil-
ity, and deformation analyses require design ground motions (acceleration versus time histories) to
estimate the amplifcation or de-amplifcation of the ground motion from the appropriate input hori-
zon (e.g., top of rock, top of stif soil, or base of structure), up through the foundation and overlying
structure; calculate peak accelerations at critical locations within the foundation and structure; con-
duct more refned and sophisticated triggering analyses; calculate seismically-induced stresses and
strains; and identify the likely modes and estimate the range in magnitude of deformation.
7.7.3.7 Simplifed Design Ground Motion for Areas of Low Seismic Hazard
As noted in the introduction to Section 7.7.3, a substantial portion of the areas mined for coal in the
central and eastern U.S. have a low seismic hazard. The defnition of what constitutes low seismic
hazard is judgmental, but coal refuse disposal facilities in the area where the Documentation for the
2002 Update of the National Seismic Hazard Map (Frankel et al., 2002) indicates that the PGA with a
return period in 2,500 years is less than or equal to 0.10g can be considered to be representative of
low seismic hazard (Figure 7.19). This zonation is efectively similar to the area where the return
period for an earthquake of M > 5.5 within 50 km of a given location is greater than 10,000 years
(Figure 7.21).
For structures in areas of low seismic hazard that warrant evaluation for a MCE, a simplifed
approach can be used. The simplifed minimum design earthquake should have a moment magni-
tude M of at least 5.5 (M 5.5) at a distance no farther than 50 km from the source. For evaluations
of triggering or deformations requiring both M and PGA, adopt 150 percent of the PGA associated
with a 2,500-year return period based on the USGS PSHA hazard maps, but not less than PGA =
0.10g. When needed, select earthquake ground motion records based on M and the other consid-
erations outlined in Section 7.7.3.4. In areas of low seismic hazard, it is not necessary to consider
more distant earthquakes with M > 5.5.
Commentary: A simplifed approach is recommended by the authors to defne a minimum standard for seismic
design in low-seismic-hazard areas, with inputs believed to be reasonably conservative for coal refuse impound-
ments. It is based on published seismic hazard mapping, subject to periodic updating and widely available from
the USGS. Based on the project site location, a designer can directly determine the M and PGA using the above
procedures for the simplifed minimum design earthquake. If a designer judges that other (or potentially less
conservative) inputs are appropriate, it is acceptable to perform a seismic hazard assessment. As indicated in
Table 7.6, the minimum standard is appropriate for projects only in low-seismic-hazard areas, and a seismic
hazard assessment is necessary in moderate to high seismic hazard areas.
< PREVIOUS VIEW
7-87
Seismic Design: Stability and Deformation Analyses
MAY 2009
Screening of site seismic hazard based on a return period of 2,500 years
1. Determine PGA at site from USGS seismic hazard maps.
2. Determine M (within 50 km) from USGS Earthquake Probability Maps.
FIGURE 7.23 SEISMIC HAZARD ASSESSMENT FLOW DIAGRAM
YES NO
DETERMINATION OF EARTHQUAKE PARAMETERS
IS THE PGA < 0.10g
AND
M < 5.5?
Select 3 actual earthquake ground motion records for
locations within 50 km of an earthquake event with M no
less than 5.5 (or design M, if higher). Potential sources
include USGS and COSMOS. In selecting ground
motions, consider duration of motion, frequency content
and ground/foundation characteristics comparable to
those found in the area of interest.
LOW SEISMIC HAZARD
DETERMINE DESIGN EARTHQUAKE PARAMETERS
MODERATE TO HIGH SEISMIC HAZARD
DESIGN OF NEW HIGH-HAZARD DAMS, EXPANSION OF
EXISTING HIGH-HAZARD DAMS AND ASSESSMENT
OF DESIGN STANDARDS FOR EXISTING HIGH-HAZARD DAMS
Apply PSHA data from USGS or data from available study
for similarly critical nearby structure, previous site-specific
seismic hazard study (whether from PSHA or DSHA) if still
valid, or from a site-specific PSHA or DSHA study.
1. Perform a site-specific DSHA of PSHA.
2. Verify that the PGA from a site-specific DSHA is
an extremely remote event and not less than a Safe
Shutdown Earthquake (SSE) for a nuclear power
plant in a similar seismic zone, if such data are
available. If not, reassess postulated earthquakes
and DSHA evaluations.
3. Verify that the PGA from a site-specific PSHA is an
extremely remote event and can be equated to the
controlling MCE.
NOTE:
1. As support to development of a site-specific DSHA,
alternative approaches for demonstrating the
conservativeness of the DSHA assumptions,
including PSHA are encouraged.
2. Several representative ground motion records
covering the potential range of worst-case ground
motions consistent with the results of the DSHA and
accounting for site-specific ground conditions should
be selected to define the seismic input for new high
hazard dams in an area of medium to high seismic
hazard.
OTHER CASES
Adopt the following simplified minimum design earthquake
for the site:
1. Earthquake Magnitude M > 5.5
2. PGA = 1.5 times PGA for 2500-year return period >
0.10g
Use available earthquake record (see box below) or DSHA
or PSHA study for similar critical or high hazard structure.
SELECT REPRESENTATIVE DESIGN EARTHQUAKE
GROUND MOTION RECORDS
FIGURE 7.23 SEISMIC HAZARD ASSESSMENT FLOW DIAGRAM
FIGURE 7.23 SEISMIC HAZARD ASSESSMENT FLOW DIAGRAM
< PREVIOUS VIEW
7-88
Chapter 7
MAY 2009
Selection of time histories of acceleration for structures in areas of low seismic hazard should con-
sider earthquake records that are refective of the site-specifc design parameters (M and PGA) and
foundation conditions. Atributes of the time histories of acceleration should include: (1) the M and
PGA are at least equivalent to the site-specifc design parameters, (2) near-feld event with hypo-
central distance of 50 kilometers or less, (3) foundation conditions similar to site-specifc conditions
(otherwise, some adjustment of the time history may be warranted), (4) several cycles with a repre-
sentative band of acceleration in proportion to the design PGA, and (5) some cycles at or near the
frequencies of interest for the structure (low frequencies for dams). Section 7.7.3.4 presents reference
sources for ground motion histories. The following time histories illustrate the selection of candidate
records for low-seismic-hazard areas:
Example 1 (Figure 7.24) Time history for low-seismic-hazard area, and the site is
distant from widely recognized seismic source zones. This example illustrates a time
history that can be considered for a site in the eastern or central U.S., with a design
M and PGA of at least 5.5 and 0.1g, respectively. This example has a high PGA (much
greater than the site specifc design PGA) and lacks cycles of substantive acceleration
at lower frequencies, which is not uncommon.
Example 2 (Figure 7.25) Time history for low-seismic-hazard area, but the site is
close to a boundary of moderate seismic hazard (as refected by a site-specifc PGA
much greater than 0.1g). This example illustrates a time history that may be con-
sidered for a site in the eastern or central U.S. (close to a recognized seismic source
zone) with a design M of at least 5.5 and a design PGA much greater than 0.1g (e.g.,
0.12 to 0.15g) with signifcant low frequency content. This example could also be
conservatively applied to sites in low-seismic-hazard areas that are distant from
widely recognized source zones.
Example 3 (Figure 7.26) Time history for low-seismic-hazard area, and the site is
distant from widely recognized seismic source zones. This example illustrates a time
history for an actual eastern North American earthquake that could be considered
for a low-seismic-hazard site in the eastern US.
7.8 SEISMIC DESIGN OVERVIEW
Gardner and Wu (2002) presented an overview of the challenges in evaluating strength loss and seis-
mic stability at coal refuse disposal facilities, including a summary of available pore-pressure-based
empirical methods and strain-based laboratory methods that MSHA has used in their review and
approval of impoundment plans. Wu et al. (2003) cite common MSHA review issues with seismic
stability including: (1) defning materials subject to strength loss, (2) sampling and testing proto-
cols, (3) defning ground motion parameters applicable to the site, (4) determining the appropriate
margin of safety, (5) and dealing with questions concerning the applicability of available techni-
cal information for fne coal refuse. There are no documented case histories of a fne coal refuse
impoundment being afected by signifcant earthquake loading, and as a result, there is litle direct
confrmation of methods that have been employed for estimating strength loss. Thus, MSHA has
been faced with reviewing impoundment design plans with minimal factors of safety, ofen with-
out mitigating design features that would enhance seismic stability such as provisions for internal
drainage of fne refuse deposits.
This chapter has presented credible, documented methods and procedures for the evaluation of seis-
mic design including stability and deformation analyses considered applicable for materials encoun-
tered at coal refuse disposal facilities. A fow chart illustrating the steps involved in evaluating seismic
design, stability and deformation was presented in Figures 7.1a, 7.1b and 7.1c, and Section 7.1.2 pro-
vides a discussion of the basis for the design guidance, including:
< PREVIOUS VIEW
7-89
Seismic Design: Stability and Deformation Analyses
MAY 2009
WHITTIER NARROWS - STATION: 24399 MT. WILSON - CIT SEISMIC STATION
OCTOBER 1, 1987 M = 5.99, DISTANCE FROM FAULT RUPTURE = 21.2 KM, V = 822 M/SEC (ROCK)
7.24a OVERALL RECORD
7.24b SAMPLE TIME SEGMENT
ATTRIBUTES: 1. M AND PGA > SITE-SPECIFIC DESIGN M AND PGA
2. NEARER FIELD, FAULT-DRIVEN EVENT WITHIN 50 KM
3. FIRM ROCK COMPARABLE TO PROJECT SITE CONDITIONS
4. MANY CYCLES WITHIN REPRESENTATIVE BAND OF PGA.
5. LIMITED CYCLES AT LOWER FREQUENCY.
FIGURE 7.24 EXAMPLE 1: TIME HISTORY FOR LOW SEISMIC HAZARD AREA
P
G
A
(
F
R
A
C
T
I
O
N
O
F
g
)
TIME (SEC)
-0.25
-0.20
-0.15
-0.10
-0.05
0.00
0.05
0.10
0.15
0.20
0 5 10 15 20 25 30 35 40
TIME (SEC)
P
G
A
(
F
R
A
C
T
I
O
N
O
F
g
)
-0.25
-0.20
-0.15
-0.10
-0.05
0.00
0.05
0.10
0.15
0.20
1.5 2.5 3.5 4.5 5.5 6.5
NOTE: SITE IS DISTANT FROM WIDELY RECOGNIZED SEISMIC SOURCE ZONES
(UNIV. OF CAL., 2007a)
FIGURE 7.24 EXAMPLE 1: TIME HISTORY FOR LOW-SEISMIC-HAZARD AREA
FIGURE 7.24 EXAMPLE 1: TIME HISTORY FOR LOW-SEISMIC-HAZARD AREA
< PREVIOUS VIEW
7-90
Chapter 7
MAY 2009
FIGURE 7.25 EXAMPLE 2: TIME HISTORY FOR LOW SEISMIC HAZARD AREA
7.25a OVERALL RECORD
TIME (SEC)
P
G
A
(
F
R
A
C
T
I
O
N
O
F
g
)
-0.50
-0.40
-0.30
-0.20
-0.10
0.00
0.10
0.20
0.30
0.40
0 5 10 15 20 25
7.25b SAMPLE TIME SEGMENT
TIME (SEC)
P
G
A
(
F
R
A
C
T
I
O
N
O
F
g
)
-0.5
-0.4
-0.3
-0.2
-0.1
0.0
0.1
0.2
0.3
0.4
2 3 4 5 6 7
COYOTE LAKE - GILROY ARROY #6, 230 (CDMG STATION 57383)
AUGUST 6, 1979 M = 5.7, HYPOCENTRAL DISTANCE = 9 km, V = 663 m/sec (SOFT ROCK)
ATTRIBUTES: 1. M AND PGA > SITE-SPECIFIC DESIGN M AND PGA
2. NEAR FIELD EVENT MUCH CLOSER THAN 50 km
3. SOFT ROCK COMPARABLE TO PROJECT SITE CONDITIONS
4. INCLUDES PERTINENT LOW FREQUENCY CYCLES.
5. SIGNIFICANT ACCELERATION SPIKES AT LOWER FREQUENCY.
NOTE: SITE IS CLOSE TO THE BOUNDARY OF MODERATE SEISMIC HAZARD
(UNIV. OF CAL., 2007b)
FIGURE 7.25 EXAMPLE 2: TIME HISTORY FOR LOW-SEISMIC-HAZARD AREA
FIGURE 7.25 EXAMPLE 2: TIME HISTORY FOR LOW-SEISMIC-HAZARD AREA
< PREVIOUS VIEW
7-91
Seismic Design: Stability and Deformation Analyses
MAY 2009
7.26b SAMPLE TIME SEGMENT
FIGURE 7.26 EXAMPLE 3: TIME HISTORY FOR LOW SEISMIC HAZARD AREA
7.26a OVERALL RECORD
TIME (SEC)
P
G
A
(
F
R
A
C
T
I
O
N
O
F
g
)
-0.15
-0.10
-0.05
0.00
0.05
0.10
0.15
0 5 10 15 20 25 30 35
TIME (SEC)
P
G
A
(
F
R
A
C
T
I
O
N
O
F
g
)
-0.15
-0.10
-0.05
0.00
0.05
0.10
0.15
10 11 12 13 15 14
SAGUENAY - W. CHICOUTIMI NORD (SITE 16 T)
NOVEMBER 25, 1988 M = 5.9, HYPOCENTRAL DISTANCE = 50 KM (HARD ROCK)
ATTRIBUTES: 1. M AND PGA > SITE-SPECIFIC DESIGN M AND PGA
2. EASTERN NORTH AMERICA
3. ROCK RECORD
4. MANY CYCLES WITHIN REPRESENTATIVE BAND OF PGA
5. RELATIVELY LONG DURATION OF STRONG SHAKING > 0.05g (~12 SEC)
NOTE: EASTERN U.S. EARTHQUAKE, DISTANT FROM WIDELY RECOGNIZED SEISMIC SOURCE ZONES
(UNIV. OF CAL., 2007c)
FIGURE 7.26 EXAMPLE 3: TIME HISTORY FOR LOW-SEISMIC-HAZARD AREA
FIGURE 7.26 EXAMPLE 3: TIME HISTORY FOR LOW-SEISMIC-HAZARD AREA
< PREVIOUS VIEW
7-92
Chapter 7
MAY 2009
Appropriate levels of analysis, depending on the type of facility.
Methods for identifying and evaluating material susceptibility to strength loss,
including available feld and laboratory techniques.
Simplifed methods for estimating post-earthquake strength, as well as more sophis-
ticated methods for evaluating if triggering of strength loss occurs.
Alternatives for evaluating seismicity depending on the level of seismic hazard of
the region.
A recommended factor of safety for seismic stability of 1.2 based on a static stabil-
ity analysis using post-earthquake strengths, which also helps achieve designs with
predicted deformations within acceptable limits.
Section 7.1.5 provides recognized, simplifed steps and procedures that utilize basic site data and
component material properties for development of conservative embankment dam designs in regions
of low seismic hazard that constitute most of the areas of coal mining, as discussed in Section 7.7.
More sophisticated analyses and procedures may be needed for some facilities built by upstream
construction, and these are also discussed in Section 7.1.5, with additional details presented in Sec-
tion 7.4. These more sophisticated analyses and procedures can reduce conservatism and potentially
provide the experienced designer with more expertise additional options for design. These types of
analyses may be needed in those limited regions where seismic hazard is moderate to high. Basic
screening methods for evaluating deformations are discussed in Section 7.1.5, and the applicability
of both screening and analysis procedures is presented in Section 7.5.
Other credible methods and procedures may be acceptable or may become accepted in practice, pro-
vided that documentation and testing support their application, and these methods should not be
precluded from use in seismic design.
< PREVIOUS VIEW
Appendix 7A
DERIVATIONS OF BASIC EQUATIONS FOR
STEADY-STATE LABORATORY TESTING
7A1 MAY 2009
7A.1 COMPARISON OF AND ENVELOPES AT STEADY STATE
a
3
1
b
3
1
MOHR'S CIRCLE AT STEADY STATE
ENVELOPE DRAWN THROUGH MAXIMUM SHEAR
STRESS ON MOHR'S CIRCLE
ENVELOPE DRAWN TANGENT TO MOHR'S CIRCLE
2
MEASURED IN TRIAXIAL TEST AT STEADY STATE DEFINE p =
1 +
3
c
c
2
MEASURED IN TRIAXIAL TEST AT STEADY STATE DEFINE q =
3
a = b = q
tan =
a
c
sin =
b
c
THEREFORE sin = tan
c = p =
3
+ q
a
=
b
= sin
-1
q
p
AND
< PREVIOUS VIEW
7A2
Appendix 7A
MAY 2009
7A.2 RELATIONSHIP BETWEEN S
US
AND
3
AT STEADY STATE
b
a
3
3 + q 1
MOHR'S CIRCLE AT STEADY STATE
b = q =
1 - 3
2
= RADIUS OF MOHR'S CIRCLE
a = b cos = SUS
THEREFORE: SUS
= q cos
FROM TRIANGLE GEOMETRY: sin =
b
3
+ q
=
q
3
+ q
sin =
3
+
=
cos
cos
cos SUS
3
+
cos
3
sin + sin =
cos
3
sin = SUS (1 - sin
THEREFORE:
SUS
SUS
SUS
SUS
SUS
SUS =
3
1 - sin
cos sin
SUS
q
< PREVIOUS VIEW
Appendix 7B
VOID-RATIO MEASUREMENTS DURING UNDISTURBED
TUBE SAMPLING AND LABORATORY TESTING
7B1 MAY 2009
7B.1 INTRODUCTION
The purpose of this appendix is to present methods for accurate measurement of the void ratio of
triaxial specimens and the changes in void ratio of undisturbed tube samples from conditions in situ
to conditions during shear in the triaxial test. Accurate void ratio measurements are needed in order
to correct the undrained steady-state (residual) strengths measured in the triaxial test to obtain in-
situ strengths using the procedures described in Poulos et al. (1985). Specifcally, the required void-
ratio values are:
Undisturbed specimens both the void ratio in situ and the void ratio during
undrained shear in the triaxial test must be determined. To obtain these values one
measures the void ratio changes that occur at various stages of the sampling and
testing process, as subsequently discussed.
Remolded specimens (tested to obtain the slope of the steady-state line) the actual
void ratio in the triaxial cell during shear is needed. For drained tests, the void ratio
changes during shear, and the value when the specimen reaches steady state is
needed.
In the discussion that follows, undisturbed specimens are addressed frst, and then the diferences
in the procedures for remolded specimens are presented. The procedures discussed below include
some measure ments that are not directly needed in the Poulos et al. (1985) procedure. However,
these extra measurements provide important checks of the measurements that are directly needed.
These procedures relate to soils ranging from slightly silty sands to slightly clayey silts. Note that,
in this appendix, the term sample is used to refer to an entire tube sample obtained from the
feld. The term specimen refers to a portion of the sample trimmed from a tube or prepared in
the laboratory.
7B.2 USE OF AVERAGE VOID RATIO
In all cases it is assumed that the average void ratio of the entire specimen during shear is representative
of the strengths and pore-pressure measurements being made in the triaxial test. If localized shear
straining develops, as evidenced by shear planes or localized bulging, this assumption is no longer
valid because the average void ratio of the specimen is no longer representative of the soil being
sheared. Tests with shear planes or localized bulging cannot be used unless localized measurements
of void ratio are carried out using esoteric techniques such as impregnation with resins followed by
microscopic examination.
Similarly, tests on specimens with signifcant stratifcation within the specimen (e.g., layers of slightly
silty sand alternating with layers of slightly clayey silt) cannot be used, because the average void
ratio is not representative of either soil type. Past experience with coal refuse indicates that careful
< PREVIOUS VIEW
7B2
Appendix 7B
MAY 2009
selection of sampling zones based on CPT data has resulted in stratifcation not being an issue and
not causing concentration of deformations in any zone within the test specimen.
7B.3 UNDISTURBED TUBE SAMPLING
Sampling in borings can be accomplished using a thin-wall tube (3-inch O.D. with wall thickness
of one-sixteenth inch) with either a fxed-piston sampler with piston rods extending to the ground
surface (Hvorslev sampler) or a hydraulically-actuated piston sampler (Gus or Osterberg sampler).
Hand-carved samples can be obtained in test pits using a tripod tube sampler. Detailed sampling
procedures for the Hvorslev sampler are presented in Appendix 7C. Similar procedures should be
used for other samplers.
High quality galvanized-steel, brass, or stainless-steel tubes should be used. The cuting edge should
have a clearance ratio of between 0.5 and 1.0 percent (Appendix 7C) and should be fled sharp and
smooth immediately prior to use.
Measurements required during sampling are:
Sampler penetration into the material (to the nearest millimeter).
Length of recovered sample (to the nearest millimeter).
Inside diameter of the sample tube cuting edge (to nearest 0.1 millimeter, using calipers).
Inside diameter of the sample tube (to nearest 0.1 millimeter, using calipers).
The volume change during sampling is computed from:
In-situ sample volume = distance pushed times tube area at cuting edge
In-tube sample volume = sample length recovered times inside area of tube
When measuring the length recovered, it is expected that the botom of the sample is fush with the
botom of the tube. If a section of the botom of the sample is missing, it can be assumed that it fell
during withdrawal of the sampler from the borehole and thus should be considered to be part of the
measured length recovered.
There are potential errors in this part of the procedure. In general, the errors overstate any decrease
in volume. For example, the sample may be fush with the botom, but may have slid downwards.
The shorter measured recovery would be assumed to be a volume decrease with an associated
decrease in void ratio, which is conservative in the overall estimation of in-situ strengths. However,
very large apparent void ratio decreases should be considered unreasonable, and the sample should
not be used for testing.
Experience indicates that it is possible to keep volume changes during sampling generally below
one percent, with the change being usually in compression, but some slightly clayey silts may actually
expand slightly during sampling.
Afer the length of recovered sample is measured, a section of the sample at the botom should be
removed to allow installation of the packer. The tube should be maintained in an upright position
at all times. The packer may have drain holes to allow drainage in sands, but should be solid for
soils exhibiting any plasticity. The top should be cleaned and a loose packer or other disk placed
on top of the sample. The distance from the top of the tube to the disk should then be accurately
measured. This measurement should later be rechecked to determine if the sample setled during
transport.
< PREVIOUS VIEW
7B3
Void-Ratio Measurements During Undisturbed Tube Sampling and Laboratory Testing
MAY 2009
7B.4 HANDLING AND TRANSPORTATION
Sample tubes must kept upright at all stages of handling and transportation. The tubes should be
transported in private vehicles and cushioning should be provided around each tube and between
the tube rack and the vehicle to minimize the transmission of vibrations from the vehicle to the
samples. Freezing of the samples must be avoided. Any volume change that occurs during handling
and transportation can be computed as the change in vertical distance from the top of the tube to the
top of the disk resting on the sample. Volume changes that occur during transportation and handling
are normally negligible even for sand, if reasonable care is taken.
7B.5 TUBE CUTTING AND EXTRUSION
Tube cuting must be performed using a method that does not produce vibration or deformations
of the sample tube. Typically a tube cuter is used. Low pressure is applied to the tube, and stifener
rings are placed above and below the cut to minimize tube deformation. The tube must be maintained
in a vertical position. The distance to the top of the sample will indicate if volume changes have
occurred. If the described procedure is carefully followed, the volume changes during tube cuting
should be negligible.
Afer the tube section containing the triaxial test specimen is cut, the ends of the specimen are typically
trimmed away from each end of the tube section by about one-half inch, so that: (1) the ends of the
tube section can be deburred and (2) a packer can be installed in the tube to keep the sample in place.
The following measurements should be made:
Total weight of the sample and tube (this is a back-up measurement)
Length of tube section length of the tube minus the distance from each end of
tube to the ends of specimen (all measurements to the nearest 0.5 millimeter; each
measurement should be the average of three points taken around the circumference
of the tube).
Inside diameter of the tube section (to the nearest 0.1 millimeter, using calipers).
The volume of the specimen in the sample tube afer cuting and trimming can be computed from the
above measurements.
From this point on, all soil in the tube section must be recovered and the dry weight measured.
This requires that extreme care be taken because: (1) some soil will stick to the sample tube
during extrusion, (2) some will stick to the membrane afer triaxial shear, and (3) some will be
used for index testing afer the triaxial shear testing is performed. The total dry weight of soil,
encompassing all of the above, and the measured volume of the specimen should be used to
compute the void ratio.
The specimen must be extruded upward from the tube, i.e., in the same direction that it entered the
tube during sampling. During extrusion, any soil that stuck to the tube and any other soil that for
whatever reason did not make it to the triaxial cell must be recovered.
Depending on the ability of the soil sample to stand vertically, it may need to be extruded directly into
a close-tolerance membrane stretcher using a membrane that is then used to transfer the sample to the
triaxial cell pedestal. A small vacuum may need to be applied to the sample to complete placement
of a second membrane and atachment of the drainage lines to the top cap. Membrane thicknesses
should be measured prior to use. Afer the sample is safely in the triaxial cell under a small vacuum,
measurements of sample length and diameter should be made, with the later requiring correction for
membrane thickness. Note that these are backup measurements.
< PREVIOUS VIEW
7B4
Appendix 7B
MAY 2009
7B.6 SATURATION
Once the triaxial cell is assembled, a cell pressure must be applied at the same time that the vacuum
is released so that the efective stresses in the specimen remain about constant. As the vacuum is
released, some water may enter the specimen as air bubbles are reduced in size, but this does not
mean that there was a volume change.
Backpressure should then be applied keeping the efective confning stress constant. Note that the
process must be done very slowly so that: (1) the backpressure is equalized at all times throughout
the sample and (2) the pressure regulators are keeping up with the required pressures. The water that
enters the specimen in this process does not correspond to a volume change but to compression of
air bubbles in the voids of the soil and eventually their solution into the pore water. For undisturbed
samples obtained below the water table, the volume change due to saturation should be negligible.
For remolded specimens prepared at low water contents, there may be a collapse due to saturation
and a substantial volume change, particularly for silts or very silty sands.
7B.7 CONSOLIDATION
Volume change during consolidation of the saturated specimens can be accurately determined from
the water expelled from the samples. If the loading piston is atached to the top cap, height change
measurements can also be made accurately, provided that an axial load is applied to the piston to
compensate for the upward force in the piston caused by the cell pressure (in order to maintain the
desired consolidation stress ratio typically one). It is advisable to consolidate the sample in stages
so that a compression curve (e versus log p) can be obtained. Comparison of compression curves
among the various tests can be used as a means of judging unusual results. The changes in void ratio
during the consolidation phase are by far larger than for any of the other stages including sampling,
transportation, and extrusion.
7B.8 SHEAR AND DISASSEMBLY OF TRIAXIAL CELL
Since undisturbed samples are usually sheared undrained, there is no volume change during shear
of saturated samples.
The procedure for disassembling the triaxial cell and removing the sample at the end of a test
should allow for two key measurements, namely the water content and the dry weight of the full
specimen. The specimen should be sliced in half longitudinally, and any stratifcation should be
noted. A vertical slice should be taken for specifc gravity measurement. For coal tailings, a specifc
gravity measurement should be made for every test specimen. For natural soil, 2 or 3 measurements
should be adequate for a given soil layer. Other vertical slices may be taken for Aterberg-limits,
grain-size, or hydrometer tests, but at least half of the specimen should be used for the water
content measurement. The wet weight of the water content specimen should be measured quickly,
before it begins to dry.
As mentioned above, all soil from the triaxial specimen must be recovered so that the full dry weight
can be determined. This includes the soil used for index testing, soil stuck to the membrane or end
caps, and any other soil trimmings. The wet weight of the water content specimen should be measured
to the nearest 0.01 gram. The dry weight of the water content specimen and of all other material
from the triaxial specimen and from the specimen removed during tube trimming should also be
measured to the nearest 0.01 gram.
To make the test specimen frmer during disassembly of the triaxial apparatus, it may be desirable
to reconsolidate the specimen at the end of the test, but, if this is done, the amount of water expelled
during reconsolidation must be recorded. The water expelled during reconsolidation must be included
in the computation of the water content of the specimen during shear.
< PREVIOUS VIEW
7B5
Void-Ratio Measurements During Undisturbed Tube Sampling and Laboratory Testing
MAY 2009
7B.9 COMPUTATIONS OF VOID RATIO
Void ratio computations are discussed in the following text and are summarized in Table 7B.1. The
following notation is used in the table:
W
s
dry weight
D diameter
H height or length
V volume
w water content
e void ratio
unit weight
w
unit weight of water
G
s
specifc gravity of soil grains
incremental change
Stages of sampling and testing are noted by the following:
is in situ
t in tube afer sampling
t in tube afer transportation
ts in tube section afer cuting the sampling tube and trimming the ends
of the test specimen
txi in triaxial cell initial condition under small vacuum
txc in triaxial cell afer consolidation and during undrained shear
The required values of void ratio are:
Void ratio during shear in the triaxial cell
Void ratio in situ
The void ratio during triaxial shear is computed simply from the measured water content of the
specimen at the end of the test, using the relationship G
s
w = e (assuming 100 percent saturation). As
mentioned previously, for coal tailings a specifc gravity measurement should be made for every
test specimen. For natural soil, 2 or 3 specifc-gravity measurements should be adequate for a given
soil layer.
Sampling, transportation and tube cuting can result in changes in void ratio from the in-situ condition
that can be determined directly from the measured changes in volume, using the formula:
e/(1 + e) = V/V
Changes in void ratio should be averaged for the full tube sample. Subsequent to tube cuting, volume
changes and resulting void ratio changes are associated with the tube section from which the triaxial
specimen is extruded.
The table at the end of this appendix lists the measured and computed quantities considered to be
the primary values, the formulas used, and the secondary values that can be used to compare with
the primary measured or computed values. The primary measured and computed quantities should
< PREVIOUS VIEW
7B6
Appendix 7B
MAY 2009
TABLE 7B.1 VOID RATIO MEASUREMENT FOR UNDISTURBED TUBE SAMPLES
(1)
Stage Volume (V)
Dry Weight
(W
s
)
Void
Ratio
(e)
Comments
1. In situ
Change V (meas.)
e(is)
e (2)
e(is) = e(t) + e (during sampling)
Measure V from V (sampled) versus V (recovered).
Compute e.
2. Full tube after
sampling
Change V (meas.)
e(t)
e (2)
e(t) = e(tt) + e (during transport)
Measure V from the height change of the sample
during transport between the feld and the laboratory
3. Full tube after
transport, but
before cutting
Change V
e(tt)
e (2)
e(tt) = e(ts) + e (during cutting)
Measure V from the height change of the sample
measured during tube cutting. V is normally zero.
Compute e (normally zero).
4. Tube section
after cutting
and trimming
Change
V(ts) W
s
(ts)
W
s
e(ts) (3)
Measure V(ts) from the I.D. of the tube and sample
length. Compute e(ts) for V(ts) and W
s
(ts). From this
point on, all solids should be saved so that W
s
(ts) can
be measured. The sample in the tube section should
also be weighed. This can later be used to estimate
the water content and saturation of the sample at this
stage.
W
s
is the soil left in the tube during extrusion.
5. In triaxial cell,
following ex-
trusion from
tube
Change
V(txi) (meas.)
V(txi) (comp.)
V
W
s
(txi)
= W
s
(txi)
e(txi) (3)
V(txi) (meas.) is from measurements of triaxial
sample when set up in the cell under small vacuum.
V(txi) (comp.) is from V(txc) and V during
consolidation. Large differences between the two
indicate a potential problem.
V is measured during consolidation.
6. In triaxial cell
after consolid
ation and during
undrained shear
V(txc) (3,4)
(Not neces-
sarily needed,
but good
practice.)
W
s
(txc)
w(txc)
(Computation
starts after
above are
measured.)
e(txc) (4)
Use w and G
s
after test (txc) to compute consolidated
as-tested void ratio, considering that the sample is
saturated at this point. This is the primary method
for obtaining e(txc). Also measure W
s
for the entire
sample. W
s
is needed for computation of e(ts) in
Stage 4. W
s
may also be used to obtain V(txc). V(txc)
and V measured during consolidation can then be
used as a check on V(txi).
Note: 1. Measurement of V are made from Stage 1 onward, as sampling and testing proceed. The arrows indicate
the order of void ratio computations. Computations begin following measurement of W
s
(txc) and w(txc) in
Stage 6. Computations proceed along the path of the arrows.
2. e = (1 + e) V/ V
3. e = (G
s
w
V/W
s
) 1
V = W
s
(1 + e) / (G
s
w
)
4. e = G
s
w (w is in percent, assumed to be 100 corresponding to complete saturation.)
be used for estimating in-situ strengths. However, there may be cases where comparison of primary
and secondary quantities may indicate that the secondary quantities are more believable and should
be used. Some measurements mentioned in the text are not included in the table, but they can also be
used to compare to the primary values.
< PREVIOUS VIEW
7B7
Void-Ratio Measurements During Undisturbed Tube Sampling and Laboratory Testing
MAY 2009
As previously discussed, consolidation should be performed in stages. However, for clarity,
consolidation is treated as a one-stage process in the table.
7B.10 REMOLDED SAMPLES
The purpose of testing remolded samples prepared at various void ratios from a uniform batch of
soil is to determine the steady-state line for the soil batch, the slope of which is needed for correcting
the results from testing of the undisturbed samples. Thus, primarily, we are interested in the void
ratio during shear, i.e., the last line in Table 7B.1. However, in planning the tests one targets specifc
afer-consolidation void ratios and confning pressures at which to perform tests so as to achieve
sufciently wide coverage of void ratio and stresses in the steady-state plot. For this purpose, it is
important to develop the compression curves for each test (by consolidating in stages) and also to
estimate void ratio changes, if any, that would be expected upon saturation.
Assuming that the samples are prepared in a mold of known volume, the process of computing void
ratios at various stages is similar to the one shown in the table for undisturbed samples, except that
it starts at the tube section afer trimming stage, which is equivalent to an in the compaction mold
stage for the remolded samples.
Ofen when samples are compacted at relatively low water contents, there is substantial compression
upon saturation. Thus the volume change from the txi to the txc stage includes not only the efect of
consolidation, but also the efect of saturation. Since saturation would occur under backpressure, it
would occur with the cell assembled, and thus there would not be access to the sample to measure its
height and diameter afer saturation but prior to consolidation. An alternative procedure would be to
cause near-saturation by circulating de-aired water very slowly through the specimen with vacuum
applied at the upper end while slightly opening the lower valve to allow the de-aired water to be
sucked in, but not opening the valve so much that the vacuum is lost at the lower end of the sample.
This process will not result in full saturation, but it will cause collapse and thus allow measurement
of the sample prior to cell assembly and the application of cell and backpressures.
< PREVIOUS VIEW
Appendix 7C
PROCEDURE FOR UNDISTURBED, FIXED-PISTON
SAMPLING OF COHESIONLESS SOIL
7C1 MAY 2009
The procedures presented herein have been developed for a fxed-piston sampler with piston rods (actu-
ating rods) that extend up to the ground surface (ofen referred to as a Hvorslev sampler). Similar proce-
dures should be employed for sampling using hydraulically-actuated samplers or tripod-tube samplers.
7C.1 ADVANCING THE HOLE
7C.1.1 General
1. The hole should be started as close to vertical as possible. This is necessary because the
actuating rods will be placed through a bracket 10 feet directly above the borehole. If the
hole is not vertical, the actuating rods will bend. Also, when the drill rods are pushed,
they may slip of the drill head if they are not vertical.
2. The drill head should be almost all the way forward when starting the hole. This allows for
maximum horizontal travel of the head. Usually, drillers will do this as routine procedure.
7C.1.2 Casing
1. It is best to use 4-inch-I.D. casing. If larger-diameter casing is used, the velocity of the
drill fuid return is reduced, which increases the amount of coarse wash in the borehole.
If the casing is smaller, the sampler will be a tight ft. Either fush-joint casing or hollow-
stem augers can be used depending on soil conditions.
2. In general, the casing should be advanced to within about 1 foot of the proposed sam-
pling depth. However, this may not be necessary if thick drilling fuid is used and the
borehole remains open below the casing. This is referred to as open-hole drilling. If
Revert drilling fuid is used in open-hole drilling, it is prudent to advance the casing to
the botom of the borehole at the end of the work shif.
7C.1.3 Drilling Fluid
1. It is generally best to use a drilling fuid that has a high unit weight, such as bentonite
mixed with water, rather than plain water. This will help to carry cutings to the surface
and will increase the pressure on the botom of the sample during withdrawal from the
borehole. Usually, the governing factor in preparing the drill fuid is the ability of the
pumping equipment to circulate the fuid through the drill rods. It is advantageous to
use a drill fuid that has a slippery feel.
2. When drilling through a dam, the potential for hydraulic fracturing at the botom of the
borehole should be evaluated. If hydraulic fracturing is possible, special drilling proce-
dures should be used. These procedures may involve maintaining the fuid level in the
borehole well below the ground surface and using augers to advance the boring. Tube
samples can be obtained through large hollow-stem augers, or small-diameter solid
augers may be used to clean out the hole inside casing.
< PREVIOUS VIEW
7C2
Appendix 7C
MAY 2009
3. Several types of drilling fuids are available. Some of these include Revert, Quick Gel,
and Kim-Mud. Revert is an organic material that decomposes a few days afer mixing
with water. It is commonly used if piezometers or a groundwater observation well are
to be installed in the borehole. All three of these muds have worked well for removing
sand-sized particles from boreholes.
7C.1.4 Mud Tub
It is important to have a mud tub with bafes that create setling basins for the cutings. A mud tub
with three or more setling basins is preferred. Cutings should be shoveled out of the mud tub at
frequent intervals to keep the drill fuid clean.
7C.1.5 Hole Advancement
1. Drill bits must have defectors that prevent drill fuid from jeting downward and dis-
turbing the material below the drill bit. Tricone roller bits are available with a defector
atached to one of the cones that prevents drill fuid from jeting directly downward. It
may be necessary to weld a bead on fshtail or chopping bits to defect the drill fuid.
Some bits have side or upward discharge of the drill fuid. The bit should be checked by
fushing water through it above the ground surface.
2. Drill bits should be slightly smaller than the I.D. of the casing. If 4-inch casing is used,
the drill rods should be N or NW size to reduce the annular space between the casing
and drill rods. This will increase the upward velocity of cutings and will help to reduce
wash at the botom of the borehole.
3. Advancement of the bit should be done at a slow rate of about 1 foot per minute maxi-
mum. Drill fuid pressure during drilling should be kept low to minimize disturbance
to the material. Afer the required depth is reached, the bit should slowly be lifed a dis-
tance of about 3 feet above the botom of the borehole. Then the drill fuid pressure can
be raised to wash cutings out of the hole.
4. The drill bit and rods should never be allowed to rest on the botom of the borehole.
They should always be suspended from the drill rig. This is important so that potentially
loose material is not densifed or disturbed by the weight of the drill rods.
7C.2 PREPARATION OF SAMPLER
7C.2.1 Defnition of Sampler Parts
TABLE 7C.1 STATIONARY PISTON SAMPLER WITH STEEL TUBE
Size Sample Tube Length
Rod
Conn.
Part No.
Weight
(lb) (kg)
2 O.D. x 1 I.D.
(1)
(50.8 x 47.6 mm)
30
762.0 mm
AW 22056-16 25.0 11.3
2 O.D. x 2 I.D.
(63.5 x 60.3 mm)
30
762.0 mm
AW 22053-16 28.0 12.6
3 O.D. x 2 I.D.
(1)
(76.2 x 73.0 mm)
30
762.0 mm
NW 22041-43 30.0 13.6
3 O.D. x 3 I.D.
(88.9 x 85.7 mm)
30
762.0 mm
NW 22057-34 32.0 14.5
4 O.D. x 40 I.D.
(113.7 x 110.5 mm)
30
762.0 mm
NW 22065-34 36.0 16.3
Note: 1. Meets ASTM, AASHTO, DCDMA Standards for tubes.
< PREVIOUS VIEW
7C3
Procedure for Undisturbed, Fixed-Piston Sampling of Cohesionless Soil
MAY 2009
7C.2.2 General Operation of Sampler
With the sampler at the botom of the borehole and the actuating rods fxed relative to the ground
surface, the drill rods are pushed down. All of the parts, except the actuating rods and piston, move
down. The locking cone allows one-way (upward) movement of the actuating rods and piston rela-
tive to the sampling tube. Afer the push, the piston cannot move down because the locking cone jams
the actuating rods in place.
7C.2.3 Sampler Tube Preparation
7C.2.3.1 Cutting Edge
1. Before the drilling program begins, the cuting edge of each sampling tube should be
machined to achieve the desired clearance ratio (CR) defned as:
CR =
D
IT
- D
CE
D
CE
where:
D
IT
= inside diameter of tube (should be 2 inch)
D
CE
= inside diameter of cuting edge
Usually, there is less chance of sand sliding out the botom of the tube if the clearance
ratio is relatively high. However, more disturbance of the sand occurs when the clearance
ratio is relatively high because the sand expands outward to meet the sides of the tube.
Typically, sampling tubes obtained from Acker Drilling Company have clearance ratios
of about 1.3 to 1.5 percent. For sand sampling, this clearance ratio is too high. Tubes
should be machined to have clearance ratios between about 0.5 and 1.0 percent.
< PREVIOUS VIEW
7C4
Appendix 7C
MAY 2009
TABLE 7C.2 OPTIONS AND SPARE PARTS STATIONARY PISTON SAMPLER
(1)
No.
Diameter and
Head Thread
Connection
2 O.D. 2 O.D. 3 O.D. 3 O.D. 4 O.D.
AW
Wt.
(lb)
(kg)
AW
Wt.
(lb)
(kg)
NW
Wt.
(lb)
(kg)
NW
Wt.
(lb)
(kg)
NW
Wt.
(lb)
(kg) Name Part No. Part No. Part No. Part No. Part No.
1 Head 120140-9
2.0
0.9
1201140-7
6.0
2.7
120140-10
6.0
2.7
120140-12
8.0
3.6
120140-16
11.0
4.9
2 Clamp Spring 120136 (1) 120136 (1) 120136 (1) 120136 (1) 120136 (1)
3
Cone Clamp
Assembly
220042-1 (1) 22042-1 (1) 22042-1 (1) 22042-1 (1) 22042-1 (1)
4 Plug 120221
3.0
1.3
120189
5.0
2.2
120142
7.0
3.1
120226
10.0
4.5
120243
12.0
5.4
5
Socket Head
Cap Screw
(4 required)
120660 (1) 120652 (1) 120652 (1) 120652 (1) 120652 (1)
6
Packing Cup
(2 required)
150045 (1) 150045-14 (1) 150045-11 (1) 150045-12 (1) 150045-18 (1)
7
Piston Spacer
(2 required)
120222 (1) 120190 (1) 120138
1.0
0.45
120145
1.0
0.45
120297
1.0
0.45
8 Piston 120223 (1) 120295
1.0
0.45
120143
1.0
0.45
120225
1.0
0.45
120242
2.0
0.90
9 Lock washer 90399-04 (1) 90399-065 (1) 90399-08 (1) 90399-10 (1) 90399-15 (1)
10 Locknut 90400-04 (1) 90400-064 (1) 90400-08 (1) 90400-10 (1) 90400-15 (1)
11 Steel Tube 30 120021-4
3.0
1.3
120086-4
4.0
1.8
120037-4
5.0
2.2
120093-11
7.0
3.1
120095-4
7.0
3.1
11
Brass Tube
30
120245-4
3.5
1.5
120246-4
4.00
1.8
120038-4
6.0
2.7
120092-10
6.0
2.7
120094-4
7.0
3.1
11
Stainless Steel
Tube 30
120245-4
3.5
1.5
120246-4
4.0
1.8
120230-1
5.0
2.2
120027-7
6.5
2.9
120244-4
7.0
3.1
12
Piston Rod-
Master
120139-1
1.5
0.67
120139-1
1.5
0.67
120139-1
1.5
0.67
120139-1
1.5
0.67
120139-1
1.5
0.67
(3) Act. Rod 2 ft 120219-2
1.5
0.67
120219-2
1.5
0.67
120219-2
1.5
0.67
120219-2
1.5
0.67
120219-2
1.5
0.67
(3) Act. Rod 5 ft 120219-4
5.0
2.2
120219-4
5.0
2.2
120219-4
5.0
2.2
120219-4
5.0
2.2
120219-4
5.0
2.2
(3)
Act. Rod
10 ft.
120219-5
7.0
3.1
120219-5
7.0
3.1
210219-5
7.0
3.1
210219-5
7.0
3.1
210219-5
7.0
3.1
Note: 1. For optional tubes, see Item 11.
1. Less than one pound or 0.45 kilogram.
2. Not shown.
(REPRODUCED FROM ACKER DRILL COMPANY INC. SOIL SAMPLING TOOLS CATALOG, 1978)
2. Any nicks in the cuting edge should be repaired. A small machinists fle is useful for
this purpose.
3. The cuting edge I.D. should be measured at three to four locations so that an average
cuting edge diameter can be determined. Also, the I.D. and O.D. of the tube should
be checked. They are normally exactly 2 and 3 inches. Field calipers should be used.
Make sure that the calipers are aligned properly.
7C.2.3.2 Deburring and Cleaning
1. The four holes at the top of the tube should be deburred using a fle so that the plug
can be inserted into the top of the tube.
2. Wash all steel flings from the inside of the tube.
< PREVIOUS VIEW
7C5
Procedure for Undisturbed, Fixed-Piston Sampling of Cohesionless Soil
MAY 2009
7C.2.4 Sampler Assembly
7C.2.4.1 Initial Cleaning of Parts
1. Locking Cone should be free of all sand. Ball bearings should rotate freely. Spray with
WD-40 (or similar light lubricant) to lubricate bearings.
2. Spring a new spring should have the ends turned inward so that the ends do not
become jammed between the locking cone and the plug. Clean spring of all sand.
3. Piston Rod make sure that the beveled edge occurs on the threads of the piston rod
only. The bevel should not extend below the threads. The piston rod must be deburred
(fled smooth) above the threads so that the locking cone slides freely along the shaf.
This should be done between each sampling atempt. The female threaded end of the
piston rod should be thoroughly cleaned and lubricated with grease.
4. Plug squirt water through the ports to clean out the sand and mud. Clean the female
threads in the four side holes. Make sure the beveled hole at the top is free of sand and
that the male threads are free of sand.
5. Head clean thoroughly. Make sure there is no sand in the female threads.
6. Piston make sure that sand is removed from the area behind the leathers. Clean the
female threads with water. Make sure that the piston rod screws into the piston very
easily (do not grease these threads). Leathers, which are referred to as packing cups (Part
6 in Table 7c.2), should be kept in water so that they remain pliable. Leathers should not
have major cracks. The leathers should be coated with a thin flm of high vacuum grease.
Initially, the piston and piston rod should be checked to make sure they are able to hold a vacuum
beneath the piston. Put the piston near the botom of a sampling tube with the piston rod screwed in
position and place in a bucket of water. Pull up on the piston rod to check if water can be raised into
the sampling tube.
7C.2.4.2 Assembly
1. Screw the piston rod tightly to the piston.
2. Place the tube horizontally on a table.
3. Push the piston from the top to the botom of the tube. Make sure the piston botom is
fush with the cuting edge.
4. Install the plug and four screws. The screws should be fnger tight.
5. Slide the cone into position. Use a screwdriver to push it in as far as possible.
6. Push hard on the piston rod to make sure the cone is seated. Check that piston has not
moved out of the tube more than one-sixteenth inch.
7. Measure the stick-up of the piston rod above the plug top. This measurement will be
used to determine the actual movement of the piston during sampling.
8. Install a rubber or plastic seal around the actuating rod and push it down over the lock-
ing cone. This seal can be made from a plastic end cap typically used to seal the ends
of the tubes or from an old rubber glove. This seal minimizes the entry of sand into the
cone. Failure of the locking cone has been atributed to sand geting below the cone pre-
venting movement of the cone required for locking the piston rod.
9. Install the spring over the cone.
10. Install the head. Make sure that the spring seats properly into the recess in the head. Put
thick strings between the head and plug and grease the threads. Tighten as tight as pos-
sible by hand. Later, this connection will be broken while soil is in tube. The string and
grease make it easier to break the connection without shaking the tube.
< PREVIOUS VIEW
7C6
Appendix 7C
MAY 2009
11. If an NW-AW sub is used, record the stick-up of the actuating rod above the sub. This
will be used to determine if the actuating rods and piston moved up in the tube while
the tube was being lowered down the borehole.
12. Put the plastic cap over the cuting edge of the tube to protect it.
13. Do not rest the sampler on the piston because the piston can move up into tube.
7C.3 SAMPLING PROCEDURE
7C.3.1 Preparation of Actuating Rods inside Drill Rods
All actuating rod threads should be cleaned and greased.
7C.3.1.1 Separate Drill Stem and Sampling Stem
It is ofen more efcient to use two sets of drill rods (with only one set down the hole at any one
time). One set would have actuating rods inside and be used for sampling only. Using two sets of
drill rods is especially benefcial when the depth of the hole extends beyond 30 feet.
7C.3.1.2 One Drill Stem
1. Afer drilling is completed, actuating rods can be placed down the drill rods and rest on
the roller bit so that the entire string can be lifed out as a unit.
2. Protection for the male actuating rod thread at the botom should be provided.
3. It is important to have a bleeder pipe (2-foot length of rod with holes through the side)
just above the drill bit to allow the escape of drilling fuid from the rods.
4. It is desirable to have a short length of actuating rod at the botom of the actuating rod
string so that the actuating rods stick up above each corresponding drill rod by about 2
to 18 inches.
7C.3.2 Attaching Sampler to Drill Rods
1. Have drill rods ready with the actuating rods inside.
2. Make sure that the bleeder pipe is at the botom of the drill stem.
3. Screw the sub part way onto the botom of the drill rods.
4. Put a vise grip at the botom of the actuating rods and a vise grip on the piston rod.
5. Raise the drill rods high enough to atach the sampler.
6. The driller should hold drill rods so they dont move. They usually extend far above the
drill rig and must be held securely.
7. The helper and inspector atach the tube do not allow the botom of tube to rest on any
surface.
a. Connect the actuating rod to the piston rod with vise grips, making sure the
piston rod is not turned.
b. Unscrew the sub from the drill rods and screw the sub onto the sampler
head. Make sure a string is in the joint.
c. Lif the sampler and the actuating rods that have been atached to the piston
rod, and screw the sampler to the drill rods (this is the hardest part).
8. Remove the plastic cap and check that the piston has not moved. Note how much the
piston is protruding below the cuting edge. This procedure is easier if NW drill rods are
used because then an NW-AW sub is not required. The procedure for ataching the sam-
pler to the drill stem will vary from one project to another and with the available drilling
equipment.
< PREVIOUS VIEW
7C7
Procedure for Undisturbed, Fixed-Piston Sampling of Cohesionless Soil
MAY 2009
7C.3.3 Lowering Sampler down Hole
1. Carefully place the tube into the hole without nicking the cuting edge. If the cuting
edge is nicked, start over.
2. Slowly lower the sampler down the hole. When connecting actuating rods, never
turn the lower actuating rod as this will release the vacuum port in the piston. Mea-
surements of the stick-up of the actuating rods above the drill rods should be made
whenever drill rods are joined as a check on the location of the piston in the sam-
pling tube.
3. Stop lowering sampler about 1 to 2 feet above the botom of the hole.
4. Measure the stick-up on the actuating rods. Is it correct? If the sampler hit a lot of
wash, the actuating rods will be higher. String between joints of drill rods may make
the drill stem slightly longer. Measure the drill rods if a problem is suspected.
5. Fix the actuating rods to the drill rods with a special coupling with a set screw. See
sketch below.
6. Put a keel line just above the coupling.
7. Very slowly, with the hydraulic cable (or cathead and rope) and pull ring, lower the
sampler until it rests on the botom. First put a keel mark on the drill rods so that the
keel mark will line up with the top of the casing when sampler is supposed to reach
botom. This will warn when the sampler is geting close.
8. Check to make sure that the actuating rods did not slip in the coupling.
7C.3.4 Preparation for Push
Make sure that there is a stable surface for making measurements of the drill rod relative to the
casing. Use the top of the casing itself, or the pull-plate resting on the casing.
1. As soon as the sampler is resting on botom of borehole, measure the actual depth of
the botom of the piston sampler.
2. Put a mark on the drill rod 24 inches above reference. Do this immediately, in case
the sampler creeps down before push.
3. Set up frame to fx actuating rods. The frame shown on the following page is
suggested:
< PREVIOUS VIEW
7C8
Appendix 7C
MAY 2009
The 2-f AW rods driven into the ground should be as close as possible to the mud tub to minimize
the span length of the upper cross bracket.
Fixing the actuating rods to the rig has proven unreliable because the rig can move (lif up) during
sampling. Using cables to hold the actuating rods is very bad because they allow movement due to
slack in cables.
4. Tighten set screw in top cross-bracket (very tight). Place a vise grip very tightly on
actuating rod just above top coupling as a safety measure to keep actuating rods from
moving down.
5. Measure the distance from the botom of the top coupling to the reference plate (a
1
) to
the nearest one-sixteenth inch.
6. Measure distance from reference plate to an arbitrary point on the actuating rod just
above drill rods (b
1
) to nearest one-sixteenth inch. This will give the actual movement of
the actuating rod, including slippage if it occurs.
< PREVIOUS VIEW
7C9
Procedure for Undisturbed, Fixed-Piston Sampling of Cohesionless Soil
MAY 2009
7. Check that mark on drill rods is still 24 inches (c
1
) above reference plate. This will be
used to determine actual push length.
8. Measure distance from a reference plate on the ground to a mark on the AW rods (d
1
).
This will be used to determine whether the support footings move during sampling.
9. Determine the required rate of push and have driller set controls to achieved desired
rate. The driller can experiment with rig controls.
On many drilling rigs, the hydraulic pressure gauge indicating downward pressure on
the drill head will be very useful. If gravel is present in sandy soils, the tube may crum-
ple if the cuting edge hits a large piece of gravel. By experimentation, a limit hydraulic
pressure can be set which, if exceeded, would mean that the tube was hiting gravel,
and, if not exceeded, would mean that the tub penetrated relatively easily for the entire
push length.
On some drilling rigs, the hydraulic pressure gauge will not be informative because it
is not sensitive to resistance against the drill head. Before sampling, determine how the
hydraulic pressure gauge works.
< PREVIOUS VIEW
7C10
Appendix 7C
MAY 2009
a. Determine hydraulic pressure for the drill head moving down with no
resistance (reference pressure).
b. Determine hydraulic pressure change for the case of the rig weight
acting on the drill stem.
c. Decide if monitoring the gauge during sampling will be useful
10. It is important to have a stable connection between the drill head and the drill rods
during the push.
On most drilling rigs, the drill heads are rounded and do not provide a good pushing
surface. A special bracket can be fabricated to allow the head to push down on the drill
rods, as shown in the sketch below:
Alternatively, a right-angle block that allows direct pushing on the drill rods can be
atached to the drill head.
7C.3.5 Advancing Tube
7C.3.5.1 From Weight of Rods
With the actuating rods fxed to top bracket, release the actuating rods from the drill rods. Then re-
measure a, b, c, and d and record on data sheet. Compute the following:
1. Penetration due to weight of rods = c
1
c
2
2. Movement of actuating rods = b
1
b
2
3. Movement of cross bar = a
1
a
2
4. Movement of support rods = d
1
d
2
< PREVIOUS VIEW
7C11
Procedure for Undisturbed, Fixed-Piston Sampling of Cohesionless Soil
MAY 2009
7C.3.5.2 From Drill Head Push
1. The tube should be pushed in a smooth continuous motion, if possible.
2. The driller should be given a mark on the drill stem to watch.
3. It is prudent to push only 22 to 23 inches because there is not much space (less than 1
inch) between the top of the piston and the botom of the plug when a 24-inch push is
completed. If fnes have setled out on top of the piston and the fnes encounter the plug,
the soil in the tube will be compressed and the actuating rods will probably slip in the
cross frame.
4. The inspector or helper should time the push. If necessary, the hydraulic pressure gauge
should be observed by the inspector. The driller should watch the mark on the rods.
5. Afer pushing the tube, back of the drill head of the rig.
7C.3.6 Post-Push Measurements
1. Total Penetration
The total penetration (c
1
c
3
) includes penetration from the
weight of the drill rods.
2. Movement of Actuating Rods Measure and record b
3
. Record the net movement of
the actuating rods
3. Movement of Actuating Rod Support The actuating rods are now fxed to the
bracket and are in tension and may have pulled the support bracket down. Measure
and record a
3
. Record the net movement of actuating rod supports.
4. Movement of Support Footings Measure d
3
and record. Also record the net move-
ment of the support footings.
5. Release Actuating Rods from Bracket Have the driller release the actuating rods
from the top of the bracket. Measure and record a
4
(top bracket may have sprung
upward when it was released from the act. rods). Measure and record d
4
. Record net
movements of a, b, and d.
< PREVIOUS VIEW
7C12
Appendix 7C
MAY 2009
7C.3.7 Lifting Sampler up Borehole
1. Disassemble actuating rod support frame.
2. Put a full head of drill fuid in the borehole.
3. Rotate tube about twice initially, the drill rod joints will come together. The actuating
rods should not be tightened to drill rods. As soon as the actuating rods begin turning
along with the drill rods, stop rotation. Measure b
5
and c
5
(initial distances from a ref-
erence to points on the actuator rods and the drill rods). Rotate the rods two full revo-
lutions (or less, if specifed), then re-measure b
6
and c
6
. Compute c
5
- c
6
(penetration
during shear). Rotation should be performed more than 10 minutes afer the push.
4. Lifing the tube the frst 2 feet is one of the most critical parts of the operation. As the
sampler is lifed, a vacuum will develop at the botom of the tube tending to suck the soil
out of the tube. The tube should be lifed as slowly as possible without jarring or vibrat-
ing the drill stem. At the same time that the tube is being lifed, the drill stem should be
rotated carefully in a manner such that the rods do not whip and the wrench does not
slip. Lifing the frst 2 feet has been accomplished as follows:
a. A swivel can be atached to the drill stem and hydraulic cables used to lif the
tube. The hydraulic cables should be operated slowly enough to lif the tube
in 20 to 30 seconds.
b. On some drilling rigs, the cable hydraulics are not sensitive enough to lif the
tube smoothly. In such cases, a rope can be wrapped around the drill head
and tied around the drill rods. The drill head can then be raised as slowly
as possible, using the rope to facilitate rotation of the drill stem. A full head
of drilling fuid should be main tained in the borehole during this opera-
tion. Note if the drill fuid drops temporarily when the tube has been lifed
approximately 24 inches. If the fuid drops, this may indicate that the soil is
in the tube and the fuid has dropped into the hole formed.
5. If two drill rod stems are being used, slowly lif the drill rods and actuating rods out of
hole at less than 1 foot per second. Maintain head. If the rod joints must be broken, make
sure that rope has been placed in the joints. When breaking actuating rods, use vise grips
to make sure that lowermost actuating rod does not move.
6. If one drill rod stem is being used, before lifing rods, remove as many actuating rods
from hole as possible. If this procedure is used, the actuating rod joint at the top of the
< PREVIOUS VIEW
7C13
Procedure for Undisturbed, Fixed-Piston Sampling of Cohesionless Soil
MAY 2009
piston rod should have been made the loosest and be well greased and not tightened
all the way. Other actuating rod joints should have been tightened very tight with vise
grips. Then, when unscrewing the actuating rods at the surface, all the actuating rods
above the piston rod can be removed. This speeds up the operation considerably.
7C.4 SURFACE HANDLING OF SOIL-FILLED SAMPLER
7C.4.1 Removal of Sampler from Borehole
1. Maintain a head of drilling fuid at all times. When the sampler nears the surface, make
sure that the water fowing into the casing is not too turbulent, which could cause ero-
sion of the soil at the botom of the tube. A stocking over the end of the hose can be used
to reduce turbulence.
2. When the bleeder pipe is observed, reduce the rate of movement to as slow as possible,
but do not decelerate too rapidly.
3. As soon as the botom of the sampler surfaces, slide a plate under the sampler to prevent
any soil from falling out. Do not put fngers under the cuting edge of the tube. Wait
until the drill stem is stable.
4. Start sliding the plate away. Be very careful that fngers are not placed directly under the
cuting edge.
a. If soil starts falling out, put the plate back. Have a piece of foam rubber
approximately 3 inches in diameter and approximately 2 inches thick ready
for placing at the botom of the tube. Make a quick switch of plate and foam
rubber and place a plastic cap at the botom of the tube. Tape the plastic cap
in place. The foam rubber will conform to the soil surface and provide resis-
tance to slippage of the soil.
b. If soil stays in, which it probably will, make a quick observation of the shape
of the botom soil surface and record later. If soil is fush with the botom,
install plastic cap and tape. If soil is missing up to 2 inches from the cuting
edge, put a cylindrical piece of foam in the botom and cap. If soil is far up
into tube, install a vented packer with an extension used for tightening. Filter
paper should be placed on the packer.
5. Disconnecting the Sampler. Make sure that the sampler does not rest on a surface with the
weight of the drill rods on it, because it could bend or kink.
a. If the actuating rods are lef inside the drill rods during withdrawal of the
sampler, the following procedure is suggested: (a) break the joint at the
botom of the 2-foot bleeder pipe, (b) lif the drill stem with rig hydrau-
lics, and (c) unscrew the actuating rods.
Do not let the sampler drop. A piece of foam can be placed under the
sampler so that when it is unscrewed it will drop approximately to
inch onto foam (there may be considerable weight above from the actuat-
ing rods and soil in the tube).
b. If the actuating rods are removed prior to bringing the sampler to ground
surface, unscrew the sampler at a convenient joint.
7C.4.2 Installation of Bottom Packer
The botom packer is installed while the piston is still in place so that a vacuum holds the soil in the
tube. Do not turn the tube upside down.
< PREVIOUS VIEW
7C14
Appendix 7C
MAY 2009
The sampler must be suspended above ground so that the botom can be worked on. This can be done
by mounting a chain clamp to a stationary object free from vibrations (i.e., a work bench or tree, but
not the drill rig or other equipment). The head of the sampler (not the tube should be strapped in,
allowing the botom to be worked on.
Alternatively, the sampler can be suspended by a rope from the derrick of the drill rig. The 2-foot
bleeder pipe should be kept on the sampler so that a swivel can be atached. The rope should be tied
of at the rig. This method is less efcient than the chain clamp method because the drillers cannot
advance the hole while the sample is being worked on by the inspector.
1. Remove temporary botom cap. Note condition of cuting edge and sketch if it is
distorted.
2. Measure distance from cuting edge to soil surface. Sketch profle. Record.
3. Remove soil with puty knife. Save cutings in baggie and jar.
4. Prepare fnal botom surface with trimming tool. Make fat. Final surface should be
1 inch or more from cuting edge for packer to ft.
5. Clean inside wall of tube so that packer will not slip.
6. Measure to soil at three locations.
7. Install packer, vented, with flter paper. Pull down hard on packer to confrm that it
is tight.
8. Measure at three locations to botom of packer.
9. Install botom plastic cap with holes for drainage.
7C.4.3 Removal of Piston
Carefully place the sampler on foam and support it so it is stable and wrapped with foam or
mounted in a chain clamp to minimize vibrations.
Carefully unscrew the joint between the head and the plug. Remove the head.
1. Remove spring.
2. Clean top of plug and around cone. A squeeze botle may be used.
3. Measure from top of plug to top of actuating rod. Record distance from tube botom
to piston. Is this the same as the penetration distance?
4. Remove cone: using vise grips, rotate piston rod clockwise while puting sideways
force on piston rod. The cone will climb the piston rod. Afer removing the cone,
screw the piston rod back down to plug piston.
5. Remove plug.
6. Measure distance from top of tube to top of wash resting on piston. Did the wash
push up against plug?
< PREVIOUS VIEW
7C15
Procedure for Undisturbed, Fixed-Piston Sampling of Cohesionless Soil
MAY 2009
7. Remove fuid and wash resting on top of piston.
a. This can be accomplished by sucking into a long tube and/or using a
squeeze botle.
b. Be sure to clean the piston around the piston rod so that air (not fuid) can
travel around piston rod to release the vacuum.
8. Unscrew piston rod to release vacuum. Screw piston rod back in part way to pull out
piston.
9. Pull out piston slowly. Use piston rod to loosen leathers by swaying piston rod back
and forth. Can use vise grips and wood block to lif if it cannot be done by hand.
If vacuum is not releasing, soil may be stuck to piston. Stick a wire through port to
pierce soil. It is important to prevent soil and fuid above piston from geting below
piston.
10. When the piston is removed, note if soil is stuck to the piston and record thickness.
This is part of gross recovery.
7C.4.4 Cleanout of Top
1. Measure distance from top of tube to fuid above soil and record.
2. Measure distance from top of tube to soil at three locations and record.
3. Record gross recovery and net recovery
4. Was the fuid level at the same locations as the botom of the piston? Remove a
sample of the drill fuid above the soil. Is it water or drilling fuid? Is it from the
pores of the soil?
5. Remove all fuid resting on top of soil.
6. Use trimming tool to remove wash from top.
7. Note: If suction developed below the piston while the piston was being removed,
some soil may have lifed. Slight pressure from a ruler will cause this to drop down.
8. Use foam rubber to remove slop from top of soil. Make sure top of soil is nearly
level.
9. Measure to top of soil in three places. Record.
7C.4.5 Installation of Top Packer
1. Prepare the botom half of a vented packer with a flter paper and a wire for future
retrieval.
2. Place packer on top of soil. This provides a good level surface for measuring.
< PREVIOUS VIEW
7C16
Appendix 7C
MAY 2009
3. Measure accurately at three locations. These data will later be used for monitoring
height changes during handling and shipment. Record.
4. Put a moistened paper towel in top.
5. Put a vented plastic cap on top and tape in place.
< PREVIOUS VIEW
7C17
Procedure for Undisturbed, Fixed-Piston Sampling of Cohesionless Soil
MAY 2009
7C.4.6 Preparation for Shipment
1. Label with the following:
Top
Project Number
Project Name
Boring Number, Sample Number
Depths
Penetration, Gross Recovery, Net Recovery
2. If needed, apply vacuum below botom packer to draw water out of sample. This
will increase the stifness of sand without signifcantly changing the void ratio.
< PREVIOUS VIEW
Appendix 7D
COMBINING NEWMARK-TYPE DISPLACEMENTS
COMPUTED USING 1D SITE-RESPONSE ANALYSES
7D1 MAY 2009
The computer program SHAKE calculates time histories of horizontal shear stress acting at the
boundaries of soil layers. To calculate a time history of acceleration for a wedge, and then a displace-
ment, the following procedure can be used (Figure 7D.1):
1. Perform SHAKE analyses for two representative profles through the potential
wedge to obtain horizontal shear stress time histories
h
(t) at two points A and B on
the sliding surface.
2. Calculate the time history of the average acceleration k(t) of the soil above the two
points:
k
A
(t) =
h
A
(t) /
vA
k
B
(t) =
h
B
(t) /
v
B
where:
h
A
(t) = earthquake shear stress at point A
h
B
(t) = earthquake shear stress at point B
v
A
= total vertical stress at point A
v
B
= total vertical stress at point B
3. Integrate the k
A
(t) and k
B
(t) time histories for various values of k
y
to develop curves of
k
y
/ k
max
versus displacement (Figure 7D.1).
4. Compute k
OVERALLmax
for the entire wedge:
k
OVERALLmax
=
k
Amax
W
A
+ k
Bmax
W
B
W
A
+ W
B
where:
k
Amax
= maximum value of k
A
(t)
k
Bmax
= maximum value of k
B
(t)
W
A
, W
B
= weights of portions A and B of failure wedge
5. Using the value of k
y
computed from the pseudostatic stability analyses, calculate
k
y
/ k
max
and enter the curves developed in Step 3 to obtain a displacement.
APPENDIX 7D
< PREVIOUS VIEW
7D2
Appendix 7D
MAY 2009
FIGURE 7D.1 DEFORMATION ANALYSIS USING SHAKE
D
I
S
P
L
A
C
E
M
E
N
T
A
B
LOG ky / kmax
kOVERALLmax =
kAmax
W
A
+ kBmax
W
B
W
A
+
W
B
FIGURE 7.D.1 DEFORMATION ANALYSIS USING SHAKE
SHAKE
PROFILE A
SHAKE
PROFILE B
V
A
V
B
W
A
W
B
(t) = k
A
kAmax = maximum of k
A
(t)
V
A
h
(t)
A
= k
B
(t) kBmax = maximum of k
B
(t)
V
B
h
(t)
B
h
(t)
B
h
(t)
A
k
y
/ kOVERALLmax
where:
h
(t) = time history of horizontal shear stress
t
= surface tensile strain (dim)
width shown. If the mined width is less than critical, it is termed subcritical, and the amount of sub-
sidence is less than the maximum subsidence experienced with wider extraction areas. If the mined
width is greater than the critical width, supercritical conditions prevail, and the subsidence trough
will exhibit a fat botom depression, as shown in Figure 8.2. The fgure also shows the angle of incli-
nation between the vertical at the edge of the workings and the point of zero vertical displacement at
the edge of the trough, which is termed the limit angle or angle of draw. Depending on the predic-
tion method used, survey standards employed, or criticality of a surface structure, determination
of the angle of draw is typically based on a minimal vertical movement of between 0.01 and 0.1 feet
(for predictive measures that assume an asymptotic approach to zero such as the tangent-hyperbolic
function). The coal extraction width can be a result of longwall mining, second mining of pillars, or
pillars crushing or punching into the foor.
< PREVIOUS VIEW
Chapter 8
8-12 MAY 2009
Subsidence lowers surface grades and elevations and can afect surface drainage and alter impound-
ment freeboard. Strains and horizontal displacements associated with subsidence can impact the
structural integrity of surface structures such as dams (including internal drains and decants),
embankments, and bridges and buried structures such as pipelines and large culverts. Zones of sur-
face tension and compression develop during mining, resulting in horizontal movement profles that
are primarily a function of the extraction width, as depicted in Figure 8.3. The most severe subsidence
impacts on many impoundment structures occur where the tensile stain is highest, while for other
structures it is the area of greatest tilt or subsidence-induced slope. The existing condition of a struc-
ture and slope of the terrain on which the structure is situated also afect subsidence impacts.
FIGURE 8.2 INFLUENCE OF EXTRACTION WIDTH ON SUBSIDENCE
SUPERCRITICAL WIDTH
CRITICAL WIDTH
SUBCRITICAL WIDTH
DEPTH OF
SEAM (D)
COAL SEAM
t
ANGLE OF
DRAW
SMAX
SUBSIDENCE FOR
SUBCRITICAL WIDTH
SUBSIDENCE FOR
CRITICAL WIDTH
SUBSIDENCE FOR
SUPERCRITICAL WIDTH
COAL SEAM
(SINGH, 1992)
FIGURE 8.2 INFLUENCE OF EXTRACTION WIDTH ON SUBSIDENCE
FIGURE 8.1 STRATA DISTURBANCE AND SUBSIDENCE CAUSED BY MINING
FIGURE 8.1 STRATA DISTURBANCE AND SUBSIDENCE CAUSED BY MINING
(ADAPTED FROM SINGH AND KENDORSKI, 1981;
PENG AND CHIANG, 1984)
ORIGINAL GROUND SURFACE
SURFACE
CRACKING
CAVING
t
3 to 10t
COAL
FLOOR
HEAVE
BED SEPARATION
SURFACE ZONE
MAIN ROOF
(CONTINUOUS
DEFORMATION
ZONE)
FRACTURED
ZONE
CAVED
ZONE
FLOOR
HEAVE
FIGURE 8.2 INFLUENCE OF EXTRACTION WIDTH ON SUBSIDENCE
FIGURE 8.1 STRATA DISTURBANCE AND SUBSIDENCE CAUSED BY MINING
< PREVIOUS VIEW
8-13
Site Mining and Foundation Issues
MAY 2009
FIGURE 8.3 DISPLACEMENT AND STRAIN FOR VARIOUS EXTRACTION WIDTHS
8.3a SUBCRITICAL WIDTH
S < SMAX
COAL SEAM
SUBCRITICAL WIDTH
EXTRACTION ZONE
HORIZONTAL
DISPLACEMENT
ANGLE OF
DRAW
HORIZONTAL
STRAIN
ANGLE OF
BREAK
SUBSIDENCE PROFILE
GROUND
SURFACE
TENSION
COMPRESSION
- -
+ +
8.3b CRITICAL WIDTH
CRITICAL WIDTH
COAL SEAM
SUBSIDENCE PROFILE
SMAX
EXTRACTION ZONE
ANGLE OF
DRAW
ANGLE OF
BREAK
TENSION
COMPRESSION
GROUND
SURFACE
HORIZONTAL
DISPLACEMENT
HORIZONTAL
STRAIN
+
- -
+
8.3c SUPERCRITICAL WIDTH
HORIZONTAL
STRAIN
HORIZONTAL
DISPLACEMENT
SMAX SMAX
SUBSIDENCE PROFILE
EXTRACTION ZONE
COAL SEAM
SUPERCRITICAL WIDTH
ANGLE OF
DRAW
ANGLE OF
BREAK
GROUND
SURFACE
TENSION
COMPRESSION
+
-
+
-
(SINGH, 1992)
FIGURE 8.3 DISPLACEMENT AND STRAIN FOR VARIOUS EXTRACTION WIDTHS
FIGURE 8.3 DISPLACEMENT AND STRAIN FOR VARIOUS EXTRACTION WIDTHS
< PREVIOUS VIEW
Chapter 8
8-14 MAY 2009
For total extraction mining, subsidence follows the advance of the mine workings, and horizontal ten-
sile and compressive strain regions move laterally with the mining, as shown in Figure 8.4. Therefore,
determination of the impacts of full extraction mining beneath embankments and impoundments
must be based upon subsidence development (i.e., dynamic subsidence) and the associated traveling
strain curve and not just the fnal subsidence profle. Areas of greatest impact generally occur near
the boundaries of the extraction zone. Figure 8.5 depicts the ground movements caused by subsid-
ence. Singh (1992) summarizes factors that afect mine subsidence, including: (1) seam thickness,
depth, and dip; (2) mine foor, roof and overburden; (3) geologic discontinuities and in-situ stresses;
(4) surface topography; (5) groundwater; and (6) percent of extraction, advance rate, backflling, and
elapsed time.
8.4.2 Potential Subsidence and Failure Mechanisms
There are several potential mechanisms that are associated with mine subsidence and mine break-
throughs. Subsidence can occur as a result of caving and fracturing of the mine overburden and
migration of the disturbance to the surface (e.g., sinkhole development), pillar crushing, or pillar
punching. Also, planned subsidence can occur as a result of: (1) total extraction mining using long-
wall methods or secondary mining of pillars or (2) room and pillar mining with partial extraction
leaving pillar remnants designed to yield. The condition of the coal and overburden and the mine
extraction method will determine the extent to which subsidence will propagate upward toward the
surface and the degree of fracturing and deformation that can occur at the surface. Longwall subsid-
A
B
C
D
E
F G H I J K
TENSION
COMPRESSION
8.4a SUBSIDENCE DEVELOPMENT
8.4b TRAVELING STRAIN CURVE
FIGURE 8.4 DEVELOPMENT OF SUBSIDENCE TROUGH
AND STRAINS WITH FACE ADVANCE
(RELLENSMANN AND WAGNER, 1957)
FACE
ADVANCE
A B C D E F G H I J K
FINAL PROFILE AFTER TIME-
DEPENDENT SUBSIDENCE
A
C
E
G
I
K
FIGURE 8.4 DEVELOPMENT OF SUBSIDENCE TROUGH AND STRAINS WITH FACE ADVANCE
FIGURE 8.4 DEVELOPMENT OF SUBSIDENCE TROUGH AND STRAINS WITH FACE ADVANCE
< PREVIOUS VIEW
8-15
Site Mining and Foundation Issues
MAY 2009
ence occurs shortly afer mining, but room and pillar subsidence may occur years or decades afer
mining, as conditions or in-mine support deteriorate.
Subsidence of underground mine workings can lead to breakthrough if an impoundment is located
directly over or near the mine. If the caved or fracture zone intersects with the surface or surface soil
layer, a breakthrough can occur. The driving force for a breakthrough is the pressure gradient (earth
and water pressure) between the impoundment and the underground mine workings. The added
weight from the embankment and reservoir, combined with increased seepage heads, increases the
stress on the underlying strata. Mechanisms that can cause a breakthrough include internal erosion
or piping, outcrop barrier failure, hillside movement and disturbance, mine seal failure, or barrier
(coal, soil and/or rock) decomposition. These mechanisms are frequently interrelated and should be
carefully evaluated as part of the analysis of the potential for breakthrough.
8.4.2.1 Sinkholes
A sinkhole is a depression or opening in the ground surface above an underground mine void where
the mine roof has fractured and fallen/caved and the disturbance has intersected the surface or soil
mantel. The fracturing of the mine roof material can eventually extend high enough that an opening,
or at least a weakened area, is created at the ground surface, particularly where the overburden is
thin. A sinkhole can serve as a direct conduit from an impoundment to underground mine work-
ings. Factors contributing to sinkhole development include: (1) the presence of a mine void, (2) low
overburden thickness, (3) mine roof material that is not sufciently strong or durable to span the
mine opening, (4) fractures in the mine roof, (5) unconsolidated soil and weathered rock above the
mine roof, and (6) pressure exerted by refuse and/or water at the surface. The action of water fowing
through fractured strata can cause deterioration and can enlarge a sinkhole beyond the depth and
extent that would occur in the absence of water.
FIGURE 8.5 GROUND MOVEMENTS CAUSED BY SUBSIDENCE
FIGURE 8.5 GROUND MOVEMENTS CAUSED BY SUBSIDENCE
CURVATURE (S")
INFLECTION POINTS
HORIZONTAL
STRAIN (V')
DEPTH OF
OVERBURDEN
COAL SEAM
ANGLE OF DIP
BREAK
ANGLE
DRAW
ANGLE
VERTICAL
DISPLACEMENT (S)
S
max
TENSION
COMPRESSION
INFLECTION
POINT
HORIZONTAL
DISPLACEMENT (V)
SLOPE (S')
(SINGH, 1978)
INFLECTION
POINT
EXTRACTION ZONE
FIGURE 8.5 GROUND MOVEMENTS CAUSED BY SUBSIDENCE
< PREVIOUS VIEW
Chapter 8
8-16 MAY 2009
A sinkhole will not develop if the rock strata above the mine are either strong enough to span open-
ings without collapsing, the strata above the mine are thick enough that an arch forms over open-
ings and prevents the collapse feature from propagating to the surface, or the swell of falling rock
checks the propagation of the collapse before it afects the mantle of unconsolidated material. One
study in Pennsylvania (Gray et al., 1977) found that 81 percent of sinkholes occurred where there
was less than 100 feet of cover, with most occurring at 50 to 60 feet of cover. However, there have
been unusual cases, such as a sinkhole in Illinois (DuMontelle et al., 1981) where the cover was
165 feet, although the overburden thickness in this case included a substantial amount of glacial
till. The mine depth associated with sinkholes reported from mines located primarily in Colorado,
Utah and Wyoming (Dunrud and Osterwald, 1980) was generally 10 to 15 times the seam thickness.
Analysis of roof strata to determine if they will indefnitely span mine entries is difcult because:
(1) the strength of the rock is not easily defned, (2) the efect of joints on the integrity of the roof
(especially at shallow depths) adds considerable uncertainty, and (3) long-term integrity of coal pil-
lars and roof support system can be an issue. For these reasons, the adequacy of the mine overbur-
den to prevent sinkhole development is commonly assessed by applying certain rule-of-thumb
type guidelines that have been developed based on experience. Guidelines developed by Babcock
and Hooker (1977) are illustrated in Figure 8.6. These guidelines are generally considered to be
conservative, and they should be carefully reviewed before detailed analyses to assess the potential
for subsidence and related diferential movement and strain are performed. Mine development
work has been performed and production areas have been successfully implemented following
the guidelines shown in Figure 8.6. However, where subsidence has the potential to afect a high-
hazard-potential impounding embankment or to allow a breakthrough that could afect the safety
of miners or the public, these guidelines should be used only in conjunction with more rigorous
site-specifc analysis.
With respect to sinkholes, the guidelines recommend for frst mining only that the thickness of solid
strata should be equal to at least 5 times the entry width or 10 times the extraction height, which-
ever is greater. These guidelines are consistent with fndings that compression arches in overbur-
den are normally stable if the mining width is limited to one-fourth to one-half of the overburden
height. In other words, the compression arch is typically stable if the thickness of the rock strata
above the mine is from 2 times (strong strata) to 4 times (weaker strata) the entry width. Adding a
margin of safety, the rule-of-thumb is that the overburden thickness should be 5 times the entry
width. Similarly, the criterion related to the extraction height is based on adding a margin of safety
to the observation that the height of collapse above a mine entry generally does not exceed 3 to 5
times the extraction height, likely because of the swell of the collapsed material.
A key point in the use of these guidelines is that the strata thickness refers to solid overburden
strata. Soil, weathered rock, or weak rock should not be included in the solid strata thickness since
they may not provide the strength needed to resist sinkhole development (MSHA, 2003).
The guidelines in Babcock and Hooker (1977) suggest that a lesser strata thickness may be accept-
able if the overlying strata consist of competent rock (e.g., competent bed of sandstone or similar
material) with a thickness of at least 1.75 times the entry width. Competent rock is normally taken
to mean a homogeneous, massive layer of sandstone or limestone (MSHA, 2003). This guideline is
based on analyses of the overlying strata as a beam with minimum tensile strength (20 psi). While
any intact piece of rock will likely exceed this tensile strength, joints or discontinuities will reduce
the efective strength of the beam. Because the potential impact of joints and weathering, especially
near an outcrop, cannot be modeled with confdence, this approach should normally not be used for
a critical subsidence or breakthrough situation. Furthermore, there is a possibility that the strength of
the overlying strata will deteriorate over time due to seepage and weathering, especially as the mine
pool level rises.
< PREVIOUS VIEW
8-17
Site Mining and Foundation Issues
MAY 2009
8.6a SAFETY ZONE BENEATH BODY OF SURFACE WATER
FIGURE 8.6 SAFETY ZONE GUIDELINES FOR MINING IN THE VICINTY
OF DAMS AND IMPOUNDMENTS
SURFACE WATER
(BABCOCK AND HOOKER, 1977)
ROOM AND PILLAR
PANEL
TOTAL EXTRACTION
D = 5s or 10t, whichever is larger
D = 3p or 270 feet, whichever is larger
D = 60t
s = maximum entry width (ft)
p = panel width (ft)
t = extraction height (ft)
8.6b SAFETY ZONE BENEATH DAM AND IMPOUNDED BODY OF SURFACE WATER
NOTE: THE ZONE OF NO EXTRACTION TO DEPTH "D" IS PRESUMED TO COMPRISE SOLID ROCK STRATA (IF MATERIAL
OTHER THAN SOLID ROCK COVER IS INCLUDED, IT IS NECESSARY TO DEMONSTRATE THE NATURE AND
PERMEABILITY OF SUCH MATERIAL). THE ZONE OF NO EXTRACTION MAY BE INCREASED OR DECREASED,
IF JUSTIFIED BY LOCAL OBSERVATION AND/OR EXPERIENCE.
ROOM AND PILLAR
PANEL
TOTAL EXTRACTION
D = 5s or 10t, whichever is larger
D = 3p or 270 feet, whichever is larger
D = 60t
s = maximum entry width (ft)
p = panel width (ft)
t = extraction height (ft)
D
NORMAL
EXTRACTION
PERMITTED ZONE OF NO EXTRACTION
3
5
0
'
3
5
0
'
65
ZONE OF EXTRACTION
USING GUIDELINES
65
NORMAL
EXTRACTION
PERMITTED
ZONE OF NO EXTRACTION
D
200'
200'
200'
IMPOUNDMENT
200'
NORMAL
EXTRACTION
PERMITTED
NORMAL
EXTRACTION
PERMITTED
200'
3
5
0
'
D
3
5
0
'ZONE OF NO
EXTRACTION
ZONE OF NO
EXTRACTION TO
PROTECT DAM
3
5
0
'
ZONE OF EXTRACTION
USING GUIDELINES
65
DAM
65 65
FIGURE 8.6 SAFETY ZONE GUIDELINES FOR MINING
IN THE VICINITY
OF DAMS AND IMPOUNDMENTS
FIGURE 8.6 SAFETY ZONE GUIDELINES FOR MINING IN THE VICINITY
OF DAMS AND IMPOUNDMENTS
< PREVIOUS VIEW
Chapter 8
8-18 MAY 2009
In guidance for evaluating breakthrough potential, MSHA (2003) indicates that any location where
the cover over a mine entry is less than 100 feet of solid strata (especially at locations near the outcrop
where additional weathering and stress relief has occurred) is a concern for sinkhole development.
Accordingly, if sinkhole development cannot be reliably ruled out, preventive measures should be
considered.
8.4.2.2 Pillar Failure
Loss of support due to coal pillar failure causes a mine roof to sag or collapse. This can create or open
fractures in the overburden. These fractures may cause a roof fall and consequent sinkholes, or the
fractures may create zones where internal erosion can occur. Furthermore, failure of one pillar trans-
fers the load to surrounding pillars and may lead to progressive pillar failure (sudden or gradual) or
excessive displacements over a relatively large area.
Pillar crushing occurs when the load on a pillar exceeds its strength. This can be caused by exist-
ing loads, additional loading from impounded slurry and/or water, loss of strength in the coal from
chemical decomposition or from slow oxidation, mine fre, or loss of buoyant pressure resulting
from a lowered mine pool. In addition to pillar strength, the pillar width to height ratio (w/ h) is also
important (Mark, 2006). For slender pillars (w/ h < 4), failure ofen results in nearly complete loss
of load-bearing capacity, sometimes with sudden and total collapse. Pillars with w/h between about 4
and 10 are largely elastic with a possible plastic core, and failures tend to occur gradually with post-
failure residual strength essentially constant. The pillars deform until they have shed enough load to
stop the process. Pillars with w/h greater than 10 (referred to as squat) have a plastic core and may
strain harden once the loss of initial strength due to crushing or yielding of the outer elastic portion of
the pillar occurs. Afer this initial crushing, the pillars gain strength as they deform. The implications
for surface structures of the failure of slender pillars with shallow cover are much more signifcant
than those associated with yielding of squat pillars at great depth.
A number of formulas for analyzing the strength of a pillar have been developed, and computer pro-
grams for performing pillar analyses are available. One example is the program ARMPS (Analysis of
Retreat Mining Pillar Stability) developed by National Institute for Occupational Safety and Health
(NIOSH). This program uses the Mark-Bieniawski formula to determine pillar strength, and it has the
capability to account for loadings on barriers and abutment pressures.
Pillar stability formulas can be divided into two categories analytical and empirical. Analytical
formulas require extensive material testing, an understanding of loading under varying conditions,
and a safety factor (typically about 2) based on knowledge and understanding of all variables. These
relationships are best applied by engineers who are experienced in mining rock mechanics. One such
relationship, Wilsons equation, which is one of the frst analytical models developed for estimating
pillar strength, is directly calculated, thus making it more fexible and adaptable to actual conditions
than any empirical equation. It can be used to estimate the stress distribution from the edge of a pillar
to the center based on the confned core theory (Wilson and Ashwin, 1972). Wilsons equation uses
the Mohr-Coulomb failure criterion for modeling the coal and surrounding rock; however, at high
confnement (high w/h ratio) coal strength is not linear with the result that it overestimates pillar
strengths. Scovazzo (1995) modifed Wilsons equation to incorporate more appropriate coal and rock
failure criteria, specifcally those presented by Kalamaras and Bieniawski (1993).
The most commonly used pillar stability formulas are empirical equations where few parameters
need to be defned. Since empirical equations are based on statistical analysis of failures and success-
ful designs, a stability factor (not to be confused with safety factor) is employed. One commonly used
empirical equation is the Mark-Bieniawski formula. For the Mark-Bieniawski formula, the recom-
mended stability factor is 1.5 for mines less than 750 feet deep (Mark, 2006). A smaller stability factor
< PREVIOUS VIEW
8-19
Site Mining and Foundation Issues
MAY 2009
is generally used for mines greater than 750 feet deep; for example, a stability factor of 0.9 is used for
pillars with over 1,250 feet of cover. These recommended stability factors are based on the assump-
tion of equal area loading. Empirical equations are applicable to specifc mining regions and coal
seams and should not be applied outside the region for which they were developed unless statistical
analyses are performed for the region in which the equation is intended to be used. While some long-
term pillar instability can be tolerated in certain mining situations, impoundment designers should
consider a higher margin of safety for overburden support and control of detrimental diferential
movements within the dam and foundation.
Conservative coal strengths based on statistical analysis of failure should be used for pillar failure
analysis. Laboratory tests for verifying the strength of the coal can be conducted, although this rarely
occurs and is not encouraged. The limited use of testing is in part due to the scater typically encoun-
tered in uniaxial rock tests, a problem that is exacerbated by coal cleating and sofness, which make
test samples difcult to prepare. A laboratory testing program should include sufcient samples and
statistical analysis to verify that data are reliable. As a result, the mass uniaxial compressive strength
of coal is ofen assumed to be 900 psi.
The uniaxial compressive strength of coal determined in the laboratory is many times the in-situ
strength of coal in a pillar. This size efect is caused by faws in the coal that are present in the larger
mass of the pillar but not in sample sizes tested in the laboratory. It would take test samples with the
impractical minimum dimension in the range of 3 feet (Pariseau, 1975) to 5 feet (Bieniawski, 1968) for
accurate determination of the mass uniaxial compressive strength of coal. The most accepted method
for reducing laboratory compressive strengths for coal to refect in-situ strength of the coal was devel-
oped by Hustrulid (1976) and is recommended by Bieniawski (1992) for use with his pillar formula.
In any analysis of pillar stability, it is best if the method being used is calibrated to conditions at the
subject mine. Thus, the model should be applied to a number of locations in the mine to determine
how well it refects actual conditions or the actual performance of mine pillars. Pillar analysis is more
complicated when multiple seams are mined. In such cases, the loading conditions are much more
complex due to load transfer and the potential for stress concentrations. Simplifed sofware models,
such as LAMODEL, are available for performing this type of analysis. Higher pillar stability factors
should be used for multiple seam analyses to account for the additional uncertainty. Both LAMODEL
and ARMPS are available from the NIOSH web site.
8.4.2.3 Pillar Punching (Floor Failure)
Pillar punching (pillar foundation failure) occurs when a pillar pushes into the mine foor allowing
the roof to sag. This sagging can create fractures and/or open existing joints in the overburden. If the
punching occurs over a large area, the surface will be afected in a similar manner (normally without
bed separation) to the situation where total extraction of a thinner seam has occurred.
Pillar punching occurs when the load on a pillar exceeds the bearing capacity of the mine foor beneath
it. Pillar punching can be caused by existing loads, saturation of the foor material causing sofening
and loss of strength, additional loading from impounded slurry and/or water, and loss of buoyant
pressure from a lowered mine pool.
A key factor to consider in evaluating the potential for pillar punching is experience elsewhere in the
mine. A review of foor and pillar performance data, particularly for wet conditions, should be made
as part of foundation bearing capacity analyses for evaluation of pillar punching. It may be appropri-
ate to reduce the foor strength when both slender pillars and wet conditions are present or are antici-
pated as the result of construction of an impoundment. Additional information on pillar punching
and associated foor heave can be found in Ganow (1975) and in Adler and Sun (1968).
< PREVIOUS VIEW
Chapter 8
8-20 MAY 2009
8.4.2.4 Outcrop Barrier Failure by Shear or Punching
In an outcrop barrier situation, a potential failure mode that must be considered is whether the pres-
sure from water and/or slurry in the impoundment may become high enough to overcome the shear
strength that is holding in place the plug of material separating the impoundment from the mine
works. Failure can occur through the coal seam itself, through the strata above the coal seam, through
a weakened foor layer, or along an interface between strata. Figure 8.7 illustrates a typical outcrop
barrier cross section, including the coal barrier, jointing and stress relief fractures. Surface features
such as slope instability and regrading for access roads are also shown on the fgure.
Analysis of this failure mode for a postulated plug separating the impoundment from the mine
involves comparison of the cumulative force tending to push the plug into the mine to the avail-
able resistance from shear forces around the perimeter of the plug. The pressure driving the plug is
normally taken as the hydrostatic head from the impoundment plus applicable lateral earth pres-
sure from setled fnes. A major issue with this type of analysis is judging what to use for the shear
strength along the top, botom and sides of the plug. Factors such as the presence of weathered
joints, the presence of cleats in the coal, the sofening of the foor from saturation, and the difculty
of obtaining test data can result in uncertainties with respect to shear strength. Furthermore, as
seepage into the mine occurs, the strength along potential shearing surfaces may degrade over
time, and, if sloughing or a roof fall occurs, the thickness of the plug may be reduced. Additionally,
the potential hydraulic gradient across the plug must be carefully considered, as water pressures
may be elevated on planes of discontinuity within the plug and present a more severe loading con-
dition than at the plug boundary. Given these uncertainties, any shear-type failure analysis should
be based upon conservative values for shear strength and a conservative factor of safety of at least
2.0 (MSHA, 2003).
8.4.2.5 Internal Erosion
Internal erosion, or piping, is the movement of material (typically soil particles) under seepage forces,
and it can occur where the gradation is such that particles are dislodged and carried along by seep-
age and drag forces. When soil overlies rock, movement of material can occur along or into joints,
stress relief fractures, or subsidence fractures. Figure 8.7 illustrates typical conditions encountered
near a coal seam outcrop where weathered joints may represent seepage pathways from a future
impoundment. The presence of stress relief fractures or subsidence fractures would exacerbate the
concern for seepage and internal erosion. Seeping water creates a drag force on the material that it is
seeping through or around. If this drag force is greater than the frictional or cohesive forces holding
material particles in place, they will be transported by the seeping water. As smaller particles are car-
ried away, additional fow occurs, increasing the drag force and dislodging even larger particles. The
process can continue until a channel (also referred to as a pipe) is formed. Piping into foundation
discontinuities can lead to failure of an embankment. In a breakthrough situation, where the source
of the seepage is an impoundment and internal erosion occurs through the material between the
pool and the mine workings, the pipe can gradually enlarge and lengthen to the point where there
is a signifcant discharge into the mine. Piping is dependent on the type, gradation, consistency, and
cementation of the intervening material; relative hydraulic conductivities of the materials along the
critical seepage front; and the hydraulic gradient.
One technique for preventing internal erosion is to incorporate flter layers (material with grain-size
distributions that prevent particles from moving along a seepage path) at critical locations. Granu-
lar flter material or various geosynthetic materials can be used for this purpose. Filter criteria are
discussed in Section 6.6.2.3. Also, the gradient can be reduced by using lower-hydraulic-conductiv-
ity materials along the upstream portion of the seepage path to increase head loss before the water
reaches a critical zone. Excess water pressure can also be controlled by providing drainage to reduce
pressures upstream of more erodible zones, such as by installing an internal drain protected by flters.
< PREVIOUS VIEW
8-21
Site Mining and Foundation Issues
MAY 2009
In breakthrough situations, seepage into an underground mine can occur through the roof, foor,
or coal seam. If the material at the location(s) where the seepage enters the mine is weathered, fne-
grained, or loosened as coal may be along a face or rib, then the force of the seeping water may carry
particles away and/or cause the in-situ material to slake or unravel. Over time, this can cause the area
to weaken, slough, or progressively deteriorate. This is especially a concern near an outcrop barrier,
where the internal erosion could cause the ground to give way or a pipe to form through the afected
zone. Either situation could lead to an uncontrolled hydraulic connection between an impoundment
and the mine workings.
Whenever seepage is fowing outward and exiting at an unconfned surface, as may be the case
near the toe of a dam or impounding embankment, the value of the critical gradient with respect to
piping is approximately one. However, when seepage is generally horizontal, such as through a rib,
or vertically downward, such as through the mine roof, internal erosion can develop under smaller
gradients.
Determining the hydraulic gradient between an impoundment and a mine requires accurate char-
acterization of the permeabilities of the intervening materials. Near an outcrop, the seepage may be
governed by fow along weathered joints and thus may be difcult to characterize and model. Gra-
dients can be estimated by drawing fow nets or by using a fnite element seepage program. Once
the gradient has been estimated, the efect of the seepage force on the seepage medium can be deter-
mined. The seepage force, per unit volume of seepage medium, is equal to the product of the gradient
and the unit weight of water. Whether the medium material will be dislodged by a combination of
seepage and gravity forces depends on the frictional and cohesive strength of the material.
Cohesionless soils, particularly silts and fne sands, are most susceptible to piping. Clays are more
resistant to piping because the cohesive strength of these soils helps to prevent particles from being
carried away. However, clayey soils are not immune to internal erosion, especially if they abut open-
FIGURE 8.7 HILLSIDE DISTURBANCE AT COAL OUTCROP
ACCESS
ROAD
STEEP SLOPE AT JOINT
IN RESISTANT ROCK
SHALE
SANDSTONE WITH
INTERBEDDED SHALE
COAL SEAM
SHALE
REMAINING COAL BARRIER MINE WORKINGS
SHALE
SANDSTONE
SANDSTONE
SANDSTONE WITH
INTERBEDDED SHALE
SANDSTONE
TALUS
UNDERCLAY
HIGHLY WEATHERED
MATERIAL
JOINTS
FIGURE 8.7 HILLSIDE DISTURBANCE AT COAL OUTCROP
(ADAPTED FROM SAMES AND MOEBS, 1992)
FIGURE 8.7 HILLSIDE DISTURBANCE AT COAL OUTCROP
< PREVIOUS VIEW
Chapter 8
8-22 MAY 2009
graded materials, open joints or fractures. Also, sof rocks, such as weakly cemented sandstones, have
been associated with piping failures. Even shales, which are usually considered resistant to piping,
have developed piping voids under conditions of very high gradients (Sowers and Sowers, 1970). In
these unusual cases involving rock strata, weaker characteristics of the rock mass presumably led to
pathways of increased seepage that intersected and then led to piping of more erodible materials.
If the materials between the mine works and an impoundment are prone to internal erosion or are
sensitive to progressive deterioration, then the impoundment should be designed: (1) with suitable
flters and drains so that seepage can be collected and released in a controlled fashion, (2) with seepage
barriers so that pressures are minimized, or (3) with some combination of these two approaches.
8.4.2.6 Bulkhead Failure
If there is a mine opening in a potential breakthrough area, then the potential for failure of the bulk-
head used to block the mine opening must be evaluated. Depending on their thickness and shape,
bulkheads can fail when the pressure acting on them causes the bending or shear strength of the
bulkhead material to be exceeded. Bulkheads can also fail if the material that they are anchored or
keyed into is not strong enough to resist the applied pressures and the pressures from water seeping
around the bulkhead, or if the bulkhead is not adequately anchored to the surrounding rock strata.
Hydraulic fracturing within foor strata can aggravate seepage and transmission of pressures result-
ing in the typical recommendation to remove underclay/claystone at bulkhead locations. Analysis of
bulkheads is addressed in Section 8.5.2.
8.4.2.7 Trough Subsidence and Subsidence Cracks
If room and pillar workings are located near the footprint of an impoundment (with inadequate sta-
bility factors on the remaining pillars), or if either longwall mining or secondary mining of pillars has
occurred, an analysis of the potential impact of trough subsidence should be performed. In this situ-
ation, zones of tension and compression stress or deformation extend from the mine workings to the
ground surface. The area at the ground surface afected by total extraction mining is typically larger
than the mined area and is related to the draw angle. The assumed draw angle should be consistent
with local experience, with past subsidence associated with mining in the coal seam, site topography,
and the method of mining.
The efect that subsidence has on steep natural slopes is a mater of debate (Luo et al., 1996). It
appears that the efect is greater than the generally accepted regional angle of draw and is depen-
dent upon local geologic conditions and the direction of mining. The approach of mining from the
downslope and towards the upslope appears to result in the greatest horizontal displacement and
accompanying stress.
If mining is approaching an impoundment, operations should be terminated at a point where ground
deformations will not adversely afect the impoundment. It is rare for the angle of draw to exceed
25 in the U.S., and typical values range from 10 to 25. In the Northern Appalachian Coalfeld, the
maximum angle of draw is normally 25, while in the Black Warrior Basin and Southern Appalachian
Coalfeld, it is typically around 12. Work by Scovazzo (2008) at three mines in the Arkoma Basin
places this value at 19. For the Illinois Basin, the angle of draw typically is in the range of 20 to 25 in
the southern portion, increasing to 30 in the northern portion as glacial till thickness increases. The
angle of draw varies with the dip of the coal seam and the slope of the ground surface and decreases
as the percentage of hard rock (sandstone and limestone) in the overburden increases. No patern for
the western U.S. has been identifed.
For total extraction mining (longwall mining or secondary mining of pillars) directly below an
impoundment, the U.S. Bureau of Mines (USBM) guidelines provided in Babcock and Hooker (1977)
< PREVIOUS VIEW
8-23
Site Mining and Foundation Issues
MAY 2009
indicate that the amount of cover should be at least 60 times the mining height, as illustrated in Figure
8.6. This guideline was based on studies that looked at disturbance of the rock strata, and the change in
the hydraulic conductivity of these strata, above mined areas. Data indicate that it is unlikely that the
rock strata above an area of total extraction mining will be disturbed for more than 25 to 35 times the
mining height. This thickness guideline (Babcock and Hooker, 1977) is based on the concept of a con-
strained zone. When total extraction occurs, cracks and joints open at the surface and in the mine roof,
but if the overburden is thick enough, the induced stress is absorbed or resisted without fracturing. It
should be noted that the Babcock and Hooker (1977) strata-thickness criterion is for evaluation of the
potential for developing a hydraulic connection between the surface and the mine; it does not address
the potential for other adverse efects such as tensile strains at the surface and loss of freeboard.
Another USBM research report titled, Criteria for Determining When a Body of Surface Water Constitutes
a Hazard to Mining (Kendorski et al., 1979), recommends cover thicknesses greater than 60 times the
mining height when the mining height is less than 7.5 feet. For example, this report recommends that
the overburden thickness should be 71, 80, 95, and 117 times the mining height for mining heights of
6, 5, 4, and 3 feet, respectively. The report also indicates that, where infow can be tolerated, the thick-
ness of cover can be reduced if certain types of strata (e.g., claystone or shales that are less prone to
cracking and have low hydraulic conductivity) are present.
The preceding guidelines refer to bodies of surface water and do not specifcally consider mitigat-
ing factors or measures such as fne refuse deposits that may be associated with a slurry impound-
ment. These guidelines are important historical literature and should be reviewed as part of initial
evaluations; however, it is recommended that subsidence analyses and prediction models be used
for evaluating potential deformations and strains on embankments and impoundments. On the basis
of these analyses, smaller overburden bufers than identifed in the aforementioned references may
be sufcient to prevent a breakthrough to a reservoir or slurry impoundment and may satisfactorily
limit impoundment leakage/seepage.
A number of subsidence prediction models are available. Two commonly used models include SDPS
(Surface Deformation Prediction System) developed at Virginia Polytechnic Institute and State Uni-
versity and CISPM (Comprehensive and Integrated Subsidence Prediction Model) developed at West
Virginia University. These programs can be used to estimate surface subsidence and ground strain
caused by mining. Two cautions are ofered concerning the use of these types of programs: (1) they
should only be used for the type of topography and mining conditions for which they were developed
and (2) the strain computations by these programs typically do not account for strain concentration
along existing discontinuities. Concerning the second point, studies have shown that site topogra-
phy may have a substantial efect on the development and concentration of horizontal strain.
Furthermore, where the confnement and continuity in the overburden is diminished, such as on
hillsides and in highly weathered and fractured material, the tensile strain that is induced by mining
may accumulate along one or more of the joints rather than being more evenly distributed. This can
be signifcant in subsidence and breakthrough potential evaluations where open joints can provide
seepage pathways. For total extraction mining, the strata and ground surface above the mine are
afected regardless of the mining depth, and the efect of tensile strains and strata disturbance on
dam stability and breakthrough potential should be evaluated.
8.4.2.8 Failures Related to Auger and Highwall Mining
Auger or highwall mining openings in an abutment can have adverse efects on an impounding
embankment. Deformation of the abutment and embankment can occur if the webs between the
holes deteriorate or fail. The holes can also provide seepage paths. Breakthroughs can occur through
auger or highwall mining holes, through the coal remaining at the ends of auger holes, or as a result
< PREVIOUS VIEW
Chapter 8
8-24 MAY 2009
of the collapse of these holes. The importance of identifying the locations of these holes during site
exploration cannot be overemphasized. During highwall reclamation these holes are typically cov-
ered without backflling. If they are not discovered and accounted for in the impoundment design,
then seepage pressures under high impoundment head can cause movement of backfll and water
into auger holes. If the auger holes are accidentally connected to the mine, then there is a direct path
for a breakthrough or piping. If the auger holes were terminated close to the mine workings, the
remaining coal barrier can potentially fail in a shear or punch-type mode. Also, the narrow pillars or
webs of coal between these holes may be marginally stable and may subsequently deteriorate and
collapse, resulting in highwall instability and ground deformations that could adversely afect the
embankment or lead to a breakthrough. NIOSH has a computer program (ARMPS-HWM) that per-
forms an empirical analysis of highwall mine pillars.
8.4.2.9 Hillside Movement or Disturbance
Hillside movement and disturbance (landslide, creep, or human impacts) can occur at and above a coal
outcrop. Weathering tends to reduce the strength of surface soils and rocks while gravity provides a
driving force to move soil and rock down the hillside. Surface mining (contour mining) and site devel-
opment activities, such as road and channel construction, can accelerate the process through removal
of soil and rock from the hillside or by aggravating or causing landslides. Any movement or removal of
soil and rock from the hillside can reduce and/or disturb the outcrop and overburden barriers between
the underground mine and an overlying impoundment, thus contributing to the occurrence of a sink-
hole, internal erosion, or shear-type failure. Figure 8.7 illustrates the efect of hillside disturbance on the
coal outcrop barrier.
8.4.2.10 Mine Blowout
A mine blowout occurs when water impounded in an underground mine breaks out through either
a sealed mine opening or a thin section of the coal outcrop barrier. Such a blowout may be a sec-
ondary consequence of an impoundment breaking into an underground mine. Additionally, a mine
blowout in an impoundment watershed may represent an infow source that should be considered
in impoundment design. Such an event could cause a signifcant amount of water to fow into the
impoundment, damage impoundment structures, and potentially afect impoundment safety. A
guidance manual for design published by the Ofce of Surface Mining (OSM) is, Outcrop Barrier
Design for Above Drainage Coal Mines, by Kohli and Block (2007). This document presents com-
piled information on case histories as well as design studies by Pearson et al. (1981). If seepage from
an underground mine is occurring at an embankment or impoundment site, the conditions and
potential efects should be evaluated. As a minimum, the facility design should account for collect-
ing seepage and discharging it in a controlled manner that will not adversely afect the embankment
or other facility structures.
8.4.2.11 Seismic and Blasting Events
Typically, earthquakes and surface blasting do not afect underground mines. The exception is
when a fault crosses through a mine or when surface waves can afect the stability of the portals at
or near their contact with the surface. The presence of an active fault passing through a U.S. coal
mine is not a situation that normally occurs. If such a fault is identifed, it would be necessary to
conduct comprehensive stability analyses not just for the mine but also for the associated surface
facilities including the refuse embankment and impoundment. Bhabdaru and Arora (1987), Nich-
olls et al. (1971), Rupert and Clark (1977), Fourie and Green (1993), and Fernandez and Van Der
Heever (1996) describe blasting and earthquake impacts on mines.
When blasting is planned within 500 feet of an active underground mine, the Surface Mining
Control and Reclamation Act of 1977 requires that the Operators blasting plan be approved by
OSM (or appropriate state agency) and MSHA. Blasting regulations are provided in 30 CFR
< PREVIOUS VIEW
8-25
Site Mining and Foundation Issues
MAY 2009
780.13 and 30 CFR 816.61 through 816.68 and also 816.79. State criteria may also be applicable.
The blasting plan should be prepared by a professional licensed in the state where the blasting is
to be performed. Potential concerns when an impoundment is present include fracturing of abut-
ments, impacts to pipes and other rigid structures, or possibly upstream construction. Where an
impoundment is present, the potential impacts should be evaluated, and monitoring of particle
velocity with a seismograph may be appropriate. Section 6.6.7 addresses potential blasting impacts
on impoundment structures.
8.4.3 Mine Subsidence Potential and Analysis
8.4.3.1 Pillar Evaluation and Analysis
For room and pillar mining, subsidence potential can be evaluated through analysis of the overbur-
den stress imposed on coal pillars and comparison of this stress to the pillar strength. To determine
the average stress on a coal pillar, the tributary area approach may be used with the assumption that
the overburden pressure increases at a rate of 1.1 psi per foot of depth:
S
p
= 1.1 H[(w + B) /w] [(L + B) / L] (8-1)
where:
S
p
= average pillar stress (psi)
H = depth below ground surface (f)
w = pillar width (f)
L = pillar length (f)
B = opening width (f)
Other assumptions inherent in this formula include:
The coal seam is subject only to vertical pressure.
Each pillar supports the column of rock and other overburden overlying the pillar
and a proportionate share of surrounding openings.
The load is distributed uniformly over the pillar cross section.
Most empirical pillar equations have been developed and the results statistically analyzed based on
these assumptions, particularly with respect to tributary area loading. Therefore, empirical equa-
tions should only be used for uniformly distributed tributary loading. Numerical models such as
LAMODEL can be used to analyze non-uniform pillar loading (transfer of load to fanking barrier pil-
lars). However, if pillars are to be analyzed without considering tributary area loading, then the pillar
loads should be determined through a detailed rock mechanics analysis using numerical modeling
sofware. Typically, a higher safety factor is used for such a pillar analysis because of uncertainty as
to the loading and the variability in the numerical methods typically employed.
Pillar strength formulas used for design of ground control measures are discussed in Bieniawski
(1992). For subsidence evaluations for coal refuse disposal facilities, pillar strength formulas consis-
tent with conditions in the underlying mine should be adopted; otherwise formulas applicable to
local mines should be employed. Several empirical formulas that have been used to determine pillar
strength are presented in Figure 8.8. The Pitsburgh Coal seam with a pillar height of 10 feet is used
as an example of the infuence of the width to height ratio on the overall pillar strength. The potential
for subsidence can be determined by dividing the pillar strength by the pillar load. An important
consideration when comparing the pillar strength formulas is the value of the recommended stabil-
< PREVIOUS VIEW
Chapter 8
8-26 MAY 2009
FIGURE 8.8 COMPARISON OF PILLAR STRENGTH FORMULAS
WITH RESPECT TO WIDTH-TO-HEIGHT RATIO
PILLAR WIDTH-TO-HEIGHT RATIO, w/h
= Experimental data from actual mine
1 2 3 4 5 6 7 8 9 10 11 12
2
3
4
5
6
BIENIAWSKI FORMULA p = 0.64 + 0.36 (w/h)
p
=
w
/h
O
B
E
R
T
-D
U
V
A
L
L
F
O
R
M
U
L
A
p
=
0
.7
8
+
0
.2
2
(w
/h
)
1
10 20 30 40 50 60 70 80 90
0
1000
2000
3000
PILLAR WIDTH, w (feet)
P
I
L
L
A
R
S
T
R
E
N
G
T
H
,
p
(
p
s
i
)
PITTSBURGH SEAM
p
=
k w
h
O
B
E
R
T
-D
U
V
A
L
L
F
O
R
M
U
L
A
w
h
p
=
1
(0
.7
7
8
+
0
.2
2
2
)
B
I
E
N
I
A
W
S
K
I
F
O
R
M
U
L
A
p
=
1
(
0
.
6
4
+
0
.
3
6
)
w
h
H
O
L
L
A
N
D
F
O
R
M
U
L
A
p
=
1
w
h
S
A
L
A
M
O
N
-M
U
N
R
O
p
=
w
0.46
h
0.66
k
1
2
1 = SAMPLE STRENGTH = 930 psi
k = SCALING FACTOR = 5580
PILLAR HEIGHT = 10 FT.
(BIENAWSKI, 1987)
N
O
R
M
A
L
I
Z
E
D
P
I
L
L
A
R
S
T
R
E
N
G
T
H
,
1
/
FIGURE 8.8 COMPARISON OF PILLAR STRENGTH FORMULAS WITH RESPECT
TO WIDTH-TO-HEIGHT RATIO
FIGURE 8.8 COMPARISON OF PILLAR STRENGTH FORMULAS WITH RESPECT
TO WIDTH-TO-HEIGHT RATIO
< PREVIOUS VIEW
8-27
Site Mining and Foundation Issues
MAY 2009
ity factor, which varies depending on the empirical formula used (Bieniawski, 1992). Mark (2006), in
addressing the stability factor for the Mark-Bieniawski Formula, recommends a stability factor of 1.5
for pillars with under 750 feet of cover and 0.9 for pillars with over 1,250 feet of cover. As previously
noted, the term stability factor should not be confused with safety factor. While some long-term
pillar stability can be tolerated in certain mining situations, impoundment designers should consider
a higher margin of safety for overburden support and control of detrimental diferential movement
within the dam and foundation.
While the occurrence of subsidence over mine workings in some areas is a function of time, the gen-
eral assumption for pillar analyses is that coal does not deteriorate over time. The reason is that coal is
highly jointed due to cleating, and additional cracking of the coal does not reduce the confnement of
the roof and foor, nor does it markedly reduce the coals mass strength. Initial pillar designs typically
take into account spalling for coal seams susceptible to spalling. This is done by reducing the efective
dimensions of the pillars by the anticipated spalled width and widening the opening width by this
spalled dimension. Oxidation of the coal can occur for seams above drainage, but this efect tends to
be limited to the yielded outer portions of the pillar. Once pumping and ventilation is discontinued,
mines below drainage will food, thus mitigating further deterioration of the coal.
The weathering of sof (high-clay-content) partings (non-coal layers in the pillars) reduces pillar
strength over the long term, although the impact appears minor and may not occur afer the mine is
fooded (Biswas et al., 1995). Biswas et al. (1999) presented an approach for estimating time-depen-
dent deterioration of partings, as well as coal. The authors indicated that much work remains before
time efects on coal pillar strength are fully understood.
Computer programs for pillar stability analysis include the ARMPS and LAMODEL models. These
programs can be used to evaluate stress distribution and factor of safety. Analyses can be performed
considering the pillar sizes and arrangement within the mine, recognizing that some smaller pillars
may shed load to larger pillars.
In addition to pillar stability, foor stability is also related to subsidence potential. Bearing capacity
analyses have been applied to evaluation of pillar punching into sof foor materials, using the follow-
ing formulas (Bieniawski, 1992; Brady and Brown, 2004):
For long (L/B > 10) pillars:
Q
u
= 0.5 (
BN
1
) + ( cN
c
) (8-2)
For pillars with L/B < 10:
Q
u
= 0.5 (
BN
1
S
1
) + ( c cot N
q
S
q
c cot ) (8-3)
where:
N
c
= (N
q
1) cot
N
1
= 1.5 (N
q
+ 1) tan
N
q
= exp [ tan ] [ tan
2
(/4 + /2]
S
1
= 1.0 0.4 (B/ L) (dimensionless shape factor)
S
q
= 1.0 sin (B/ L) (dimensionless shape factor)
slurry
w
)
2S
F
rock
=
[
w
2
]
(
rock
w
)
2S
< PREVIOUS VIEW
8-31
Site Mining and Foundation Issues
MAY 2009
F
coal
= [ w h
seam
] (
coal
w
)
F
water
= h
w
h
seam
w
and:
F
slurry
= efective weight of slurrry (lb/f)
F
rock
= efective weight of overburden rock (lb/f)
F
coal
= efective weight of coal (lb/f)
F
water
= efective force of water (lb/f)
F
K
o
= lateral force on barrier from overburden (lb/f)
h
w
= net hydrostatic head on coal seam (f)
h
seam
= coal seam height (f)
w
= unit weight of water (pcf)
slurry
= unit weight of slurry (pcf)
rock
= unit weight of overburden rock (pcf)
coal
= unit weight of coal (pcf)
= shear strength for barrier contact (degrees)
w = outcrop barrier width (f)
S = ground slope at outcrop (dimensionless)
The efective lateral force on the barrier from overburden F
K
o
is given by the following relationship,
based on the properties of the slurry and conservatively treating the soil/weathered coal as slurry:
F
K
o
= [ (
slurry
w
) h
slurry
h
seam
] ( 1 sin
slurry
) (8-5)
where:
h
slurry
= depth of slurry to barrier (f)
h
seam
= coal seam height (f)
slurry
= shear strength of slurry (degrees)
The preceding equations can be modifed to include resistance along the roof-coal interface and to
incorporate protective embankments at the outcrop. In applying the analysis, the weight of overbur-
den rock and coal may be taken as the total weight only if the mine workings are not and will not be
fooded. The efective (or buoyant) weight is applicable if the coal and overburden are afected by
seepage and the mine workings are fooded. Conservative assumptions associated with this analysis
include:
Hydrostatic impoundment head is assumed to be acting at the barrier.
There is no cohesion along the coal-foor interface, and the resistance along the roof-
coal interface is ignored consistent with Figure 8.9.
Three-dimensional efects are not considered (side resistance of the wedge of coal
should generally be ignored because of the potential for discontinuities).
< PREVIOUS VIEW
Chapter 8
8-32 MAY 2009
Selection of the barrier width w requires a thorough evaluation of mine maps and exploration infor-
mation to identify the representative minimum dimension, based on the following:
Extent of weathering of the outcrop, which can best be evaluated by exploration and
sampling.
Potential for spalling and deterioration of the barrier rib within the mine work-
ings, which can best be judged based on pillar performance within the mine.
Biswas et al. (1999) introduce a method for estimation of the decline of pillar
strength with time.
FIGURE 8.9 OUTCROP BARRIER BREAKTHROUGH
ANALYSIS FOR SHEAR FAILURE
8.9b SIMPLIFIED FREE-BODY DIAGRAM OF EFFECTIVE BARRIER SEGMENT
(ADAPTED FROM NEWMAN, 2003)
8.9a ILLUSTRATION OF OUTCROP BARRIER
MINE VOID MINE WORKINGS
SLURRY
SOIL OVERBURDEN
ROCK OVERBURDEN
UNWEATHERED COAL
POOL LEVEL
hSLURRY
hSEAM
BARRIER SEGMENT
w
WEATHERED
COAL
OUTCROP BARRIER
S
1
NOTE: SLURRY AND SOIL ARE
TREATED TOGETHER
AS SLURRY.
FSLURRY
FCOAL
FROCK
FWATER
SHEAR RESISTANCE = (VERTICAL FORCES) TAN
MINE VOID
FKO
= (w hSLURRY) - (w
2
/2S) (SLURRY - W)
= w
2
/2S
(ROCK - W)
(COAL - W)
w hSEAM
=
FIGURE 8.9 OUTCROP BARRIER BREAKTHROUGH ANALYSIS FOR SHEAR FAILURE
FIGURE 8.9 OUTCROP BARRIER BREAKTHROUGH ANALYSIS FOR SHEAR FAILURE
< PREVIOUS VIEW
8-33
Site Mining and Foundation Issues
MAY 2009
Discontinuities in coal seam or overburden (e.g., stress relief fractures to surface) that
may allow hydrostatic pressures to reach the interior of the barrier, the potential for
which can be evaluated based on local geologic studies and surfcial reconnaissance
above the outcrop.
The preceding analysis is based on the assumption of a horizontal outcrop barrier. Adjustments to
account for the dip of the coal seam and barrier may be warranted as well as evaluation of possible
wedge failure across overburden bedding planes where fracture systems with such orientation are
present.
If a man-made barrier instead of coal is being evaluated, the preceding analyses should be performed
using the properties of the barrier. The following guidance is provided:
Earthen materials and aggregate barriers The strength properties of earthen materi-
als or aggregate should be established based upon how the material was placed to
form the barrier and accounting for resistance at the barrier-foor interface only (i.e.,
resistance between the barrier and the coal seam or roof is ignored).
Concrete, concrete block, and grouted aggregate barriers Construction details
associated with the barrier keys should be included in the analysis of resistance to
failure, along with the interface friction between the barrier and foor/roof, as subse-
quently discussed in Section 8.5.
Fractures in overburden above an outcrop may have the potential for seepage and internal erosion
that could propagate to an impoundment and lead to breakthrough. Natural stress relief fractures, if
close to mine workings or subsidence cracks can lead to the potential for breakthrough from internal
erosion. Breakthrough potential for this mode of failure can be determined by evaluating: (1) the
presence of or potential for fractures and their infuence on mine stability, (2) the occurrence and
magnitude of subsidence strains and deformations, and (3) the type and efectiveness of any mitiga-
tion measures employed. Mitigation measures may include features to control susceptibility to inter-
nal erosion such as self-healing granular soils within the overburden and design of flls and flters to
prevent piping.
8.4.4.2 Overburden Barriers
When mine workings extend below drainage and beneath the botom of an impoundment, atention
should be focused on the overburden, roof, pillar, and foor stability, as illustrated in Figure 8.10. The
potential for breakthrough could be associated with the following scenarios:
Sinkhole subsidence under relatively thin overburden strata where a roof fall propa-
gates into the overburden and eventually daylights into the impoundment.
Fracturing of the overburden due to pillar collapse or foor failure, leading to inter-
nal erosion propagating to the impoundment.
Analysis of roof stability and sinkhole development is addressed in Section 8.4.2. The potential for
these phenomena can be determined empirically using the Rock Mass Rating (RMR) (Bieniawski,
1992) or the Coal Mine Roof Rating (CMRR) (Molinda et al., 2001) and analytically by calculating
beam or arch safety factors for each stratum. The analysis of pillar and foor stability, which was
addressed in Section 8.4.3, is based on the stability factors of pillars that may lie within a reasonable
angle of draw around the periphery of the impoundment, considering room and pillar dimensions,
mining height, overburden, and coal strength. Newman (2003) presents an analysis methodology
and associated case history for evaluation of breakthrough potential for an overburden barrier.
< PREVIOUS VIEW
Chapter 8
8-34 MAY 2009
8.5 SUBSIDENCE AND BREAKTHROUGH MITIGATION METHODS
Mitigation measures to address the potential impacts of subsidence or breakthrough may include
some combination of:
Providing a safety zone around the embankment and impoundment, limiting
mining to only entry development beneath the dam.
Providing support by backflling portions of the mine.
FIGURE 8.10 ILLUSTRATION OF OVERBURDEN BARRIER
8.10a ILLUSTRATION OF OVERBURDEN BARRIER
FIGURE 8.10 OVERBURDEN BARRIER ILLUSTRATION
ROCK OVERBURDEN
MINE WORKINGS IN COAL SEAM
SLURRY
SOIL
8.10b SCHEMATIC PLAN OF PILLARS
PILLAR TRIBUTARY
SUPPORT AREA
B
B
8.10c TYPICAL PILLAR DETAIL
W
L
0.5B
FIGURE 8.10 ILLUSTRATION OF OVERBURDEN BARRIER
< PREVIOUS VIEW
8-35
Site Mining and Foundation Issues
MAY 2009
Improving the in-situ materials by grouting.
Constructing an engineered barrier.
Isolating the structure from the area of infuence of the mining or altering the
mining sequence or plan.
Constructing secondary defense measures such as bulkheads to contain a
breakthrough within the mine.
Other engineered measures or impoundment operating procedures to control
seepage and reduce pressures in the areas of potential breakthrough.
Subsidence mitigation measures are dependent on the structure that must be protected. Mine work-
ings beneath a dam must be stabilized such that support for the impounding embankment and abut-
ments is achieved. Overburden or outcrop barriers beneath an impoundment must be sufcient to
prevent breakthrough. Mitigation measures for breakthroughs are discussed in MSHA (2003) and
NRC (2002) and are summarized in the following sections, which describe potential mitigation alter-
natives and methods for the design of bulkheads to seal mine entries.
8.5.1 Mine Subsidence and Breakthrough Mitigation
8.5.1.1 Use of Safety Zones
For new impoundments, the most efective method for preventing damage to the dam or a break-
through is to leave an unmined safety zone between the mine workings and the impoundment
so that any mining-induced ground disturbance cannot cause a breakthrough or other signifcant
adverse efects. At new facilities, siting of the impoundment at locations that are not or will not
be undermined is preferred. If mine workings cannot be avoided, other mitigation measures may
be feasible provided that support for the impounding embankment and appurtenant structures
is achieved.
For existing impoundments where the mine workings are already close enough to potentially cause
a problem, a safety zone can be created (if necessary) by backflling the mine workings. Guidelines
for sizing safety zones around impoundments are provided in Babcock and Hooker (1977). Figure 8.6
provides an illustration of these guidelines. Kendorski et al. (1979) provide criteria for determining
when a surface water body represents a hazard to mining. Peng and Luo (1993) provide guidance for
establishing safety zones for sensitive structures.
8.5.1.2 Mine Backflling
If the thickness and natural characteristics of the overburden barrier cannot be relied on to prevent
a sinkhole, subsidence cracks, or other subsidence-related failure mechanisms, flling previously-
mined areas with grout or other material (commonly referred to as stowing) may be a necessary
remedy. The lateral extent of backflling must include critical areas within the angle of draw based on
the results of subsidence analyses described in Section 8.4.2.
The mine backfll material can have minimal strength, and even a partial backfll will ofer confne-
ment to mine pillars, thereby reducing spalling and dramatically increasing pillar strength. However,
it is good practice to backfll to the roof of the mine, using material with sufcient strength (typically
above 100 psi) to reduce consolidation and prevent erosion. This can be difcult to accomplish, as full
contact with the mine roof is ofen not possible because of roof irregularity and the rolling nature of
coal seams. However, partial roof contact will reduce roof falls and will dramatically limit roof fall
propagation into the overburden.
Pozzolan slurry makes a good backfll, because it readily fows into irregular spaces. Because of the
high sulfate environment in a coal mine, any fll containing cement should employ sulfate-resistant
< PREVIOUS VIEW
Chapter 8
8-36 MAY 2009
cement, preferably Type V, but in some mine environments Type II cement will sufce. Depending
on cement availability, Type II or Type I cement in combination with pozzolans (e.g., fy ash, slag
cement) may be required in order to provide adequate sulfate resistance. The extent of the backflled
areas must be sufcient to support the overburden and/or hillside that could potentially be inun-
dated by the impoundment and to protect existing or planned embankment dams and mine seepage
barriers from adverse subsidence efects. Boreholes should be advanced to determine if the backfll-
ing program has achieved performance criteria. These boreholes can also be used to locate remaining
voids, to perform secondary grouting, and to obtain samples from within the backflled mine work-
ings for strength testing.
To verify that the backfll strength and coverage is adequate, subsidence and/or pillar analyses should
be conducted as part of the mine backflling design. As a minimum, the following should be specifed
as part of the backfll design:
Strength of the backfll material
Area to be backflled
Methods for verifcation that the design strength
and area of backflling are acceptable
When used with cement, most types of fy ash improve fowability, increase sulfate resistance, and
sometimes increase the compressive strength of backfll grout. Fly ash improves the fowability of
grout because the spherical particles act like ball bearings that allow the grout to move more freely
and the small particle size promotes beter flling of voids. In addition, the pozzolanic properties of
most types of fy ash increase compressive strength. Use of fy ash also reduces shrinkage and slows
setup time, which is important if grout pumping must be interrupted for a few hours. The proper-
ties of fy ash will vary dependent upon the coal source used at the originating power plant. Fly ash
properties are generally determined by the power companies that generated it and these properties
can ofen be obtained from them.
Fly ash can pose potential environmental and health risks because it contains trace amounts of toxic
metals such as boron, molybdenum, selenium, and arsenic. Portland cement also contains these ele-
ments, and they can occur naturally in soil and water. (USEPA, 2000) discusses environmental impact
considerations associated with the use of fy ash in mine environments.
8.5.1.3 Stowing of Mine Openings and Associated Barrier Construction
It has been common practice to seal mine entries by stowing them with competent rock or other fll
in conjunction with constructing a compacted earth or coarse coal refuse embankment on the out-
side against the openings. This approach can also be used with auger holes. The openings should be
flled far enough back into the mine to prevent adverse subsidence or sinkholes that could extend to
the ground surface under the dam or impoundment. Typically, aggregate or other rock materials are
pushed or rammed into openings to the extent practical with construction equipment. If possible,
stowing should extend into the opening for a distance sufcient to mitigate subsidence. In some
applications, including areas beneath a dam or abutment, grouting is performed in order to extend
the backflling of the opening, support the roof, and mitigate seepage. In an impoundment area,
when it is impractical or infeasible to mitigate subsidence by stowing, other options can be consid-
ered such as: (1) earthen barriers designed to provide protection against potential sinkholes or cracks
and (2) overexcavation to remove shallow overburden subject to sinkholes and establish a highwall
cut. Monitoring systems for detecting movement and/or hydraulic pressure should be considered.
Earth or coarse coal refuse barriers should be protected by appropriate geotextile or graded flters
to prevent piping into the mine, open joints, and stowing material. Drains should be installed as
< PREVIOUS VIEW
8-37
Site Mining and Foundation Issues
MAY 2009
FIGURE 8.11 STOWED MINE OPENINGS AND BARRIER CONSTRUCTION
MINE OPENING
STOWED WITH ROCK
GEOTEXTILE
COMPACTED
BACKFILL
8.11a BACKFILL BARRIER WITH GEOTEXTILE
OVERBURDEN
COAL SEAM
WITH WORKINGS
FINE AGGREGATE
(e.g., AASHTO NO. 57)
8.11b BACKFILL BARRIER WITH GEOMEMBRANE
OVERBURDEN
COAL SEAM
WITH WORKINGS
GEOMEMBRANE
COMPACTED
BACKFILL
FINE AGGREGATE
(e.g., AASHTO NO. 57) MINE OPENING
STOWED WITH ROCK
FINE AGGREGATE
CUSHION LAYER
(e.g., ASTM C33)
needed to control the hydrostatic pressure. The system of the stowed and external barrier materi-
als must be designed to have sufcient strength and dimensions to resist shear or punching failure
into the mine resulting from hydraulic and earth-load forces. The stowed material must also have
sufcient bulk and material gradation to provide adequate seepage resistance and thus prevent
failure of the embankment or adjoining strata due to piping or hydraulic fracturing into the mine.
Figure 8.11a illustrates a stowed opening protected with geotextile and backfll. The fne-aggregate
flter shown in the fgure should cover the area where fractures or open joints are present or are
likely to be a concern.
If a coarse coal refuse embankment is to be constructed over a coal seam outcrop with workings, the
exposed coal seam must be covered and sealed with soil or other inert material to provide a fre bar-
rier and to minimize the potential for spontaneous combustion. If water is draining out of the mine
from an opening that is to be covered, a drain to prevent water from building up in the mine may be
needed. For these cases, the sealing cover should include a drainage system that will release the mine
water and prevent hydrostatic pressures from building up and causing problems with respect to
saturation, piping, or structural instability. A discussion of internal drain and flter design is provided
in Section 6.6.2. If the mine discharge is acidic, drain materials capable of resisting degradation will
be required (Section 11.7), and downstream water collection and treatment could also be required
(Section 10.3.3).
FIGURE 8.11 STOWED MINE OPENINGS AND BARRIER CONSTRUCTION
FIGURE 8.11 STOWED MINE OPENINGS AND BARRIER CONSTRUCTION
< PREVIOUS VIEW
Chapter 8
8-38 MAY 2009
Impervious membranes have been used in conjunction with mine opening seals. A membrane should
extend a sufcient distance past the perimeter of an opening to provide an efective barrier encom-
passing mining-induced fractures and open joints. To prevent seepage from fowing around the
membrane, the edges of the membrane should be anchored or embedded. The membrane should be
surrounded by a layer of fner-grained, cushioning material to prevent puncture by sharp rocks that
may be present in the embankment or highwall. Figure 8.11b illustrates the use of a geomembrane as
part of a mine opening seal. Liner systems are discussed Section 10.4.
If mine openings are safely accessible to workers or can be rehabilitated so that they are safely acces-
sible and plans for entry are submited and approved by MSHA (30 CFR Subchapters G, H I and
O), form-work or bulkheads can be constructed to restrict the extent of stowed material to a desired
depth into the mine. In this situation, workers could enter the mine to install grout pipes or to do
other needed work. This approach has been used successfully for both the pneumatic stowing and
grouting of gravel and for installation of a grouted rock plug. Pneumatic stowing and grouting meth-
ods are described in the U.S. Bureau of Mines Information Circular 9359 (Walker, 1993). The follow-
ing considerations apply: (1) some material types or uncontrolled (uncompacted) placement can be
subject to erosion under water fow and (2) the operation exposes personnel to dust. These concerns
can be overcome by grouting, although the use of cement adds cost, and environmental impacts
can be an issue. If the mine opening can not be made safely accessible to workers, rams mounted on
heavy construction equipment can be used to pack the stowing material into the opening, but the
depth to which the opening can be backflled is limited.
In addition to stowing the mine opening, measures to isolate the sealed opening from the impound-
ment and to control hydraulic pressure may also be needed. A compacted embankment fll with
an internal drain protected by a flter discharging beyond the limits of the impoundment provides
additional protection and redundancy for mitigating breakthrough. Figure 8.12 shows an example of
a compacted fll (prior to refuse placement) placed against an outcrop and with a drain wrapped in
geotextile and spoil. Drain and flter design requirements are discussed in Section 6.6.2.
8.5.1.4 Bulkhead Construction
Construction of bulkheads inside a mine or at mine portals can be a feature of a breakthrough pre-
vention program. Bulkheads are designed to withstand fuid and/or earth pressures. They may be
relatively thin reinforced-concrete structures or relatively thick structures consisting of concrete,
grouted rock or polyurethane foam.
COAL SEAM WITH WORKINGS
BENCH
REFUSE
SPOIL
LOW PERMEABILITY BARRIER DRAIN
GEOTEXTILE
FIGURE 8.12 BARRIER AGAINST OUTCROP WITH LOW
PERMEABILITY FILL AND DRAIN
(ADAPTED FROM COWHERD ET AL., 2002)
OVERBURDEN
FIGURE 8.12 BARRIER AGAINST OUTCROP WITH LOW-PERMEABILITY FILL AND DRAIN
FIGURE 8.12 BARRIER AGAINST OUTCROP WITH LOW-PERMEABILITY FILL AND DRAIN
< PREVIOUS VIEW
8-39
Site Mining and Foundation Issues
MAY 2009
In situations where there is uncertainty about the level of breakthrough protection provided by other
preventative measures, remote bulkheads can be used as a secondary defense against the discharge
of water or slurry through and from a mine. These bulkheads are constructed in mine openings
where the water/slurry would fow to (and fow out of) in the event of a breakthrough. For such an
application, designers need to consider the rapid build up of air pressure and subsequent impact of
water and debris against the secondary structure. Furthermore, the impounding water or slurry in a
mine creates the potential for a blowout of an outcrop barrier or bulkhead that should be evaluated.
The design of bulkhead seals is presented in Section 8.5.2.
8.5.1.5 Construction of Compacted Earthen Barriers on the Surface
Construction of a compacted earth-fll barrier around the botom or perimeter of the impoundment
area is a design measure that has been used to mitigate potential breakthrough conditions. A com-
pacted earth-fll barrier provides additional bulk between the impoundment and the mine workings
and, in combination with properly designed internal drainage, can lower the water pressure against
an outcrop barrier. The water pressure can be reduced if internal drains are used to draw down the
hydrostatic level in the fll and also to provide an outlet so that seepage discharges in a controlled
manner. Compacted earth-fll barriers can be placed and raised as the impoundment level rises. The
design of a compacted earth-fll barrier should take into account the potential efect of subsidence,
including sinkholes resulting from underlying mine workings. Measures that have been incorpo-
rated into barriers to resist potential subsidence movement, cracking and internal erosion include
geogrid reinforcement and graded flters.
For designing a compacted earth-fll barrier as a breakthrough prevention measure, one approach
is to conduct reconnaissance and excavation of the coal seam in the vicinity of the reservoir that is
known or suspected to be constructed over mine voids. The concept is to identify and intercept any
mine voids, and thereby expose known or suspected workings near the outcrop. Weathered or frac-
tured strata near the coal seam outcrop should be removed as part of the excavation, thus allowing
the barrier to be founded on competent rock. Surface mining of suspect areas is a comprehensive
approach for addressing large areas with unknown conditions. The stability of highwalls should also
be evaluated, particularly when workings are encountered during construction.
Another approach is to construct the earth-fll barrier over the existing coal seam outcrop without
excavation of the seam. The support and internal stability of the earth-fll barrier due to the presence
of mine voids should be evaluated. For this approach, there may be exposed highwall if the coal seam
has previously been surface-mined, augered or highwall-mined, or there may be an undisturbed out-
crop that has not been afected by surface mining.
The construction of compacted earth-fll and coarse coal refuse barriers should be based upon the
same concepts regarding material selection and engineering properties, foundation preparation, pro-
visions for internal drainage, slope stability and development of construction specifcations used for
the design and construction of earth-fll dams and refuse embankments. Where feasible, fne coal
refuse slurry should be deposited around these barriers, as the development of a delta of fne coal
refuse above the normal pool level will provide additional resistance to breakthrough.
8.5.1.6 Conversion to Slurry Cells
As described in Section 3.4, slurry cells can be used to dispose of fne coal refuse. One approach that
can be taken to reduce the breakthrough hazard is to convert from a full slurry impoundment into
a slurry cell confguration using compacted coarse refuse to construct small, individual cells or a
series of cells over previously placed fnes. Some cell designs have reached depths of 12 feet (8 feet
of fne refuse covered by 4 feet of compacted coarse refuse). The depth and number of active cells
(i.e., cells that are not capped with backfll) at any given time is usually determined by the volume of
< PREVIOUS VIEW
Chapter 8
8-40 MAY 2009
storage and cell confguration that will result in a low-hazard potential classifcation for the facility,
as described in Section 3.4. The beneft is that the coarse refuse dikes and covering layers combined
with the thin layers of fnes allow the fnes to dewater and consolidate and make the total mass less
fowable. Additionally, with the fnes compartmentalized, a problem at one cell location is less likely
to afect the entire facility or result in a catastrophic event.
The downstream containment structure (i.e., structural zone) for slurry cells is designed in the same
manner as a dam embankment with appropriate width, slopes, benches, internal drainage system,
and embankment-material strengths needed to achieve suitable safety factors for slope stability.
Slurry cells are most efcient when the depth of fnes in the cells is kept relatively shallow, thus pro-
moting drainage from the fnes before and during covering.
Some disadvantages of slurry cells include:
Frequent construction of diversion ditches, new cells, and cell spillways is required
as the site elevation increases.
There is limited fexibility in that a relatively large ratio of coarse refuse to fne refuse
is required to keep cell construction ahead of slurry placement, and the fne refuse
must setle quickly with limited clarifed water retention.
Close planning and supervision of the site is required so that the construction, flling,
and backflling of cells is accomplished in the proper sequence to make the system
function as intended.
Slurry cell operations are not generally compatible with high production rates at
some coal preparation plants.
Slurry cell operations are not generally compatible with sites that employ conveyor
belt/dozer push disposal techniques.
If a slurry cell site is large enough to be classifed as having high-hazard potential,
the facility spillways must be designed to handle the runof from the associated
design storm (e.g., PMF).
8.5.1.7 Sealing Sources of Leakage
Discontinuities such as open joints or cracks in a refuse embankment foundation or impoundment
area may be treated by grouting or other measures to prevent them from transmiting high hydrostatic
pressures and to eliminate potential paths for internal erosion. As a secondary protection measure
against leakage, fne coal waste (slurry) can be deposited to form a delta that provides an additional
layer of material between the impoundment and potential seepage problem areas. This technique
has been used successfully to reduce leakage through embankment and abutment areas and around
the perimeter of the impoundment. The long-term beneft of this secondary measure is limited to
situations where the bedding planes or joints are relatively small and the sand-sized portion of the
slurry is sufcient to allow the gradual formation of a natural flter in joint openings and the eventual
sealing of the openings with the clay- and silt-sized particles. If a seepage problem develops afer
an impoundment is constructed and the technique of distributing the slurry upstream of the seep-
age area does not correct the problem, then grouting, construction of an impervious barrier, or other
measures may need to be taken.
8.5.1.8 Stabilization of Fines
The potential for breakthrough of the contents of a slurry impoundment into underground mine
workings can be signifcantly lessened through stabilization of the fne refuse. Stabilization alter-
natives include providing drainage measures or treating the fnes with additives to increase their
strength and/or reduce their water content.
< PREVIOUS VIEW
8-41
Site Mining and Foundation Issues
MAY 2009
Stabilization of coal-waste fnes has sometimes been accomplished by the addition of portland cement
or lime-based products. Since the late 1970s and continuing more recently (Fiscor, 2002), studies for
development and evaluation of the performance of stabilizer additives have been undertaken. Labo-
ratory studies performed in the late 1970s and early 1980s showed that mixing fne coal refuse with
lime-based products may cause the material to appear to have increased strength and stifness in
comparison to untreated material. Afer samples were mixed and cured, the treated refuse exhibited a
relatively stif load-deformation behavior during initial loading, but then when loading exceeded the
apparent maximum past consolidation pressure, the stabilized refuse collapsed and the load-defor-
mation characteristics returned to the previous unstabilized behavior. In addition, these laboratory
studies showed that 5 percent or more by weight of the stabilizers were needed, which made the addi-
tives prohibitively expensive except for small treatment volumes. While not all stabilizer additives
result in similar behavior, they should always be carefully investigated and used with caution.
Shallow or deep soil mixing methodology using fy ash and cement grout are also methods for sta-
bilizing fne deposits that would be efective for mitigating breakthrough potential. Although not
for the purpose of breakthrough mitigation, a portion of a slurry impoundment in Pennsylvania was
stabilized in place by shallow and deep mixing with fy ash and cement grout (Bazan-Arias et al.,
2002), as also discussed in Section 6.6.3.3. The stabilized fnes were then used as the foundation for a
highway embankment that crossed the upstream end of the impoundment.
For an existing impoundment, an approach that might be taken is to show that the setled fnes have
consolidated and gained sufcient strength that they will not fow. Whether setled fnes will fow
depends on factors such as their degree of consolidation and cohesive strength, pore pressures, the
potential for excess pore-water pressures to be induced, and the size of opening associated with a
potential breakthrough feature. One qualitative measure of fowability is the moisture content of the
fnes as compared to their liquid limit (LL) and plastic limit (PL). If the moisture content of fnes near
the botom of an impoundment is below the LL and close to the PL, they are less likely to fow and
progress to a breakthrough unless the openings in the overburden above or adjacent to the mine are
sufciently large to afect more fowable fnes at shallower depths in the impoundment.
A change in conditions in an impoundment, such as infow of runof from a large storm, could result
in the liquefaction of some portions of the fnes leading to an unplanned release of slurry and water
through underlying, connected mine workings. Additionally, if subsidence occurs beneath saturated,
hydraulically placed fnes, the sudden increase in shear stress in the fnes may increase pore-water
pressure, triggering static liquefaction and causing the fnes to fow (Davies et al., 2002). While there
may be mitigating conditions in such a situation, measures to drain the fnes can be helpful in reduc-
ing or eliminating the consequences of a breakthrough because partially saturated fne refuse is likely
to densify under load and thus would be more resistant to liquefaction.
8.5.1.9 Overexcavation and Induced Subsidence
Overexcavation of the overburden and mine workings around impoundment perimeters and beneath
embankment structural zones may be feasible, particularly in areas of thin overburden such as along
outcrops. When applied, this approach includes removal of sufcient overburden to mitigate concern
for sinkhole subsidence over the remaining workings by creating a highwall. Barriers must then be
designed for any exposed entries found at the base of the highwall. Additionally, highwall stability
will need to be addressed as part of the overexcavation plan.
Inducing subsidence through controlled blasting may also be a feasible approach. Guidelines for
evaluating the technical and cost feasibility of using controlled blasting to induce subsidence are
provided by Workman and Satchwell (1987). In areas where old, unmapped, or poorly mapped
mine workings make evaluation of mine breakthrough potential difcult, collapsing the workings
by controlled blasting may be an option. Controlled blasting can prevent future mine subsidence
< PREVIOUS VIEW
Chapter 8
8-42 MAY 2009
from inducing cracks or sinkholes or from compromising planned mine seepage barriers. However,
not all rock overburden strata are suitable for this approach; implementation requires: (1) a geologic
investigation for determining the character of the strata, (2) a test area for initial demonstration, and
(3) monitoring programs for evaluating performance and addressing safety issues associated with
blasting and impacts to adjacent areas. Obviously, this approach should be evaluated with consider-
able diligence and should only be implemented under the guidance of persons with expertise and
experience in dam safety, explosives, and the blasting characteristics of local rock strata.
8.5.1.10 Monitoring Provisions
Whenever there is potential for subsidence to afect a dam or for a breakthrough, critical parameters
should be identifed and a monitoring program should be implemented to determine whether the dam
and overburden is performing/behaving as anticipated. Typical monitoring might include instrumenta-
tion for measuring reservoir and piezometric levels, discharge rates from mine workings and drains,
seepage quantity and quality, water levels in the mine, ground movement, and weather conditions.
Use of such instrumentation is discussed in Sections 13.2.1 and 13.2.2. Acceptable ranges and warning
or action levels should be established for all monitored values. Monitoring programs should include
requirements for ploting and evaluating data in a timely manner and review by an engineer familiar
with the design of the facility. Such practice permits detection of trends and correlations to other data.
8.5.2 Mine Entry Bulkhead Seal Design
Bulkheads, also known as hydraulic seals, may be installed across underground mine entries or in
mine openings at coal seam outcrops. Normally, bulkheads are installed across underground mine
entries to control groundwater in abandoned workings, to prevent rapid inundation of active mine
areas in the event of a breakthrough, or to serve as a retention dam in an underground disposal system
for fne coal waste. At coal seam outcrops, bulkheads have typically been used to prevent access and
to control acid mine drainage. Bulkheads have also been installed at outcrop openings within the
footprint of a refuse embankment or within the refuse impoundment to prevent stored water or
slurry from fowing into active or inactive underground mine workings. Also, when coal barriers are
found to be inadequate, bulkheads may be added on the mine side of the barrier to enhance safety.
Design considerations include: site preparation, seal type selection, design load assessment, struc-
tural resistance of the bulkhead and surrounding strata, and safety monitoring. Additional guidance
is presented in Guidelines for Permiting, Construction and Monitoring of Retention Bulkheads in Under-
ground Mines by Harteis, et al. (2008)
8.5.2.1 Site Preparation
As problems with bulkheads are ofen associated with seepage along the bulkhead and rock interface
or through the surrounding strata, the nature of the rock strata at a bulkhead location is important. To
the extent practical, bulkheads should be located in the most competent and least fractured area avail-
able (normally away from pillar corners), so that problems are minimized. The information gathered
as part of the evaluation of a potential bulkhead location should include the type and strength of the
rock in the roof and foor and the strength of the coal at the location under consideration. Data related
to roof falls, pillar punching, foor heave, or other unusual conditions should be reviewed and evalu-
ated. Coal pillars adjacent to the bulkheads should have a high factor of safety against failure, taking
into account all loading factors including transfer stress due to overlying or underlying mining and
stresses imposed by the dam and impoundment. Another consideration is that once a bulkhead is
constructed, active mining in the vicinity, whether in the same seam or in other seams, can afect the
surrounding strata and the load on the bulkhead itself.
As part of the evaluation, fractures or joints in the surrounding rock should be located and char-
acterized. Geophysical techniques can be employed to supplement visual observation for locating
rock discontinuities. The potential for subsidence and sinkhole formation over entries should also be
< PREVIOUS VIEW
8-43
Site Mining and Foundation Issues
MAY 2009
assessed. If joints, fractures, or subsidence cracks are present, then chemical or cement-based grouting
measures may be necessary in order to minimize seepage pathways that could lead to piping prob-
lems or deterioration of the rock strata. To maintain the integrity of a bulkhead, it may be necessary
to construct a grout curtain around the perimeter of the planned location. In addition to grouting,
loose, cracked, and weak foor, roof, and rib material should be removed. Regardless of the bulkhead
type, if the surrounding material consists of weathered or sof rock, then piping (internal erosion)
and hydro-fracturing of this material should be considered as potential failure mechanisms.
In-mine bulkheads have failed because of sofening of the foor material. These failures occur with a
rooted claystone referred to as freclay. When subjected to water, this material breaks down from
a rock-like material to a sof soil. The presence of water causes the electrostatic charge on the com-
ponent clay particles to break down. This process results in dispersion of the clay particles, loss of
cohesive strength, and potential erosion from seeping water. For this reason, claystone and freclay
foors should be removed from beneath the planned footprint of a bulkhead by trenching to a depth
sufcient to expose hard, competent rock. Trenching should minimize the potential for a foor failure
directly beneath a bulkhead. Floor trenches should be backflled with a seepage-resistant material
such as a cement-based or polyurethane-foam product. The surface of freclay running under the
pillars should be sealed and the bulkhead should be constructed as soon as possible afer the foor is
exposed in order to limit the tendency for the freclay to swell or absorb moisture from the air.
Shales in mine roofs, foors, and partings tend to be pyritic and thus prone to swelling. Bulkheads in
shale should be designed to accommodate additional load due to swelling and should also accom-
modate deterioration of the shale over time due to weathering and moisture efects. High-clay-con-
tent shales can break down in the same manner as freclays, and similar measures to those discussed
above can be used to mitigate potential problems.
The integrity of a bulkhead system can be afected by hydraulic fracturing. This type of fracturing can
occur when pressure from seeping water is sufciently great to cause cracks in rock strata to widen
and grow. When coal is mined to create an entry, the associated stress relief can result in stress frac-
tures or the opening of natural discontinuities in the mine roof, ribs, and foor. Specifcally, the foor
may heave upward in some mines. This opening of the rock strata can lead to additional hydraulic
fracturing if seepage pressures elevate. Strata grouting and notching of bulkheads into competent
material are methods for mitigating potential bulkhead damage due to hydraulic fracturing.
Prior to construction, supplemental roof support should be provided near the interior (inby) and exte-
rior (outby) sides of the bulkhead location. Roof support alternatives include roof jacks and cribbing
on both sides of the bulkhead. Unlike typical mine supports, these supplemental roof supports are
considered permanent installations. Therefore, corrosion protection should be incorporated into all
steel supports, and concrete or other durable cribbing material should be used. Sof foor materials
should be removed before installing supplemental roof supports. Notching of the bulkhead into the
surrounding rock strata is also recommended. A notch will create a longer fow path for seepage and
will increase the resistance to a contact failure between the bulkhead and the rock strata. In most cases,
notching will place the bulkhead in contact with more competent material as loose material is removed
to construct the notch. When roof notching is not feasible, a steel structural-angle member (with corro-
sion protection) can be bolted to the roof on the outby side of the bulkhead to provide lateral support.
Angles with dimensions of 6 inches by 6 inches by inch have been used for this purpose, but the
angle should be sized based on the design loads. Seepage resistance can also be increased by treating
the surrounding rock strata with a low-hydraulic-conductivity coating such as shotcrete.
8.5.2.2 Bulkhead Types
Several types of materials and physical arrangements have been used for bulkhead construction.
Figures 8.13 and 8.14 illustrate entry and drif opening bulkhead concepts, including bulkheads
< PREVIOUS VIEW
Chapter 8
8-44 MAY 2009
installed at mine entrances to prevent water from entering. Bulkhead arrangements typically are
either thick plugs with a straight or tapered length or thin walls with a straight or arched shape. Each
type of construction has advantages and disadvantages. In general, the thicker the bulkhead, the
more resistance there will be to seepage and piping around the perimeter.
Kirkwood and Wu (1995) present technical considerations for mine seals to withstand hydraulic heads
in underground mines. Solid concrete block walls can be designed to withstand these loads and have
been tested under hydraulic pressure of 40 psi (92 feet of head), but it is recommended that the maxi-
mum hydraulic pressure be limited to 2.6 psi (6 feet of head) because of long-term efects of strata
movement and deterioration surrounding the seal (Chekan, 1985). Reinforced-concrete walls can be
designed for signifcant hydraulic loads, provided it is recognized that the maximum safe head is
controlled by the capacity of the surrounding rock. While notching a seal will help increase the total
shearing resistance provided by the surrounding rock and increase the seepage path, the bearing
FIGURE 8.13 BASIC BULKHEAD DESIGN CONCEPTS
8.13a WALL KEYED INTO RIBS, ROOF AND FLOOR
FIGURE 8.13 BASIC BULKHEAD DESIGN CONCEPTS
8.13b TAPERED PLUG
8.13c PARALLEL PLUG
= DIRECTION OF HYDROSTATIC PRESSURE
(HARTEIS AND DOLINAR, 2006)
FIGURE 8.13 BASIC BULKHEAD DESIGN CONCEPTS
< PREVIOUS VIEW
8-45
Site Mining and Foundation Issues
MAY 2009
capacity of the foor and adjacent ribs must be addressed. The difculty of achieving adequate contact
with the roof and dealing with the limited access for working the concrete around the reinforcing can
best be handled by using experienced work crews. When constructed as a drif opening bulkhead, a
reinforced-concrete wall (Figure 8.14a) can be used to cap of an entry, but contact with the mine roof
may require notching, dowels, or other measures to improve the seal and structural integrity.
Thick concrete-plug seals overcome these limitations through development of thicker barriers with
signifcantly more contact with the mine ribs, foor and roof. Typically concrete-block walls serve
as forms at either ends of the seal, and concrete or cement-foam materials are used to construct the
plug, as shown in Figure 8.15. Grouted-rock seals are similarly constructed, but include rock (gravel
to large hand-placed rock) that is grouted in place, as shown in Figure 8.14b. Grout materials vary
from cement to rigid polyurethane foam. Grouted-rock seals are generally as thick as or thicker than
concrete-plug seals.
When cement-based products are used to construct thick plugs, consideration should be given to the
possibility of heat buildup within the fresh concrete mass during curing (Figure 8.15). This buildup
8.14a REINFORCED CONCRETE "CAP" BULKHEAD
8.14b GROUTED BULKHEAD
FIGURE 8.14 DRIFT OPENING BULKHEAD DESIGN CONCEPTS
(HARTEIS AND DOLINAR, 2006)
MINE OPENING
REINFORCED-
CONCRETE
WALL
MINE FLOOR
MINE ROOF
MINE OPENING
BLOCK RETAINING
WALL (BOTH SIDES)
MINE ROOF
MINE FLOOR
GROUTED ROCK
(e.g., POLYURETHANE-
LIMESTONE CORE)
FIGURE 8.14 DRIFT OPENING BULKHEAD DESIGN CONCEPTS
FIGURE 8.14 DRIFT OPENING BULKHEAD DESIGN CONCEPTS
< PREVIOUS VIEW
Chapter 8
8-46 MAY 2009
of temperature is termed the heat of hydration and, if not accounted for, can lead to internal crack-
ing within the plug and to property changes in the surrounding rock strata during the curing period.
Measures to minimize heat buildup include using low-heat cements, replacing a portion of the cement
with certain pozzolans, using larger aggregate, and cooling the mix water. Guidance can be found in
ACI 207.1, Guide to Mass Concrete (ACI, 2005a). Polyurethane foam will also generate heat as it cures.
The potential for this heating to cause a fre due to contact with combustible material must be taken
into account.
Also, polyurethane foam is highly sensitive to moisture that may be present at a bulkhead construc-
tion site. Moisture can cause the foam to expand more than intended, resulting in a lower density and
lower strength material. The product manufacturer should be consulted regarding measures to prevent
moisture from adversely afecting the foam.
Regardless of type of material selected for bulkhead construction, resistance to deterioration from acid
mine water should be evaluated. For example, sulfates in groundwater, coal and the surrounding rock
strata can cause deterioration or spalling of certain types of cements that are not inherently resistant to
sulfate atack. Because of the high-sulfate environment in a coal mine, any cement used in fll should
be sulfate resistant. Type V cement is preferable, but in some mine environments, Type II cement will
sufce. Depending on cement availability, Type II or Type I cement in combination with pozzolans (e.g.,
fy ash, slag cement) may be used in order to provide adequate sulfate resistance.
When formwork is used for bulkhead construction, it should be adequately braced for structural sup-
port, and vents should be installed at the top of the formwork to release entrapped air and to prevent
the formation of voids. Because concrete can shrink during curing, remaining voids between the sur-
rounding rock strata and the bulkhead should be flled by contact grouting.
In selecting the type of bulkhead, consideration should be given to the potential for roof convergence.
Some types of materials (e.g., low-density cementitious foams, lightweight concretes, and polyure-
thane foams) can accommodate this compression by deforming without cracking.
FIGURE 8.15 BULKHEAD CONFIGURATION SHOWING CONCRETE-BLOCK
RETAINING WALLS AND CONCRETE CENTER
FIGURE 8.15 BULKHEAD CONFIGURATION SHOWING CONCRETE-
BLOCK RETAINING WALLS AND CONCRETE CENTER
(ADAPTED FROM CHEKAN, 1985)
ENTRY
CONCRETE BLOCK
RETAINING WALL
1
0
T
O
3
5
F
T
RIB
RIB
CONCRETE BLOCK
RETAINING WALL
C
O
N
C
R
E
T
E
C
O
N
C
R
E
T
E
FIGURE 8.15 BULKHEAD CONFIGURATION SHOWING CONCRETE-BLOCK
RETAINING WALLS AND CONCRETE CENTER
< PREVIOUS VIEW
8-47
Site Mining and Foundation Issues
MAY 2009
8.5.2.3 Design Loads
Bulkheads should be designed to resist the estimated maximum hydraulic and geologic loads. The
hydraulic pressure is afected by the projected head of water or slurry behind the bulkhead. This level
may be infuenced by such factors as:
Other mine drif openings
Shaf or borehole openings
Flooded overlying mines
Flooded adjacent mines with inadequate barrier pillars
Partial height seals located up-dip that act as a weir spillway and divert the
water elsewhere in the mine
Changes in the mine foor slope that divert water into other parts of the mine
Maximum seasonal groundwater levels or efects of slurry injection
Maximum design storm water level in an overlying impoundment
The potential water level in overlying fooded mines is a concern if cracks in the interburden pro-
vide a hydraulic pathway between mines. There have also been instances where fooded mines located
beneath the subject seam dip such that the food water heads in the seam below are greater than the
elevation of the bulkhead in the overlying mine opening or entry. In this case, if cracks are present in the
interburden strata, the water level in the underlying mine could control the bulkhead design head. The
inlet elevation of drainage pipes extending through up-dip ventilation seals is normally not considered
the limiting design water pool level because these pipes can clog. Because of these considerations, an
accurate, up-to-date contoured mine map is essential for predicting the design pool level.
In addition to static hydraulic loads, seismic loads may need to be considered if a site is located
in a seismically active area. In this event, bulkhead design should account for both the inertial
and the hydrodynamic forces that could result from an earthquake. The hydrodynamic forces
would result from an increase in static pressure caused by acceleration of the water mass behind
the bulkhead.
The roof in a mine entry can exert a compressive load on the top of a bulkhead seal. This compressive
force should be determined based on experience with convergence or by estimating the stress that
could occur in the zone of material forming the pressure arch over the entry.
A bulkhead should be designed to have the structural capacity to resist the forces acting on it with
a factor of safety consistent with the degree of uncertainty associated with the type and magnitude
of load and the consequences of failure. A bulkhead must be able to resist the shear and bending
stresses caused by the pressures acting on its face. The bending stresses between the roof and foor
and between the adjacent ribs can be calculated based on the width and length dimensions and the
type of edge restraint (Timoshenko and Woinowsky-Krieger, 1959; Young and Budynas, 2001). For
bulkheads that are thick relative to their span, Young and Budynas (2001) provide guidance on stress
multipliers. In addition to resisting lateral loads, the bulkhead should have the capacity to resist
vertical bearing loads caused by the mine roof convergence and stress transfer. With relatively thin
bulkheads, grouting may be necessary in order to prevent excessive seepage from adversely afecting
the anchorage of the bulkhead or the surrounding strata.
8.5.2.4 Design of Solid Concrete Block Bulkheads
The analysis of a solid concrete block bulkhead requires estimation of the shear strength around the
perimeter of the wall and evaluation of the structural capacity of the wall itself. The structural capac-
< PREVIOUS VIEW
Chapter 8
8-48 MAY 2009
ity of the wall is dependent on the bending stresses near the center of the entry. Because a concrete
block wall has no steel reinforcement, it has relatively poor fexural strength, and, except in the case
of very narrow entry widths, the lack of fexural strength limits the maximum hydraulic head more
than the strength of the surrounding strata. Kendorski et al. (1979) provide guidance for block wall
design and the estimation of fexural stress. The structural capacity of the concrete block wall will
also depend on the strength of the block, the strength of the mortar, joint thickness, and the quality of
construction. Masonry design and construction guidance such as ACI 530, Building Code Requirements
for Masonry Structures (ACI, 2008), should be followed.
8.5.2.5 Design of Concrete Bulkheads
For thin bulkheads installed with an entry width-to-height ratio greater than 2, a reinforced-concrete
bulkhead can be designed as a one-way slab spanning between the roof and the foor if adequate
edge connections are provided at the mine roof and foor. To account for temperature and shrinkage
stresses, reinforcement is required in the rib-to-rib direction if the bulkhead is designed for one-way
behavior in the roof-to-foor direction. For an entry width-to-height ratio less than 2, the bulkhead can
be designed for two-way behavior provided there is adequate anchorage on all four sides. However,
reinforcement should be sized to separately carry the full bending moment if there is concern that
either the roof or foor or ribs may not provide adequate resistance. Regardless of the width-to-height
ratio, diagonal reinforcement steel should be placed in the bulkhead corners to control cracking from
torsional moments. Regardless of the load path direction assumed, if the mine roof, foor, and ribs are
notched into the surrounding rock strata or if the steel bar reinforcing mats near the inby and outby
faces are doweled into the surrounding strata, then it is possible to develop negative moment bend-
ing stresses at the edges of the bulkhead slab. Negative steel (i.e., the reinforcement near the inby
wet side face) should be sized to resist negative moment bending stresses.
Reinforced-concrete structures should be designed in accordance with the most recent version of
ACI 318, Building Code Requirements for Structural Concrete (ACI, 2005b) and ACI 350 (where appli-
cable). The codes are based on ultimate strength design in which factors are applied to the loads and
strength to account for the uncertainty of these parameters in structural design. Using this approach,
a safe design is achieved when the factored design strength of a structure component exceeds the fac-
tored loads. For fuid pressure loading, the load factor is 1.4 when the maximum height of the water
or slurry is controllable or conservatively estimated. Thick concrete bulkheads with a thickness-to-
height ratio greater than 1 should be designed in accordance with Section 8.5.2.5.2.
8.5.2.5.1 Design of Reinforced-Concrete Bulkheads for Flexure
Where the span-to-thickness ratio of the bulkhead is greater than 2, the design fexural strength of a
reinforced-concrete member (ACI, 2005b) is determined using the relationship:
M
d
= A
s
f
y [
d
a
]
(8-6)
2
where:
a =
A
s
f
y
(8-7)
0.85 f
c
b
M
d
= fexural design strength (lb-in)
= strength reduction factor = 0.90 (dimensionless)
A
s
= area of tension reinforcement (in
2
/f)
f
y
= yield strength of reinforcing steel (psi)
d = distance from extreme compression fber to centroid of tension reinforcement (in)
< PREVIOUS VIEW
8-49
Site Mining and Foundation Issues
MAY 2009
a = depth of rectangular stress block at failure (in)
f
c
= specifed minimum compressive strength of concrete (psi)
b = width of compression face of member, normally taken as 12 in for slabs (in)
For thick bulkheads with a span-to-thickness ratio less than or equal to two, the fexural design
strength from Equation 8-6 should be modifed to refect the behavior of a thick fexural member.
Because thicker members have low span to thickness ratios, a linear stress distribution is no longer
valid. According to Park and Paulay (1975), for simply supported members with span-to-depth (thick-
ness) ratios less than or equal to two, the internal lever arm can be calculated as:
z = 0.2 ( + 2h) 1 / h 2 (8-8)
z = 0.6 / h 1 (8-9)
where:
= centerline-to-centerline span distance of two bearing points or 1.15 times
the clear span, whichever is smaller (in)
h = thickness of bulkhead (in)
z = internal lever arm (in)
Applying the revised lever arm value to the standard fexural equation (Eq. 8-6), the capacity can be
estimated using:
M
d
= A
s
f
y
z (8-10)
The value of z should not exceed d a/2. If the end supports are assumed to be fxed, rather than
simply supported, the value of z should be further adjusted (Park and Paulay, 1975). Nilson et al.
(2002) recommend that tension steel in a deep fexural member be distributed over the botom third
of the member depth.
8.5.2.5.2 Design of Unreinforced-Concrete Plug Bulkheads for Shear
Unreinforced-concrete bulkheads typically have a thickness-to-height ratio greater than 1. The shear
resistance of these bulkheads may be governed by the:
Shear strength of the seal material
Shear strength of the surrounding strata
Contact interface shear strength between the seal and the strata
In cases where there is no notching of the bulkhead into the surrounding strata, the interface resis-
tance will be governed by adhesion or friction. The South African plug formulas, which are based on
the shear strength and bearing capacity of the bulkhead material and surrounding strata, are ofen
used to evaluate the required length of thick bulkheads (Garret and Campbell Pit, 1961):
=
p a b
(8-11)
2( a + b) f
s
=
p a b
(8-12)
( a + b) f
c
< PREVIOUS VIEW
Chapter 8
8-50 MAY 2009
where:
= length of bulkhead (f)
p = hydraulic pressure on bulkhead (psi)
a = width of entry (f)
b = height of entry (f)
f
s
= minimum allowable shear strength of strata or plug material, whichever is less (psi)
f
c
= minimum allowable compressive strength of strata rock or plug material,
whichever is less (psi)
The values of f
s
and f
c
obtained from sampling may not conservatively refect the strength of the de-
stressed edges of the coal pillars. The designer should select the bulkhead length based on the larger
of the values obtained from Equations 8-11 and 8-12. These equations are most applicable to high-
head situations where the bulkhead acts as a massive plug. If the bulkhead span-to-thickness ratio
is greater than one, then fexural reinforcement should be provided, as discussed in Section 8.5.2.5.1.
Selecting shear strengths for coal is problematic, as such values are typically not determined and
research in this area is limited. Coal is generally crushed or yielded at ribs and extending some dis-
tance into pillars. Some designers use the cohesive strength of the coal if the ribs are scaled and the
bulkhead is constructed before substantial additional yielding takes place. If this is not possible, the
residual cohesive strength of coal is taken as its shear strength. A common estimate for the cohesive
strength for coal is 145 psi, which, depending on the practitioner asked, may have resulted from
the calculation of 16 percent of the typical compressive strength (0.16 x 900 psi = 144 psi) or from
an assumed minimum value of 1 MPa (145 psi). Sixteen percent is a rule-of-thumb sometimes used
to estimate the cohesive strength of rock from the compressive strength. It can be difcult to obtain
representative samples of coal for testing, but a method of estimating the shear strength of coal is to
conduct direct shear tests with coal samples oriented to model in-mine conditions.
Methods for increasing bulkhead resistance include notching the bulkheads into the surround-
ing rock strata, tapering the plug, and/or installing corrosion resistant, epoxy-coated dowel rods
into the surrounding rock strata and allowing the rods to protrude into the bulkhead. The dowel
rods should have an embedded length into the surrounding strata and the bulkhead sufcient to
develop the strength of the dowel rod without a bond failure. The minimum required development
length in concrete should be determined in accordance with the most recent version of ACI 318,
Section 12.2.
Thick concrete bulkheads with a thickness-to-height ratio greater than 1 should be designed in accor-
dance with the most recent version of ACI 207.1. Reinforcement should be provided to control tem-
perature and shrinkage stresses at the surface and should be designed in accordance with the most
recent version of ACI 318. Reinforcement is not necessary when the bulkhead thickness is greater
than 6 feet, provided that steps are taken to control the efects of temperature and shrinkage. Heat of
hydration can be controlled by using low-strength concrete, low-heat cement, or replacing a portion
of the cement with pozzolans. It can also be controlled by reducing the placement temperature using
cooled mix water or aggregates. Accelerating admixtures should not be used.
8.5.2.5.3 Design of Reinforced-Concrete Bulkheads for Shear
While Equations 8-11 and 8-12 can be used to determine the required length of thick bulkhead plugs
in order to prevent failure in shear, Equation 8-13 can be used to calculate the design shear strength
of concrete for thinner, reinforced-concrete bulkheads:
< PREVIOUS VIEW
8-51
Site Mining and Foundation Issues
REV. AUG. 2010
V
c
= 2
f
c
b
w
d (8-13)
where:
V
c
= shear strength of the concrete bulkhead per unit width (lbs)
= shear strength reduction factor = 0.75 (dimensionless)
b
w
= unit width of bulkhead (12 in)
d = distance from inby side of bulkhead to centroid of tensile reinforcement (in)
If Equation 8-13 indicates inadequate concrete shear strength, a more rigorous formulation should be
obtained from the most recent version of ACI 318, Section 11.3. In addition, the contribution of steel
shear reinforcement, as discussed in ACI 318, Section 11.5, can be added to the value obtained for the
concrete to obtain a combined shear strength for the bulkhead.
For thick, reinforced-concrete bulkheads where the ratio of the clear span distance
n
to the depth d
from the inby side of the bulkhead to the centroid of the tensile steel reinforcement is less than four,
the most recent version of ACI 318, Section 11.8 should be applied. It should be noted that the span-
to-depth ratios are diferent for shear design than they are for fexural design.
If lightweight concrete with a density of 100 to 110 pcf is used, values of V
c
obtained using
f
c
in the
Equation 8-13 should be multiplied by 0.75 for all-lightweight concrete and by 0.85 for sand-light-
weight concrete, or should be in accordance with the most recent version of ACI 318, Section 11.2.
8.5.2.6 Monitoring of Bulkheads
The water pressure behind a bulkhead should be monitored so that it can be compared to the design
pressure. Warning levels that would warrant drawing down the mine pool or initiating an emergency
action plan should be established. To lower the water pressure, a corrosion-resistant pipe should be
installed through the bulkhead with a U-trap and a pressure relief valve on the downstream end
and with provisions to prevent clogging (such as a riser and trash rack) on the upstream end. Pipes
extending through a bulkhead should be equipped with external collars to cut of seepage and mini-
mize the potential for a pipe blowout.
A possible safety measure for regulating the maximum hydraulic pressure is to install an additional
pipe through the bulkhead with a corrosion-resistant rupture disk atached to the downstream end.
If the hydraulic pressure on the bulkhead reaches the rupture strength of the disk, it will fail and
thus limit the load on the bulkhead. The outlet end of the pipe should project downward to prevent
injuries to anyone near the pipe if the disk should suddenly rupture. If water could collect and food
workings, inhibiting escape for miners, an evaluation of the consequences of an unexpected water
release should be conducted.
Seepage can occur through a bulkhead or the surrounding strata, but the presence of seepage is not
necessarily a sign of distress. Since unexpected increases in seepage could indicate deterioration of
the bulkhead or adjacent strata, seepage through the bulkhead and the surrounding strata along with
the corresponding head behind the bulkhead should be monitored and evaluated. To this end, seep-
age on the active side of the bulkhead is ofen channelized and monitored using a weir to facilitate
measurement of changes in quantity.
A fnal safety measure is to establish an inspection schedule for long-term monitoring of the bulk-
head and to have a contingency evacuation plan in place for situations where problems are encoun-
< PREVIOUS VIEW
Chapter 8
8-52 MAY 2009
tered with the integrity of the bulkhead or surrounding strata. To determine the impact of a bulkhead
failure on active in-mine escapeways, the potential inundation area should be identifed.
8.6 FOUNDATION-RELATED CONSTRUCTION AND OPERATIONS MONITORING
Observations of foundation conditions during construction and operation of impounding facilities
are critical. During construction, foundation conditions including mine-related features are exposed,
allowing the facility operator and engineer to assess conditions for consistency with expectations and
to evaluate specifc foundation preparation design measures. The operator and engineer responsible
for construction certifcation should involve the designer during critical tasks such as construction/
implementation of bulkheads and seals, cutof trenches, and grouting programs. During construc-
tion, survey control to confrm locations of key features and the dimensions of associated structures
is essential. Documentation of conditions with photographs and as-built drawings and reports of
construction activity is important for certifcation of the work and for subsequent evaluation, if nec-
essary. Additional discussion of construction monitoring is presented in Chapter 12. A discussion of
construction and operations procedures and monitoring for bulkheads is provided in MSHA (2003),
and key guidance from the MSHA document is presented in this section.
Where there are mine workings near an impoundment, the manner in which an impoundment is
operated can afect the breakthrough potential. For example, measures should be taken to design
and operate the impoundment in a manner that minimizes the presence of free water. Thus, decant
raising should be staged so that the water level rises incrementally. Pumping can also be employed
to minimize the volume of water in the impoundment. Mine personnel who work on or around the
impoundment should be cognizant of key components of the operation plan and particularly of any
unusual operational requirements.
A site-specifc monitoring plan oriented toward breakthrough prevention and assessment of the poten-
tial for subsidence to afect the dam should be developed. Monitoring involves collection of information
from both visual inspection of the impoundment and from instrumentation. Coal company personnel
who inspect the impoundment, or who routinely work on or around the impoundment, should be trained
to observe potential signs of trouble that could be related to subsidence efects or a breakthrough. These
personnel should be aware of where underground mining has occurred and where to look for cracks or
other evidence of subsidence. Signs that they should look for include: (1) unusual sudden drops in the
pool level, (2) the presence of a whirlpool or bubbles in the pool, (3) cracking or sudden displacement
in embankment surfaces, (4) unusual readings in piezometers, (5) changes in seepage conditions, and
(6) changes in the quantity of fow and the amount of sedimentation in discharges from mine openings,
backflled mine openings, or outcrop areas. Instruments such as staf gages, weirs, survey monuments
and inclinometers are useful for monitoring these conditions and detecting changes.
In determining what instrumentation should be installed and monitored, the designer should iden-
tify the parameters that will indicate how the site is performing with respect to potential failure
modes. This will facilitate taking action if the facility does not perform as expected. The following
items should be considered in developing a monitoring plan:
Seepage Seepage from the impoundment should be monitored, including seep-
age through the embankment, through any internal drainage systems, and through
underground mines that receive seepage from the impoundment. Weirs or other
devices should be installed so that fow rates can be easily and consistently mea-
sured. Changes in water quantity and quality in seeps and discharges (including
mine/pump discharges) that are hydraulically connected to the impoundment
should be monitored. Changes in seepage quantity, particularly when not correlated
to rainfall or pool water levels, may indicate deteriorating or adverse conditions war-
ranting additional investigation.
< PREVIOUS VIEW
8-53
Site Mining and Foundation Issues
MAY 2009
Water levels The pool levels in the impoundment and in underground mines
should be monitored, and unexplained changes should be investigated. If there are
bulkheads in the mine, the water pressure against them should be monitored. Where
conditions with respect to breakthrough potential are uncertain, instrumentation
can be installed to provide an alarm in the event of a sudden drop in the impound-
ment water level. The alarm would alert mine personnel to check on the situation
and would facilitate early warning and emergency response in the event of a break-
through failure.
Piezometric levels Saturation levels and water pressures in the refuse embankment,
as well as in any other earthen barriers, should be monitored to determine whether
hydrostatic pressures are within design limits and whether any changes or trends
are reasonable. In situations where it is critical to be able to measure rapid or sudden
changes in pore water pressure, a closed system such as a vibrating-wire piezometer
should be used.
Rainfall data It is good practice to install a rain gauge in the vicinity of an
impoundment, but it is especially important in situations where there is break-
through potential and where discharges from a mine can be traced to seepage from
the impoundment. In such cases, rainfall data should be routinely collected so that it
can be determined whether changes in the water fow or water level data correlate to
rainfall or may be occurring for other reasons.
Ground movement When there is potential for subsidence in the vicinity of an
impound ment, the ground surface should be monitored for movement. Both hori-
zontal and vertical movements should be measured.
The types and suitability of instrumentation for accomplishing these monitoring objectives are dis-
cussed in Chapter 13. The timely review and interpretation of instrumentation data by someone
knowledgeable in the design and performance of impoundments is critical to an efective monitoring
program.
The type and frequency of monitoring required depend on impoundment conditions. Typically, mon-
itoring is performed during regular weekly inspections. Where conditions warrant, more frequent or
even continuous automated monitoring may be needed. Monitoring plans should include provisions
for ploting and evaluating the observed data in a timely manner. When a potentially hazardous
condition develops, more frequent monitoring is required. 30 CFR 77.216-3(b)(4) requires, in part,
that when a potentially hazardous condition develops, the mine operator shall immediately direct a
qualifed person to monitor all instruments and examine the structure at least once every eight hours
or more ofen, if requested by an authorized representative of MSHA.
8.7 MINE BACKFILLING DESIGN
Backflling of mine workings is performed in order to provide localized support for pillars and the
mine roof and to reduce the volume of open space that could potentially be flled with collapsed
material, thus tending to minimize the deformation of the surrounding rock mass. Support for pillars
and volume reduction of open space can be achieved by a variety of backfll types. While deforma-
tion and bulking of the roof strata can provide support for mine overburden, backflling with grout
to improve contact with the roof may be a desirable option. The most common types of fll material
are waste rock, mill tailings, quarried rock, sand and gravel, and fy and botom ash. Additives such
as portland cement, lime, fy ash, and pastes can also be mixed with the fll to alter the characteristics
and to improve the efectiveness of the backfll. Placement alternatives include stowing by hand,
gravity, mechanical, pneumatic and hydraulic methods. Hydraulic placement is generally efective
for the varying conditions encountered in mine workings.
< PREVIOUS VIEW
Chapter 8
8-54 MAY 2009
Backflling of mines has signifcantly reduced surface damage from subsidence by lending lateral
support to pillars and by limiting the volume of voids (National Academy of Sciences, 1975). The
most important consideration in mine backfll design is the planned backfll material. The mechanics
of uncemented fll can be analyzed using soil mechanics principles (Coates, 1981). The strength of
most uncemented hydraulic flls is related to frictional resistance to sliding between particles and can
be afected by pore water pressure and erosion, as well as dynamic loads such as blasts or sudden
fuctuations in the water table. Additionally, the compressibility of the backfll material is related to
its ability to provide support for the pillars and roof.
Cemented backfll has cohesion resistance gained through the addition of cement or pozzolanic
admixtures that render the backfll relatively incompressible. Backfll strength usually increases lin-
early with cement content, which can range up to 10 percent. The grain-size distribution of the fll
may also be important. Well-graded material generally has a greater strength than uniformly graded
material, although the fnes content (i.e., minus 200 mesh portion) can adversely afect the strength.
The inclusion of pozzolanic material additives such as fy ash can reduce the volume and associated
cost of portland cement in a mine backfll while signifcantly increasing strength and providing other
benefcial characteristics such as improved fuidity during placement. The strength and placeability
of candidate cement mixes should be evaluated through laboratory testing.
Typically, mine backfll provides substantial flling of mine voids, such that relatively litle bulking
from roof materials is needed to mitigate fracturing of the overburden and the advance of subsidence
in the overburden. In critical situations, the space between the backfll and the roof can be grouted. In
instances where mine backfll provides support for the coal pillars but not direct support for the over-
burden, the mechanical behavior of the cemented fll can be modeled as providing lateral restraint
according to the relationship (Cai, 1983):
h
= n
Ha
(8-14)
K
p
L
where:
h
= passive earth pressure (psf)
n = correction coefcient (dimensionless)
1
=
1
+
h
K
pl
(8-15)
where:
1
= strength of pillar supported by fll (psf)
1
= strength of pillar (psf)
h
= passive earth pressure provided by fll (psf)
< PREVIOUS VIEW
8-55
Site Mining and Foundation Issues
MAY 2009
The supported pillars can then be analyzed for stability based on the procedures discussed in Sec-
tion 8.4.3. For instances of thin overburden, Mitchell (1983) and Wizniak and Mitchell (1987) present
analytical procedures for estimating surface subsidence deformation for backfll placed to the mine
roof wherein the roof is modeled as a beam on an elastic foundation.
The design of mine backfll generally depends on the availability of fll materials and fy ash and their
associated costs.
8.8 SURFACE MINE SPOIL ISSUES
At some refuse disposal sites, surface mine spoil may be available in the foundation area, and some
of this spoil may be suitable for use in constructing structural portions of refuse embankments such
as the starter dam. Mine spoil is typically quite variable in terms of its composition of soil and rock
materials and also relative to maximum particle or fragment size and distribution, durability and
moisture content. Typical methods of spoil placement can also result in considerable segregation.
This variability makes it difcult to characterize the engineering behavior of mine spoil, and thus the
feld and laboratory procedures described in Sections 6.4 and 6.5 may need to be modifed to accom-
modate its special characteristics. This variability also can lead to difculties in estimating engineer-
ing properties such as compressibility, shear strength and hydraulic conductivity. Piping in mine
spoil materials is a concern and has been identifed as the cause of sinkholes and black-water releases
at slurry impoundment sites founded on mine spoil. When a coal refuse facility is to be constructed
over or using surface mine spoil, designers should recognize the variable character and composition
of the material and understand the potential impacts that it can have on the long-term performance
of a coal refuse embankment.
8.8.1 Surface Mine Spoil Characteristics
Surface mine spoils result from excavation and placement of overburden and interburden materials.
These operations typically range in size from several hundred to several thousand acres, and mine
life is typically fve to 30 years or more. Overburden removal is generally accomplished by continu-
ous bucket-wheel excavators, walking draglines, hydraulic excavators, stripping shovels, scrapers,
dozers, or cast blasting. Where unconsolidated materials such as glacial till or loess represent a sig-
nifcant portion of the overburden, bucket-wheel excavators are generally used.
Spoil placement is typically accomplished by dropping the spoil materials at the angle of repose to form
a ridge of piles parallel to the active pit or by placing the materials in lifs where the lif thickness and
degree of compaction depend on the equipment type, weight and number of passes.
Uncontrolled spoil placement is typically associated with contour and area mining, while placement
in lifs is generally associated with haul-back mining and head-of-hollow and valley-fll construction.
Before the late 1970s to early 1980s, spoil ridgetops were lef as deposited or were graded with a single
pass of a dozer that resulted in a ridge and trough topography. With the advent of state and federal
regulatory programs to return mining operations to near their original contours, spoil ridges have
been extensively graded with dozers leading to increased spoil handling and machinery trafc, greater
breakdown of the spoil, higher density and improved engineering behavior of the in-place spoil. A dis-
cussion of the infuence of mining method on the bulk density of spoil based on testing in the Eastern
Coal Province is presented in Phelps et al. (1981).
Because of the material handling that occurs during excavation, deposition and placement, mine spoil
materials experience signifcant changes in physical integrity. These changes are related to variations in
geologic characteristics, moisture, stress regimes, mining and reclamation methods, and other environ-
mental aspects of the materials. The physical deterioration of geologic materials caused by changes in
stress conditions or strength characteristics is referred to as slaking. The most distinctive aspect of the
slaking process is a relatively rapid decrease in the particle or fragment size of the material.
< PREVIOUS VIEW
Chapter 8
8-56 MAY 2009
The decrease in particle or fragment size caused by slaking can have a wide range of efects on the
behavior of the material in a spoil pile. These efects will depend on the gross characteristics of the
spoil pile (e.g., physical dimensions and confguration of the pile and the degree of compaction), the
durability of the pile materials, the proportion of slakable materials, and changes in the pile surfcial
or internal moisture regimes. Possible adverse efects of spoil slaking include: (1) decreases in mate-
rial strength that can reduce the stability of slopes and (2) increases in moisture content and decreases
in particle and fragment size that can increase setlement and surface erosion and afect hydrologic
regimes and vegetation. In a study of the environmental efects of slaking of surface mine spoils in
the eastern and central U.S., Andrews et al. (1980) observed that:
The rate and degree of particle breakdown is directly related to the material charac-
teristics (e.g., durability) and local environmental conditions (e.g., depth of burial).
The most active zone of slaking occurs within about three feet of the exposed mine
spoil surface.
The major observed efect of mine spoil slaking is a decrease in particle or fragment
size that results in changes to the hydrogeologic characteristics (e.g., rate of infltra-
tion, hydraulic conductivity, rate of groundwater fow) of spoil piles
The signifcance of slaking seems to be minimized by the mixing of slakable (e.g.,
mudstone and shale) and nonslakable (e.g., limestone) materials that usually occurs
during typical spoiling operations.
No gross environmental damages due to slaking were apparent at the sites visited.
Andrews et al. (1980) developed these observations through laboratory testing of bulk samples
obtained from test pits excavated in 2-, 5-, and 10-year-old spoils from four mine sites located in the
Appalachian Basin. Based on extensive qualitative and semi-quantitative data collected relative to the
behavior, causes, and efect of slaking materials, the study identifed three feld slaking modes:
1. Slaking to a constituent grain size that typically occurs in mudstones and occasion-
ally sandstones
2. Chip slaking to thin, platy fragments that generally occurs in shales, siltstones and
occasionally thinly-bedded sandstones
3. Slab or block slaking to large, approximately equidimensional fragments that gener-
ally occurs in sandstones and limestones.
The lithology, bedding and mineralogical characteristics of the spoil materials were found to have
a major efect on the mode, rate and degree of slaking. In general, mudstones, siltstones and shales
were found to be the most slake-prone lithologies. Slaking of sandstone and limestone was variable,
but generally minor. Bedding characteristics were the primary factor in the mode of slaking (i.e.,
rocks with thin bedding typically exhibited chip slaking whereas rock with a massive structure were
prone to block or slab slaking, or to slake to their constituent particle size). Spoils that slake to their
constituent particle size were found to be less stable, while spoils in which chip slaking or slab or
block slaking dominates generally were found not to be associated with stability issues.
The engineering behavior of mine spoil may be related to mining techniques. In general, overbur-
den removal by blasting, material transport and dumping by trucks, and placement and leveling by
dozers results in mechanical breakdown that leads to reduced slaking once the material is placed.
8.8.2 Design and Construction Considerations
The compressibility, shear strength and hydraulic conductivity of mine spoil can vary considerably
depending on the proportion of slakable and nonslakable materials and the methods used to place
and reclaim the spoil. Test methods that can be used to assess the slaking potential of mine spoil are
< PREVIOUS VIEW
8-57
Site Mining and Foundation Issues
MAY 2009
presented in Section 6.5.9.4. The application of these test methods as part of an overall management
control process are described in Andrews et al. (1980). Figure 8.16 presents a system of classifcation
and interpretation of spoil durability based on index and slake durability testing using the degrada-
tion index (DI), where DI = 1 I
d
and I
d
is the slake durability index. Based on conclusions and rec-
ommendations from the study, designers should consider the following measures in planning and
conducting site reconnaissance, site exploration and laboratory testing programs and in developing
designs and preparing construction documents:
Review of available surface mining records in order to document the mining and rec-
lamation practices used, the stratigraphy of overburden and interburden materials
excavated, and the time frame for mining operations.
Use of test pits to observe and document the general proportion of slakable and
nonslakable materials, the type and amount of slakable material breakdown, the
grain-size distribution and moisture content of bulk samples, and the existence of a
permanent or perched water level in the spoil mass.
Provisions to control seepage and prevent internal erosion through mine spoil zones
under refuse facility structures, including cutof trenches through spoil deposits or
impervious blankets with flter layers.
Measures to monitor the setlement and lateral displacement of foundation spoil
materials as the refuse embankment is raised. A discussion of instrumentation is pro-
vided in Chapter 13.
If mine spoil is to be used for embankment fll, the variability of the spoil should be determined so
that the embankment can be constructed in a manner that will allow the desired performance in
terms of strength and hydraulic conductivity to be achieved. The strength of mine spoil is typically
estimated by: (1) evaluating its angle of repose, (2) correlating with published values for rock fll if
the material is durable, or (3) by testing the fner portion of the material, which typically results in a
conservative estimation. Internal stability of non-uniform material can be a concern for mine spoil,
particularly if it is gap-graded. The hydraulic conductivity of mine spoil is typically estimated based
on grain size, and measures such as zoning may be required in order to address the variability of the
material. Designers should consider the following measures, if mine spoil is to be used to construct
some or all of a coal refuse embankment:
The composition and variability of the mine spoil used for borrow should be deter-
mined using test pits to observe and document the general proportion of slakable
and nonslakable materials, the type and amount of slakable material breakdown,
the grain-size distribution and moisture content of bulk samples, and the chemical
composition (i.e., pH and content of sulfde minerals in fne-grained constituents
such as shale) of the spoil mass. Placement of slakable material in upstream zones
(particularly when fne-grained) and nonslakable material in downstream zones in
an embankment can provide a means of isolating the materials and taking advantage
of their properties.
If oversize material (e.g., greater than 12 inches in maximum dimension) is present,
provisions to crush, mechanically degrade or isolate oversize materials should be
incorporated into the construction specifcations.
If the chemical composition of any portion of the spoil mass is not acceptable, these
materials should be isolated or excluded from the embankment.
The presence of mine spoil in the foundation zone of an impounding embankment may require mea-
sures to address setlement and the potential for internal erosion. The signifcance of such efects
should be evaluated based on exploration, testing and analysis of the foundation and embankment
< PREVIOUS VIEW
Chapter 8
8-58 MAY 2009
PRELIMINARY FIELD RECONNAISSANCE
EXAMINE ROCK EXPOSURES, IDENTIFY MAJOR
LITHOLOGIC UNITS; CHECK EXTENT AND MODE
OF WEATHERING (E.G., BLOCK FALLS, CHIPPING,
MUD FLOWS); CHECK LOCAL MINE SPOILING
OPERATIONS FOR SLAKING EFFECTS
DETERMINE LITHOLOGIC TYPES AND CORRELATE
WITH RECONNAISSANCE PROGRAM. CHECK FOR
CARBONATES USING HYDROCHLORIC ACID.
LIMESTONES AND NON-CALCAREOUS CEMENTED
SANDSTONES
EXPLORATION PROGRAM
NO FURTHER TESTING OR SPECIAL DESIGN
MEASURES RECOMMENDED
ALL FINE-GRAINED LITHOTYPES AND CALCAREOUS
CEMENTED SANDSTONES
PRELIMINARY LABORATORY PROGRAM
CONDUCT 1:1 pH AND CEC TESTS ON CRUSHED
POWDERED SAMPLES
1:1 pH > 7.8 or CEC > 15 M
EQ
/ 100g
DURABILITY TESTING PROGRAM
CONDUCT 5-CYCLE DURABILITY TEST (WET/DRY
OR RATE OF SLAKE), FRAGMENT SIZE > 1- INCH
(38 mm) AND SLAKE FLUID: WATER. OBSERVE
PATTERN AND RATE OF BREAKDOWN (NONE TO
SLIGHT, CHIP, PARTIAL TO COMPLETE DISAG-
GREGATION), COMPUTE 5-CYCLE DI
1:1 pH < 7.8 AND CEC < 15 M
EQ
/ 100g
NO FURTHER TESTING OR SPECIAL DESIGN
MEASURES RECOMMENDED
SETTLEMENT AND STABILITY PROB-
LEMS MAY DEVELOP; MECHANICAL
BREAKAGE DURING EXCAVATION,
HANDLING AND PLACEMENT MAY
BE SUFFICIENT TO MINIMIZE SLA-
KING AFTER BURIAL. NO SPECIAL
SEDIMENT CONTROL REQUIRED.
SETTLEMENT, STABILITY AND SED-
IMENT PROBLEMS MAY DEVELOP
IF LITHOTYPE COMPRISES MORE
THAN 50% OF THE OVERALL SPOIL
COMPOSITION.
(1)
SPECIAL DESIGN
AND MANAGEMENT TECHNIQUES
MAY BE REQUIRED.
SETTLEMENT, STABILITY AND SED-
IMENT PROBLEMS MAY DEVELOP
IF LITHOTYPE COMPRISES MORE
THAN 75% OF OVERALL SOIL COM-
POSITION.
(1)
SPECIAL DESIGN AND
MANAGEMENT TECHNIQUES MAY
BE REQUIRED.
NO SPECIAL DESIGN MEASURES
RECOMMENDED
CHIP SLAKE
PARTIAL OR
COMPLETE
DISAGGREGATION
CHIP SLAKE
PARTIAL
DISAGGREGATION
BREAKDOWN
SLIGHT TO NONE
FIGURE 8.16 CLASSIFICATION AND INTERPRETATION OF SPOIL DURABILITY
NOTE: 1. BASED ON HOMOGENEOUS MIXING
OF LITHOTYPES DURING PLACEMENT
(ANDREWS ET AL., 1980)
SLAKING MODE
DI > 50 DI < 50
SLAKING MODE
TESTING RESULTS
CEC = CATION EXCHANGE CAPACITY
M
EQ
= MILLIEQUIVALENT
DI = DEGRADATION INDEX = (1 - I
d
)
I
d
= SLAKE DURABILITY INDEX
FIGURE 8.16 CLASSIFICATION AND INTERPRETATION OF SPOIL DURABILITY
FIGURE 8.16 CLASSIFICATION AND INTERPRETATION OF SPOIL DURABILITY
< PREVIOUS VIEW
8-59
Site Mining and Foundation Issues
MAY 2009
materials. Typical measures that are employed include: (1) compensating for setlements through staged
construction by loading foundation zones using broad embankment widths and (2) providing sufcient
gradient on drainage structures. The potential for internal erosion can be addressed by a variety of site
preparation measures including the use of cutof trenches with seepage barriers and flters.
8.9 SURFACE MINE HIGHWALL ISSUES
Surface mine highwalls represent foundation concerns for coal refuse disposal facilities, in that the
change in rock surface elevation may represent an abrupt transition in an embankment abutment
with the potential for: (1) embankment cracking due to diferential setlement or (2) a zone of con-
centrated seepage due to the difculty of placement and compaction of fll materials close to a near-
vertical highwall. Additionally, surface mine benches and highwalls may contain fractured materials,
particularly if they have been subjected to auger or highwall mining activities. Subsidence associated
with such mine openings and remedial measures are discussed in Sections 8.4.2.8 and 8.5.1.3. The
discussion in this section focuses on abrupt rock transitions and associated seepage cutofs.
Whenever a coal refuse embankment or earthen dam abuts a surface mine bench or highwall, there
is a potential for performance problems due to the abrupt transition where the rock bench/highwall
adjoins the earthen embankment. These potential performance problems include:
Inadequate compaction of embankment materials against the rock surface suf-
fcient to provide a low-hydraulic-conductivity contact between the earthfll and
rock surfaces.
Incomplete flling of surface cavities and depressions in rock surfaces with compacted
embankment materials that can be sources of diferential setlement and seepage.
Overly steep rock surfaces and overhangs that limit the opportunity for maintain-
ing positive compressive pressure over the full contact between earthfll and rock
surfaces.
Earthfll placed over rock surfaces with steep or abrupt slope transitions that can
lead to diferential setlement, cracking and elevated seepage in the earthfll.
Highly fractured highwalls or benches resulting from mining or subsidence that rep-
resent potential seepage pathways that could impact downstream embankment and
abutment areas.
To minimize these potential performance problems, rock benches and highwalls should be inspected
during construction and should be treated or modifed so that unacceptable conditions and perfor-
mance problems do not occur. For earth embankment dams with an impervious core, these prob-
lems are of concern mostly where the impervious core abuts the rock surface. For coal refuse or
other embankments that are constructed as homogeneous dams, the concern may be more signifcant
because these structures do not have an impervious zone to serve as a barrier to internal seepage or
seepage along abutment contacts.
If the control measures described in the following list cannot be fully implemented in a constructed
embankment, special drainage features for intercepting and conveying seepage fows from the embank-
ment should be considered. Fell et al. (2005), USBR (1984) and Sherard et al. (1963) describe a number
of treatment and modifcation options for such performance problems, which may be encountered at
water retention dams. Additionally, alternate measures are described that may be appropriate for coal
refuse impoundments, but any such measure should be evaluated with respect to site-specifc condi-
tions. Guidance for addressing a number of conditions is provided in the following:
Loose and weathered materials The existing overburden and weathered rock
should be excavated to competent rock, and the resulting rock slope should be
< PREVIOUS VIEW
Chapter 8
8-60 MAY 2009
trimmed to a regular surface to eliminate depressions, overhangs, pinnacles or sharp
transitions. This treatment should focus on the abutment cutof zone, as illustrated
in Figure 8.17, and should extend as warranted by site conditions and the embank-
ment cross section. Sherard et al. (1963) noted that the most efective way to obtain
a tight bond between earthfll and a rock surface is to slope the rock surface suf-
ciently to permit each embankment layer to be compacted directly against the rock
using heavy compactors. Equipment operation above or below a highwall should be
preceded by an evaluation of the highwall stability and development of appropriate
procedures, if necessary, to address the threat of rockfalls or instability during grad-
ing and excavation.
Overly steep rock slope If the slope of a rock surface that will abut an earthfll is
steeper than 0.5H:1V, the rock surface should be fatened by excavation or backfll-
ing with concrete, or other measures should be taken to efectively place and com-
pact the embankment materials. Alternate measures could include incorporation of
impervious zones, incorporation of internal drainage features, or broadening of the
embankment. The same cautions discussed above relative to working near a high-
wall apply.
Hydraulic cutof If fatening is not practical due to the height of the rock slope, a
cutof keyway can be excavated into rock, or other measures to control seepage along
the rock interface may be implemented. The depth of a cutof keyway should extend
a sufcient depth (e.g., six feet as per Fell et al., 2005) into competent rock, and the
width of the keyway should be sufcient to permit compaction of earthfll in the
cutof in accordance with embankment criteria. Alternate measures to mitigate seep-
age could include incorporation of internal drainage features or broadening of the
embankment-abutment contact.
FIGURE 8.17 EXCAVATION SLOPES FOR PREPARATION OF DAM ABUTMENTS IN ROCK
FIGURE 8.17 EXCAVATION SLOPES FOR PREPARATION
OF DAM ABUTMENTS IN ROCK
ORIGINAL OVERBURDEN
SURFACE
(PRATT ET AL., 1972)
WEATHERED
ROCK SURFACE
1
1
2
1
1
2
NOTE: SEEPAGE CUTOFF NOT SHOWN.
SOUND
ROCK
SURFACE
FIGURE 8.17 EXCAVATION SLOPES FOR PREPARATION OF DAM ABUTMENTS IN ROCK
< PREVIOUS VIEW
8-61
Site Mining and Foundation Issues
MAY 2009
Steep, abrupt transitions in rock grade Steep, abrupt transitions in rock grade
should be trimmed (Figure 8.18) to minimize the potential for impacts of difer-
ential setlement in compacted earthfll susceptible to cracking (e.g., cohesive fll)
above the transition. The same cautions discussed above relative to working near a
highwall apply.
Clay-flled seams or very weathered rock Clay-flled seams or very weathered
rock should be excavated and flled with concrete (or grouted) to prevent erosion of
the seams. Treatment should focus on the abutment cutof zone and should extend
upstream and downstream to the extent warranted by the site conditions and the
embankment cross section. The U.S. Bureau of Reclamation (USBR, 1984) recom-
mends that seams narrower than 2 inches be cleaned to a depth of three times the
seam width and that seams between 2 inches and 5 feet wide be cleaned to a depth
of three times the seam width or to a depth where the seam is inch wide or less. To
avoid unnecessary excavation of stable seam material that would be held in place by
concrete and subsequent earthfll, engineering judgment should be used to deter-
mine the reasonableness of these guidelines in light of site-specifc conditions. Perin
(2000) presents a case history that addresses treatment of stress relief fractures at a
coal refuse disposal facility.
Irregularities on slopes not steeper than 0.5H:1V These irregularities may be
treated using dental concrete, pneumatically-applied mortar or slush concrete grout,
as illustrated in Figure 8.19. The extent of treatment can be limited to the abutment
cutof zone, as warranted, or may extend throughout the abutment area depending
on site-specifc conditions and the breadth of and material types in the embankment
cross section. Generally the following treatments are considered for structures that
FIGURE 8.18 SLOPE MODIFICATION TO REDUCE DIFFEREN-
TIAL SETTLEMENT
AND POTENTIAL FOR CRACKING
(FELL ET AL., 2005)
EMBANKMENT CREST
CRACKING DUE
TO DIFFERENTIAL
SETTLEMENT
POSSIBLE
ADDITIONAL
EXCAVATION
EXCAVATE
TO ELIMINATE
CRACKING
FIGURE 8.18 SLOPE MODIFICATION TO REDUCE DIFFERENTIAL SETTLEMENT
AND POTENTIAL FOR CRACKING
< PREVIOUS VIEW
Chapter 8
8-62 MAY 2009
may impose signifcant hydraulic head in the vicinity of the irregularity (e.g., water
retention dams):
Dental concrete should be used to fll joints, bedding, sheared zones, over-
hangs or excavated surfaces, particularly in the abutment cutof zone. Fell
et al. (2005) recommend that dental concrete slabs have a minimum thick-
ness of 6 inches, a minimum 28-day compressive strength of 3,000 psi, and
a maximum aggregate size not more than one-third the depth of the slab or
one-ffh the narrowest dimension between the rock surface and the edge of
the form. Feathering at the ends of slabs should not be permited, and slab
edges should be sloped no fater than 45 degrees. To achieve good bond
between the rock surface and the concrete, the rock surface should be thor-
oughly cleaned and moistened prior to concrete placement. The fnished con-
crete surface should have a roughened, broomed surface to facilitate earthfll
placement. Sulfate-resistant cement should be used in the concrete.
Pneumatically-applied mortar (shotcrete) can be used as an alternate to dental
concrete provided that care is taken that it is applied in a manner consistent
with the recommendations for dental concrete.
Slush concrete grout is a neat cement grout or sand-cement slurry used to fll
narrow surface cracks or to serve as a temporary cover over slakable materi-
als that degrade rapidly upon exposure to air and water. Slush grout may be
applied by brooming, troweling, pouring, rodding, or funneling into individual
cracks (Fell et al., 2005). To facilitate adequate bond, the rock surface and cracks
should be cleaned and moistened before the slush concrete grout is applied.
FIGURE 8.19 SLOPE MODIFICATION AND SEAM TREATMENT
FOR SEDIMENTARY ROCK STRATA
(FELL ET AL., 2005)
SHALE
SANDSTONE
SLAKING SHALE
INTERBEDDED
SILTSTONE
AND SHALE
SANDSTONE
CLEAN OUT WEATHERED SHALE AND SHOTCRETE TO FILL UNDERCUTTING
TRIM AND COVER WITH SLUSH
CONCRETE OR SHOTCRETE
FILL WITH CONCRETE
CLEAN OUT WEATHERED SEAM
TO DEPTH = 3X WIDTH OF SEAM
AND BACKFILL WITH CONCRETE
EMBANKMENT CREST
FLATTEN TO 0.5H TO 1V (LINE)
DRILL AND BLAST LIGHTLY
REMOVE SOIL, SOFT ROCK
TO CUT-OFF FOUNDATION
FILL WITH
CONCRETE
FLATTEN TO 0.5H TO 1V (LINE)
DRILL AND BLAST LIGHTLY
FIGURE 8.19 SLOPE MODIFICATION AND SEAM TREATMENT FOR SEDIMENTARY ROCK STRATA
In some situations, grouting of extensively fractured surface mine benches and highwalls may be
more appropriate than other treatments discussed above because of the depth or extent of distur-
bance. Table 8.6 presents general guidance for cement grout programs (Fell et al., 2005) that may
be useful for highly fractured rock foundations. USACE (1984a) and Fell et al. (2005) provide fur-
ther guidance for evaluation, design and implementation of grouting programs, including the use of
chemical grouts for limited applications.
FIGURE 8.19 SLOPE MODIFICATION AND SEAM TREATMENT FOR SEDIMENTARY ROCK STRATA
< PREVIOUS VIEW
8-63
Site Mining and Foundation Issues
MAY 2009
TABLE 8.6 GUIDELINES FOR CEMENT GROUT PROGRAMS
Staging of Grout
Program
Downstage
Top section of hole is drilled and washed, pressure tested, grouted and allowed
to set for 24 hours. Top section of grout is then washed out then second stage
is drilled and washed, pressure tested, grouted and allowed to set for 24
hours. Upper sections of grout are then washed out then third stage is drilled
and washed, pressure tested and grouted. Use of packer allows increased
pressures.
Upstage
Hole is drilled to full depth and washed, packer is seated at the top of the
bottom stage, pressure tested, grouted, and allowed to set for 6 hours. Set
packer at top of second bottom stage, pressure test, grout, and allow to set for
6 hours. Continue remaining stages.
Full Depth
Hole is drilled to full depth and washed, pressure tested, grouted. Only
recommended for consolidation grouting.
Erodibility of
Foundation
(1)
Pressure Test
Value Before
Grouting
(Lugeon)
(3)
Reduction in
Lugeon Value or
Grout Take from
Previous Stage
(2)
(Lugeon)
(3)
All Grout
Takes
(kg cement/
m)
Grout Hole
Spacing (m)
Closure Criteria Low/Non < 10 < 20 < 25 < 1.5
High < 7 < 15 < 25 < 1.5
Note: 1. Erodible foundations would include extremely or highly weathered rock and rock with
clay-flled joints that might erode under seepage fows.
2. For rock with joints closer than 0.5m.
3. Tabulated values are for Type F portland cement; for Type A portland cement,
adopt Lugeon values 20 percent greater. One Lugeon is a fow of 1 liter/minute/
meter of borehole under a pressure of 1000 kPa. In a 75-mm borehole, one Lugeon
equivalent to approximately 1.3x10
-7
m/sec hydraulic conductivity.
Depth and Lateral
Extent
So far as practical, grout holes should be taken to the depth and extent necessary to meet
closure criteria. Rules of thumb are not recommended. In nearly horizontally layered rock,
geologic interpretations may provide a guide where testing is unavailable in the valley foor, but
may be at different depths and orientations around the abutments due to the infuence of stress
relief, weathering, and rock types.
Cement Grout
Particle Size
Cement particles are mostly silt size, but include some fne sand particles
in conventional cement. With plasticizers, Type A and C Portland cements
have a maximum particle size of about 0.05 to 0.08 mm, while microfne
cements may be about 0.02 mm.
Fracture Size Minimum Lugeon values are indicative of rock that will accept
cement grout
Cement 1 Fracture/m 2 Fractures/m 4 Fractures/m
Type A 8 16 32
Type C 5 10 20
MC-500 (microfne) 3 5 10
Type A with dispersant 8 16 32
Type C with dispersant 5 10 20
MC-500 dispersant 1 2 4
Note: Fractures are assumed to be rough and uniform width and grout is assumed to have
been treated with plasticizer.
Grouting
Effectiveness
No further grouting is needed when:
< PREVIOUS VIEW
Chapter 8
8-64 MAY 2009
TABLE 8.6 GUIDELINES FOR CEMENT GROUT PROGRAMS
(Continued)
Approximate Penetration from Borehole of Grout (m)
Fracture Spacing
Grout
Penetration
Lugeons 1m 0.5m 0.25m
100 20 12 4
50 12 3 2
20 3 1.5 1
10 2 1 NP
5 1 NP NP
1 NP NP NP
Note: NP indicates that grout will not penetrate the fractures.
Practical Aspects
Grout Holes 30 to 60 mm percussion drilling and washing of borehole.
Standpipes Threaded galvanized pipe just larger than drill size, grouted into borehole to
enable near surface grouting.
Grout caps Necessary when grouting closely fractured or low strength rock where
standpipe cannot be sealed into rock.
Grout Mixers High speed, high shear, colloidal mixers.
Agitators Slow speed designed to prevent cement particle settling.
Grout Pumps Helical screw pumps or ram type pumps.
Packers Mechanical or infatable.
Water Cement
Ratio
Starting mix: most sites 2:1, for rock < 5 Lugeons 3:1, for rock > 30
Lugeons 1:1; for very high losses 0.8:1; for heavily fractured, dry rock
4:1, and for above water table where excess water is absorbed by dry
rock 5:1.
Thicken mix: (1) to deal with severe leaks, (2) after 1 hours with continued
take, or (3) if hole is rapidly taking grout (e.g., > 500 liters in 15 minutes).
Grout Pressure Recommend avoiding rock fracturing. Start at 100 kPa or less for 5 minutes,
then steadily increase over next 25 minutes. Occurrence of fracturing
can be detected by sudden loss of grout pressure at top of hole due to
increased take. Recommend grouting to refusal or minimum take.
Monitoring Parameters: (1) hole location, orientation, and depth, (2) stage depths,
(3) water pressure test value for each stage, (4) grout mixes, (5) grout
pressures (e.g., 15-minute intervals), (6) grouting times, (7) leaks, uplift,
(8) total grout take for each stage, (9) amount of cement in these takes,
and (10) cement takes/unit length of hole.
Water Pressure
Test
Before grouting, apply water pressure and monitor for 15 minutes.
Stage Lengths Based on geologic conditions, minimum drill run, allowable pressures in
upper part of hole, rock fracture conditions and hole stability, water fows
into the hole, and large water pressure tests or grout takes.
(ADAPTED FROM FELL ET AL., 2005)
< PREVIOUS VIEW
Chapter 9
HYDROLOGY AND
HYDRAULICS
9-1 MAY 2009
Coal refuse impoundments and embankments must handle the runof from precipitation that occurs
over the contributing watershed area. If not properly controlled, runof can jeopardize the collection
and conveyance system (channels and conduits). For impoundments, runof can cause the embank-
ment to be overtopped with the potential for failure. The principles of hydrology and hydraulics can
be used to determine and design the required combination of fow capacity and freeboard and to
select durable channel lining systems. The discussion of technical issues in this chapter is based on
the assumption that the reader is experienced in the technical areas of hydrology and hydraulics and
is familiar with the selection of hydrologic and hydraulic design parameters and the use of related
computer sofware. A number of traditional design concepts are reviewed herein, and reference is
made to additional resource materials.
The design of coal refuse disposal facilities requires a somewhat specialized approach. There are
many possible combinations of disposal facility confguration, facility staging, environmental con-
siderations and unique characteristics and properties associated with each site. Therefore, one of the
major aims of this chapter is to relate fundamental engineering principles to the unique requirements
of refuse disposal facility site design. While primarily focused on slurry impoundments, the contents
of this chapter are also applicable to other mining dams and impoundments.
The hydrologic and hydraulic information and design procedures presented in this chapter fall into
fve interrelated categories, as follows:
Basic defnitions and principles Sections 9.1 and 9.2 defne basic terms and condi-
tions applicable to coal refuse disposal facilities that relate to hydrologic and hydrau-
lic features. Table 9.1 presents a complete summary of hydrologic and hydraulic
planning and design procedures. The table also serves as an outline of this chapter
and a summary of supplemental references. The fundamental interrelationships of
runof, reservoir storage, and outfow are established. The major elements that may
afect these interrelationships at coal refuse disposal facilities are also discussed.
General design considerations Section 9.3 identifes regional and site conditions
that afect the suitability of various hydraulic conveyance structures for coal refuse
disposal facilities. In Section 9.4, these broad concepts are extended to consider the
efect of disposal facility confguration upon selection of suitable hydraulic convey-
< PREVIOUS VIEW
9-2
Chapter 9
MAY 2009
ance structures. Characteristics that distinguish coal refuse disposal facilities from
conventional embankment dams are emphasized.
Design-storm criteria Section 9.5 presents design storm precipitation criteria for
coal refuse disposal facilities. Factors such as location, facility size, and hazard
potential are discussed. Design storm criteria for short-term conditions and for
minor hydraulic structures are also addressed.
Procedures for analysis Sections 9.6, 9.7 and 9.8 discuss analytical procedures for
evaluation and design of coal refuse disposal facility hydraulic structures. Methods
for determining runof based on predicted precipitation are frst established, fol-
lowed by reservoir storage and outfow capacity requirements. Various components
of outfow structures are discussed in detail. Procedures for routing storm runof
through an impounding disposal facility and optimizing reservoir storage and out-
fow are presented.
Dam-breach analysis Section 9.9 discusses procedures for evaluation of dam breach
and potential downstream inundation for the determination of hazard potential and
for Emergency Action Plan (EAP) preparation.
9.1 GENERAL CONSIDERATIONS
The hydrologic and hydraulic design and analysis procedures discussed in this chapter apply to both
existing and new coal refuse disposal facilities. The sequence presented in Table 9.1 is normally fol-
lowed either for modifying an existing disposal facility or for constructing a completely new disposal
facility. It should be recognized that sequencing of a modifcation to an existing coal refuse disposal
facility should be continually coordinated with the ongoing mining and coal preparation operations.
The designer of a new coal refuse disposal facility normally has fexibility in site selection, stag-
ing of the embankment growth and long-term planning of related hydraulic structures. Given this
fexibility, design food requirements can typically be met throughout the entire life of the disposal
facility. Ofen the designer is able to optimize the relationships between refuse disposal operations,
embankment design, hydraulic structure construction, and the overall mining and coal preparation
operations.
A designer modifying an existing disposal facility should frst determine its conformance with cur-
rent design storm criteria and should then assess options for any necessary upgrade of the runof
collection and control system. Sometimes a facility has limited storage or hydraulic conveyance capa-
bility, may not satisfy current design and regulatory requirements, and cannot be easily modifed in a
short period of time. An efective solution may be to perform a staged modifcation program, as part
of continued refuse disposal operations, which may in fact provide materials necessary for increasing
freeboard and constructing diversions, thus improving hydraulic capacity. Under such conditions,
the modifcations to the facility are usually required to meet or exceed MSHAs short-term hydrologic
design criteria, as subsequently described in Section 9.5.2.
9.2 HYDROLOGY AND HYDRAULICS PRINCIPLES
Hydrology is the study of climatic and physical conditions that govern natural fows in rivers, streams
and channels. Hydrologic analyses are used to determine the probable and possible direct runof to
a particular site from natural causes such as precipitation or snow melt. Hydraulics is the study of
water fows in channels and conduits. Hydraulic engineering is used in the design of decant systems,
outlet works, spillways, ditches, channels, diversion structures, and other systems for controlling
fowing waters. An integrated application of hydrology and hydraulics is necessary for the develop-
ment of safe, economical and environmentally acceptable coal refuse disposal facilities.
< PREVIOUS VIEW
9-3
Hydrology and Hydraulics
MAY 2009
TABLE 9.1 HYDROLOGIC AND HYDRAULIC DESIGN PROCEDURES FOR
COAL REFUSE DISPOSAL FACILITIES
Design Considerations
Applicability
Manual
Sections for
Reference
Supplemental
References
All
Facilities
Impounding
Facilities
I. Determine Importance of Hydrologic and
Hydraulic Considerations
Type of facility X Chapter 3, 9.4 USBR (1987a)
Impounding vs. non-impounding 9.4 USBR (1987a)
Site conditions X Chapter 5, 9.5 USBR (1987a)
Downstream conditions X (Of particular
concern)
9.3, 9.5 USBR (1987a)
Startup, operation, abandonment
requirements
X Chapters 4,
6, 9
II. Establish Preliminary Facility
Confguration and Hydraulic Systems
Select structure type X Chapter 3, 9.4
Balance availability of materials for
embankment construction with facility staging
X Chapter 5,
9.3, 9.4
Determine size and potential hazard
classifcation based on dam breach analysis
and downstream inundation
X 9.5, 9.9 FEMA (2004a)
Determine appropriate design storm for long-
term operation
X 9.5 MSHA (2007)
Determine if separate design consideration
should be given to short-term conditions with
lesser design storm at any time during the
operational period of the facility
X 9.4, 9.5 MSHA (2007)
Calculate watershed contributing to major
hydraulic systems
X 9.3
Determine approximate infow rates and
volumes to be controlled by major hydraulic
systems from design storm criteria
X 9.6 NWS (2006a,b)
NRCS (2004b)
Evaluate alternative combinations of spillway
outfow and impoundment storage capacities
X 9.6 to 9.8 USBR (1987a)
NRCS (2004b)
Brater et al. (1996)
Determine preliminary spillway type, location
and approximate size (for all stages of
operation)
X Chapter 5,
9.6 to 9.8
USBR (1987a)
Brater et al. (1996)
Determine preliminary decant type, location
and approximate size (for all stages of
operation)
X Chapter 5,
9.6 to 9.8
USBR (1987a)
Brater et al. (1996)
< PREVIOUS VIEW
9-4
Chapter 9
MAY 2009
Design Considerations
Applicability
Manual
Sections for
Reference
Supplemental
References
All
Facilities
Impounding
Facilities
Determine magnitude of storm that can be
controlled and compare with appropriate
design storm for facility size and potential
hazard classifcation
X 9.5 to 9.9 USBR (1987a)
NRCS (2004b)
Evaluate modifcations to be made to
improve the facilitys hydraulic system
X Chapter 6,
9.5 to 9.8
USBR (1987a)
Evaluate advantages and disadvantages of
modifying the facility for continued use or to
a satisfactory confguration for abandonment
X Chapter 6,
9.4 to 9.8
Assign appropriate long-term design storm or
abandonment criteria
X 9.5 FEMA (2004c)
III. Determine Design Infow Rates and
Volumes for Major Hydraulic Systems
Determine if key parameter curves are
suitable for fnal design for any or all stages,
including abandonment
X 9.4, 9.6
Determine infow hydrograph parameters, if
required, for any stage of development
X 9.6 NRCS (2004b)
USBR (1987a)
IV. Design Major Hydraulic Systems
Design major diversion system to insure
against failure during appropriate design
storm
Collection of inlet area X 9.7 Chow (1959)
Establish control section of fow (inlet,
transport section or outlet)
X 9.7 USBR (1987a)
Henderson (1966)
Brater et al. (1996)
Determine requirements to prevent
failure by overtopping, erosion or
clogging
X 9.6 to 9.8 USBR (1987a)
Determine downstream outlet and/
or discharge requirements to avoid
unacceptable damage at design fow
X 9.7, 9.8 Chow (1959)
USBR (1987a)
Brater et al. (1996)
Determine optimum combination of storage
and outfow for each stage of development
(for impoundments)
Perform reservoir routing analysis of
infow hydrograph
X 9.6 to 9.8 USBR (1987a)
Design the spillway system for the
appropriate design storm for each stage of
development
TABLE 9.1 HYDROLOGIC AND HYDRAULIC DESIGN PROCEDURES FOR
COAL REFUSE DISPOSAL FACILITIES
(CONTINUED)
< PREVIOUS VIEW
9-5
Hydrology and Hydraulics
MAY 2009
Design Considerations
Applicability
Manual
Sections for
Reference
Supplemental
References
All
Facilities
Impounding
Facilities
Establish control section for all fow
conditions to assure adequate capacity
X 9.6 to 9.8 Chow (1959)
USBR (1987a)
Henderson (1966)
Brater et al. (1996)
Design the inlet including provisions to
prevent clogging
X 9.8 USBR (1987a)
Design the outlet to prevent
unacceptable damage at magnitude of
fow
X 9.8 FHWA (2006)
USBR (1987a)
Design the decant system for normal
operating conditions and to evaluate
impoundment storage of design storm
Establish fow control for all storage
levels to assure adequate capacity
X 9.6, 9.8 USBR (1987a)
Brater et al. (1996)
Design the inlet, including provisions to
avoid clogging
X 9.8 USBR (1987a)
Brater et al. (1996)
Design the transport section,
considering structural stability,
corrosion resistance, and capacity
X 9.8 USBR (1987a)
Brater et al. (1996)
FHWA (2005b)
Design the outlet to prevent
unacceptable damage
X 9.8 USBR (1987a)
FHWA (2006)
Perform dam breach analysis and evaluate
downstream inundation
X 9.9 FEMA (2004c)
V. Design Minor Hydraulic Systems
Surface drainage ditches that are not critical
to safety during design storm
X 9.6, 9.8 USBR (1987a)
FHWA (2005a)
FHWA (2006)
Minor roadway culverts X 9.6, 9.8 FHWA (2005b)
FHWA (2006)
Weirs to separate seepage from large fows,
if required, for environmental control
X 9.8 Henderson (1966)
TABLE 9.1 HYDROLOGIC AND HYDRAULIC DESIGN PROCEDURES FOR
COAL REFUSE DISPOSAL FACILITIES
(CONTINUED)
9.2.1 Basic Design Principles
The fundamental principle governing the hydrologic and hydraulic design of a coal refuse disposal
facility is that runof, natural drainage and process water must be conveyed past the embankment,
stored within the facility impoundment(s), or handled by a combination of these two methods. The
hydrologic characteristics of the applicable watershed (rainfall, tributary area, land use cover condi-
tions, soil type, slope, etc.) determine the runof hydrograph, while the physical dimensions and
hydraulic characteristics of the facility and hydraulic structures determine the required conveyance
and storage capacity. Table 9.2 presents a summary of the application of the basic design principles
to coal refuse disposal facilities.
< PREVIOUS VIEW
9-6
Chapter 9
MAY 2009
TABLE 9.2 HYDROLOGIC AND HYDRAULIC DESIGN CONSIDERATIONS
FOR COAL REFUSE DISPOSAL FACILITIES
Embankment Type
(1)
Runoff, Outfow and Storage Considerations
Valley-Fill and Side-Hill
Non-Impounding
Embankments
If placement of the embankment is started at the upper end of the valley, runoff
from the natural watershed can be diverted around the embankment and no
water has to be stored. Precipitation on the embankment can be directed
downstream.
If placement is started by forming a downstream embankment, it will have a
temporary character with interim diversion ditches sequentially replaced as the
fll is raised (sometimes the fnal diversion ditches are installed initially). Pre-
cipitation and runoff on the embankment are directly discharged downstream
with the intervening drainage between the fnal diversion ditch and the interim
diversion ditches.
Ridge and Heaped
Non-Impounding Embankments
Ridge and heaped embankments that are constructed above the natural
topography only have infow associated with direct rainfall onto the dis-
posal area. Precipitation and runoff on the embankment can be directed
downstream.
Cross-Valley
Impounding
Embankment
The cross-valley impounding embankment presents a variety of alternatives for
handling hydrologic events. Infow may include precipitation from upstream of
the embankment, including the drainage area above diversion ditches, unless
the ditches are designed not to fail from the design storm.
For a cross-valley impoundment, the three possibilities for handling design
storm infow are:
1. If the embankment crest elevation is maintained suffciently high above the
pool level, all runoff from the design storm can be stored, such that outfow
is not a requirement during the design storm. The impounded water can
then be lowered gradually by fow through a decant system.
2. If a spillway of adequate size is constructed with its crest at the normal
pool level of the impoundment, all of the storm runoff can be passed
directly through the disposal area and the storage requirement will be
minimal.
3. If the spillway crest is located above the normal pool level, but the storage
volume between the pool and spillway elevations is less than the infow
volume, the spillway must be designed to conduct a volume equal to the
difference between infow volume and storage volume in an appropriate
time interval.
Side-Hill Impounding Embankment
A side-hill impounding embankment can have all of the alternatives of a cross-
valley impoundment except that the smaller watershed and the potential for
diversion signifcantly reduce the storage and outfow requirements associated
with the design storm.
Diked-Pond
Embankment
Normally, a diked-pond embankment will have infow equal to the precipitation
falling directly into the impoundment. Total storage with limited or no outfow
during the design storm is normally the best solution, although the drawdown
requirement must be met by either a spillway, decant pipe, or pumping.
Incised Pond
An incised pond has a water surface below the normal ground surface, and
infow runoff, storage and outfow generally are not critical to safety.
Note: 1. Embankment types are discussed in Chapter 3.
< PREVIOUS VIEW
9-7
Hydrology and Hydraulics
MAY 2009
In general, for non-impounding coal refuse disposal facilities or the downstream or perimeter portions
of impounding facilities or slurry cells, runof is conveyed around the facility without retention and
storage. On the other hand, impounding embankments are designed to temporarily store runof from
upstream areas and to convey excess fows past the embankment with decant pipes and spillways.
Design criteria for impounding and non-impounding coal refuse disposal facilities include the total
volume of runof from the design storm, as discussed in Section 9.5. For a non-impounding coal
refuse disposal facility, the peak runof rate caused by a food or the design storm is of prime concern.
For an impounding facility, both the peak runof rate and the total volume of runof are of concern.
In the frst case, the hydraulic facilities must be sized to pass the peak runof rate, while in the later
case, the impoundment and hydraulic structures must be designed to store and pass the total volume
of runof.
The runof and outfow elements are infuenced by a number of critical factors, as discussed in the
following section.
9.2.2 Defnition and Discussion of Key Runof Elements
Sources of impoundment infow are shown in Figure 9.1. These sources also include ancillary fow
contributions such as process water (water or water-slurry mixture pumped from the mine or the
coal processing plant), indirect runof from adjacent watersheds, or other diverted fows such as from
underground mines. The sources of impoundment infow can be categorized as follows:
Major Sources
Direct precipitation rain or snow falling directly onto the disposal site
Runof from precipitation falling on areas upstream or upgradient from the site
and within the watershed associated with the facility
Minor Sources
Springs from groundwater fow
Base fow in a stream passing through or by the site that is relatively independent of
the most recent rainfall events, but directly related to infltration associated with ear-
lier rainfall events
Process water and other pumped fows
Minor sources of fow are typically much smaller than the major sources of runof. The volumes
associated with minor sources can be determined with relative accuracy. However, the amount of
runof resulting from a storm will vary depending upon site location. Geographic location, climatic
conditions and watershed characteristics all contribute to storm runof, as discussed in the follow-
ing sections.
9.2.2.1 Watershed Boundary and Area
The watershed is all of the catchment area that drains toward a particular point of interest. Watershed
boundaries are typically determined from site-specifc topographic maps (Section 6.4.1.1) or USGS
topographic quadrangle maps, as shown in Figure 9.2.
9.2.2.2 Precipitation
Runof results from precipitation falling on the watershed, melting of snow already on the ground
and outfow from upstream impoundments in the watershed. Snowmelt is usually a minor portion of
runof in small watersheds, such as those usually associated with a coal refuse disposal facility. The
efects of upstream impoundments should be considered on an individual site basis.
< PREVIOUS VIEW
9-8
Chapter 9
MAY 2009
FIGURE 9.1 RUNOFF AND IMPOUNDMENT INFLOW SOURCES
NET INFLOW TO
IMPOUNDMENT
FIGURE 9.1 RUNOFF AND IMPOUNDMENT INFLOW SOURCES
WATERSHED
RUNOFF
PRECIPITATION
RAINFALL
SNOWMELT
OUTFLOW
SPILLWAYS & DECANTS
STORAGE
RETENTION IN
IMPOUNDMENT
UPSTREAM
IMPOUNDMENT
RELEASE
SPRINGS, PROCESS
WATER, STREAM INFLOW
GAINS
INFILTRATION, SURFACE
RETENTION, UPSTREAM
IMPOUNDMENT
DIVERSION AROUND
IMPOUNDMENT
LOSSES
RELEASED AT
CONTROLLED RATE
FIGURE 9.1 RUNOFF AND IMPOUNDMENT INFLOW SOURCES
< PREVIOUS VIEW
9-9
Hydrology and Hydraulics
MAY 2009
9.2.2.2.1 Rainfall Curves
Calculation of the design storm rainfall (Section 9.5) involves determination of the total amount and
distribution of rainfall for the entire storm duration. The relationships between total precipitation
(cumulative rainfall depth), storm duration, and storm intensity (slope of the rainfall distribution
curve) have a direct efect on the runof rate and volume (Sections 9.6.1 and 9.6.2). For example, a
sudden short rainfall can result in a high runof rate and a small total volume of runof, while a pro-
longed rainfall of low intensity can produce a large total volume of runof with a relatively low runof
rate. Coal refuse disposal facilities should be designed to accommodate all possible precipitation/
runof conditions associated with the design storm.
9.2.2.2.2 Rainfall Intensity
The relationships between rainfall intensity, duration of the rainfall event and frequency (i.e., inten-
sity-duration-frequency or I-D-F) can be used to determine the peak runof, and are useful in the
FIGURE 9.2 WATERSHED BOUNDARY DELINEATED
ON USGS TOPOGRAPHIC MAP
LIMITS OF
WATERSHED
LONGEST
WATERCOURSE
POINT OF INTEREST
FIGURE 9.2 WATERSHED BOUNDARY DELINEATED ON USGS TOPOGRAPHIC MAP
FIGURE 9.2 WATERSHED BOUNDARY DELINEATED ON USGS TOPOGRAPHIC MAP
< PREVIOUS VIEW
9-10
Chapter 9
MAY 2009
design of hydraulic structures such as culverts, channels and ditches (Sections 9.6.3 and 9.6.4). Only
the most intense portion of the rainfall, not the entire storm history, governs the selection of culvert
size, the most efcient ditch or channel confguration, and the required erosion protection associated
with the runof fow velocity.
9.2.2.3 Watershed Characteristics
A portion of the precipitation falling on a watershed is retained in the soil and by vegetation or may
be retained in upstream impoundments. The portion of the precipitation that fows to the point of
interest is termed the runof. The watershed characteristics that determine the diference between the
amount of precipitation falling on the watershed and the amount that becomes runof include: (1) the
types of surfcial soils and their efect on infltration; (2) the condition of the ground surface (e.g., wet,
dry, snow-covered or frozen) prior to the precipitation (termed the antecedent moisture condition);
(3) the type and density of vegetation; (4) development features such as paved surfaces, channeling,
storm sewers, etc.; and (5) the presence of dams, lakes, ponds or swamps upstream from the disposal
facility that can either store water and release it at a slow rate or fail and release large volumes of
stored water at a high rate.
The runof hydrograph at the point of interest will vary as a function of the intensity distribution of
precipitation and the geometric shape and slope conditions of the watershed area. Infow and out-
fow hydrographs for a typical impoundment are shown in Figure 9.3. The fgure also shows the net
infow and volume of impoundment storage.
FIGURE 9.3 TYPICAL IMPOUNDMENT INFLOW AND OUTFLOW HYDROGRAPHS
FIGURE 9.3 TYPICAL IMPOUNDMENT INFLOW
AND OUTFLOW HYDROGRAPHS
INFLOW
HYDROGRAPH
F
L
O
W
R
A
T
E
TIME
VOLUME OF INFLOW
VOLUME OF STORAGE
VOLUME OF OUTFLOW
PEAK INFLOW RATE
OUTFLOW
HYDROGRAPH
PEAK OUTFLOW RATE
FIGURE 9.3 TYPICAL IMPOUNDMENT INFLOW AND OUTFLOW HYDROGRAPHS
< PREVIOUS VIEW
9-11
Hydrology and Hydraulics
MAY 2009
9.2.3 Key Storage and Outfow Elements
The principal factors governing the storage capacity of a reservoir or impoundment are the physical
dimensions of the embankment and ground surface and the current level of water/slurry. The out-
fow capacity is determined by the types and sizes of the hydraulic structures.
9.2.3.1 Impoundment Capacity
The storm storage capacity of an impoundment is the volume of runof that can be temporarily
retained during the applicable design storm. If a refuse disposal facility has minimal storage capacity,
the outfow hydrograph will be about the same as the infow hydrograph, and the hydraulic struc-
tures must be designed to transport the peak runof rate. The primary beneft of impoundment stor-
age is that the outfow rate can typically be reduced, permiting use of smaller hydraulic structures.
An additional beneft is that the downstream fooding risk is not exacerbated by an increase in runof
from disturbed areas in the watershed. The potential diference in peak fow rates is illustrated by the
infow and outfow hydrographs shown in Figure 9.3.
Figure 9.4 shows a typical impoundment capacity curve relating storage volume to pool elevation.
The fgure also shows the relationship between reservoir surface area and pool elevation. Such curves
are used to evaluate the storage conditions at any given pool elevation and are prepared as part of the
design of an impounding structure (Section 9.7).
Two terms used to describe the limits of acceptable pool elevation are surcharge and freeboard.
Surcharge is the vertical distance between the usual operations level of the impoundment and the
maximum allowable water surface elevation. Normal freeboard is the vertical distance between the
pool elevation and the top of the embankment at its lowest point (where the dam would begin to
be overtopped). Design storm freeboard is the vertical distance between the maximum water sur-
face elevation during the design storm and the top of the embankment. The minimum design storm
freeboard is an impoundment design criterion and should be such that waves do not overtop the
embankment crest during the design storm. Freeboard also serves to compensate for uncertainty in
hydrologic parameters.
9.2.3.2 Decants, Principal Spillways and Auxiliary Spillways
Decants are conduits that extend through an embankment and discharge under controlled conditions
at or beyond the embankment toe. As the term decant would imply, impoundment water typically
enters the conduit by fowing over the top edge of the upstream end. At coal refuse disposal facili-
ties, decants are generally not intended to discharge at high fow rates, but are designed to remove
clarifed process water, pass base stream fows or to drain the impoundment of stored water afer a
storm. However, a decant must be sufciently large that stored water from the design storm can be
drained within a reasonable period of time, so that the storage volume needed for a subsequent storm
is available. Several types of decant systems are shown in Figure 9.5.
Principal spillways are generally designed to control the discharge associated with large design
storms, to limit discharges and associated impacts downstream, and to limit the frequency and dura-
tion of fow through the emergency (auxiliary) spillway. Principal spillways are most ofen associ-
ated with fresh water impoundments and sedimentation and treatment ponds, and state regulatory
agencies typically provide specifc design storm criteria that govern the size and capacity of these
structures. In some coal refuse facility designs, the decant may also function as a principal spillway.
Principal spillways are designed to: (1) release runof at a controlled rate, (2) provide setling time for
the removal of sediment or process water solids prior to discharge, (3) provide runof detention, and
(4) function as decants to control the impoundment operational pool level. Decant systems are gen-
erally not considered in design storm food routing analyses for determining maximum impound-
ment pool level, as they do not have signifcant discharge capacity. If considered in the food routing
< PREVIOUS VIEW
9-12
Chapter 9
MAY 2009
analysis, the decant pipe should be of sufcient size that clogging is unlikely (typically, greater than
12 inches in diameter) and should be equipped with a properly designed trashrack.
Auxiliary (emergency) spillways are open channels generally used to discharge that portion of the
runof volume that cannot be stored in the impoundment or routed through the principal spillway.
Auxiliary spillways typically are capable of discharging: (1) moderate fows from storms much
smaller than the design storm (Section 9.5.1) with litle or no damage or (2) large fows resulting from
the design storm, where some localized damage may occur, but without the threat of failure of the
entire impounding embankment. Typical auxiliary spillway systems are shown in Figure 9.6.
FIGURE 9.4 TYPICAL IMPOUNDMENT AREA AND STORAGE VOLUME CURVES
FIGURE 9.4 TYPICAL IMPOUNDMENT AREA AND STORAGE VOLUME CURVES
STORAGE VOLUME (ACRE-FT)
IMPOUNDMENT SURFACE AREA (ACRES)
W
A
T
E
R
S
U
R
F
A
C
E
E
L
E
V
A
T
I
O
N
(
F
E
E
T
,
M
S
L
)
0 100 200 300 400 500
V
50 40 30 20 10 0
675
670
665
660
655
650
645
STORAGE VOLUME IMPOUNDMENT
SURFACE AREA
STORAGE VOLUME (V) AT WATER
SURFACE ELEVATION (h)
h
FIGURE 9.4 TYPICAL IMPOUNDMENT AREA AND STORAGE VOLUME CURVES
< PREVIOUS VIEW
9-13
Hydrology and Hydraulics
MAY 2009
DECANT PIPE
OUTLET
RIPRAP
9.5b DECANT THROUGH REFUSE EMBANKMENT
FIGURE 9.5 TYPICAL DECANT SYSTEMS
9.5c DECANT THROUGH SADDLE
RIPRAP
DECANT PIPE
OUTLET
DECANT PIPE
INLET
DECANT INLET ON
HILLSIDE SLOPE
9.5a DECANT FOR DIKED IMPOUNDMENT
NOTE: SPILLWAY TYPICALLY
NOT REQUIRED
RIPRAP
DECANT PIPE OUTLET
DECANT INLET ON
UPSTREAM SLOPE
FIGURE 9.5 TYPICAL DECANT SYSTEMS
FIGURE 9.5 TYPICAL DECANT SYSTEMS
< PREVIOUS VIEW
9-14
Chapter 9
MAY 2009
The design of auxiliary spillways, principal spillways and decants normally requires evaluation of
three basic components: the inlet, the transport section and the outlet. Key processes in the design of
these systems include:
Determining which component controls the outfow rate for various fow conditions.
Sizing each component to function properly for the anticipated range of fow condi-
tions.
Specifying materials for each component that will not erode excessively under the
anticipated fow velocities.
SPILLWAY CUT
INTO ROCK
DECANT INLET
POSITIONED ON
HILLSIDE SLOPE
DECANT PIPE OUTLET
FIGURE 9.6 TYPICAL DECANT AND SPILLWAY SYSTEMS
DECANT PIPE OUTLET
EXCAVATED GRASS-LINED
SPILLWAY TO ADJACENT VALLEY
DECANT INLET
POSITIONED ON
HILLSIDE SLOPE
FIGURE 9.6 TYPICAL DECANT AND SPILLWAY SYSTEMS
FIGURE 9.6 TYPICAL DECANT AND SPILLWAY SYSTEMS
< PREVIOUS VIEW
9-15
Hydrology and Hydraulics
MAY 2009
Designing inlet or transport sections that will not become clogged or otherwise fail,
causing major downstream damage or failure of the impounding embankment.
Arranging the outlet location so that the release of water does not lead to failure of
the impoundment embankment or major downstream damage.
The design of outfow systems is further discussed in Section 9.7.
9.3 GENERAL CONSIDERATIONS FOR COAL REFUSE DISPOSAL FACILITIES
9.3.1 Special Characteristics
Table 9.3 lists characteristics that distinguish the design of typical coal refuse disposal facilities from
many other structures with appurtenant hydraulic structures. In addition to the special characteris-
tics indicated in Table 9.3, the hydrologic and hydraulic design of coal refuse disposal facilities is also
governed by the considerations discussed in the following sections.
9.3.2 Site Conditions
Site selection impacts the cost and difculty in providing adequate hydraulic appurtenant structures
for use during the disposal period and subsequent abandonment of a coal refuse disposal facility.
Based upon hydrologic and hydraulic considerations, the best site will almost always have the small-
est possible upstream watershed. In some cases, however, hydrologic/hydraulic considerations are
secondary to preparation plant location and materials handling requirements. Even if this appears to
be the case, the designer should evaluate the needed hydraulic structures considering downstream
hazard potential, environmental control and construction costs prior to fnalizing the location for a
disposal facility. Large initial costs associated with the construction of hydraulic structures may be
justifed if this allows materials transportation costs to be lowered.
TABLE 9.3 FACILITY CHARACTERISTICS INFLUENCING HYDRAULIC SYSTEM DESIGN
Characteristic Signifcance In Design
The facility is designed for disposing coal refuse,
with active operations taking place for an associated
period of time, and not to collect water for food
prevention, water supply, power, or recreation.
Greater fexibility in choosing location, confguration and
construction sequence for appurtenant hydraulic structures.
The facility covers a large area, with the gradient or
drainage slope primarily in one direction.
Providing diversion facilities not subject to localized failures or
controlled overtopping during large storms is often not econom-
ically practical.
The placement of refuse occurs over many years,
during which time the facility confguration is
constantly changing.
Hydraulic systems must be designed so that they can be
expanded or decommissioned and replaced as the facility
grows.
The growth rate of the facility is estimated based
upon projected quantities of refuse production.
Actual quantities must be evaluated periodically to determine if
the rate of construction is adequate.
Water passing over or through the coal refuse can
be destructive or environmentally unacceptable.
Proper design requires that potentially adverse environmental
effects (e.g., corrosion of construction materials), and the
cost of water collection and treatment, be considered in the
evaluation of alternative hydraulic systems.
When placement of refuse is completed, the facility
typically has no continuing utility, and the hydraulic
systems are decommissioned, the impounding
capability is eliminated, and the site is abandoned in
accordance with mine reclamation requirements.
The sequence of constructing hydraulic systems must provide
an arrangement that will function until decommissioning at a
specifed future date. Planning must allow for the possibility
that decommissioning, elimination of impounding capability,
and abandonment may be required for a confguration either
larger or smaller than originally anticipated.
< PREVIOUS VIEW
9-16
Chapter 9
MAY 2009
The site conditions described in the following sections may afect decisions related to the selection
and design of facility hydraulic structures.
9.3.2.1 Topography
The importance of topography on the geotechnical aspects of site selection and disposal facility
confguration is discussed in detail in Section 6.2.2.1. As discussed in this section, the signifcance of
topography is generally limited to the planning, design and construction of hydraulic conveyance
structures.
9.3.2.1.1 Steep Terrain
In areas of steep and rugged terrain, many disposal facilities must necessarily be located in valleys
formed by small streams. Two very signifcant problems may be encountered with respect to placing
diversion ditches, spillways and conveyance channels in these areas:
Channels cut into the side of the valley will require the excavation of large amounts
of material, as illustrated in Figure 9.7a. With increased channel width, the cut
becomes more extensive and the slope of the cut must ofen be decreased to achieve
stability. These conditions combine to increase excavation quantities and costs dis-
proportionately to the fow capacity gained.
The potential for sloughing of overburden soil or weathered rock into the channel,
thus restricting its fow capacity, is increased, as illustrated in Figure 9.7b. Major
sloughing will ofen occur during a heavy rainstorm when a large fow capacity is
desired. The possibility of main spillway channels becoming obstructed by slough-
ing must be considered in the geotechnical analysis and design of the cut slope.
Diversion ditches for non-impounding coal refuse embankments are designed based on the design
storm (100-year-recurrence-interval storm). For impounding coal refuse facilities with more extreme
design storms such as the Probable Maximum Precipitation (PMP), it is usually not feasible to design
perimeter diversion ditches large enough to pass the maximum fow. While diversion ditches for
impounding facilities still perform an important function, the hydraulic design of the impoundment
generally is based on the assumption that during large foods the diversion ditches will be over-
topped and the resulting overfow will enter the impoundment. However, such overtopping should
not be permited to occur if fows in excess of the diversion ditch design storm could cause erosion
of the dam and spillway.
Although there are undesirable aspects to cuting channels into steep slopes, there may also be sig-
nifcant advantages. Bedrock is normally found near the surface in rugged terrain. Thus, channels cut
in such areas will ofen be resistant to erosion without special protection. A channel should be located
where its base will be on the most resistant material. If possible, channels should be constructed in
sound rock, particularly where fow velocities will be erosive and where failure of the channel would
create an unsafe general condition or large repair costs. Another advantage that may be realized from
cuting channels into such slopes is the concurrent production of borrow materials suitable for use
as resistant drainage material.
9.3.2.1.2 Gently Sloping Terrain
In gently sloping terrain, the disadvantages associated with hillside channel excavation are not as
pronounced as in steep terrain. As shown in Figure 9.8a, the volume of excavation is a nearly linear
function of channel width. In addition, achieving stability of the uphill cut slope is not as difcult
as for channels cut into steep hillsides. However, as illustrated in Figure 9.8b, these areas ofen do
not have rock near the surface. Therefore, the channels are more susceptible to erosion unless fow
< PREVIOUS VIEW
9-17
Hydrology and Hydraulics
MAY 2009
velocity can be kept low or some type of stabilization system (e.g., channel lining) is provided. A key
to economical design in this case is to minimize the length of channel sections where fow velocities
exceed the natural erosion resistance of the channel.
9.3.2.1.3 Efects of Slope on Facility Staging
The combination of site topography and the constantly increasing size of coal refuse disposal facili-
ties ofen creates design problems not normally encountered with other water-impounding facilities.
For example, the previously mentioned difculties involved in excavating wide auxiliary spillway
channels in steep slopes are multiplied when the design requires that multiple auxiliary spillway
channels be excavated as the height of the embankment is increased in subsequent stages. Problems
V
O
L
U
M
E
O
F
E
X
C
A
V
A
T
I
O
N
CHANNEL WIDTH
W
1
W
2
9.7a EXCAVATION VOLUME VS. CHANNEL WIDTH
FIGURE 9.7 CHANNEL CONSTRUCTION IN MODERATELY AND
STEEPLY SLOPING TERRAIN
W
1
W
2
POTENTIAL OVERFLOW
FAILURE
RESISTANT ROCK
RETARDS EROSION
SOIL OR WEATHERED
ROCK CAN SLOUGH AND
BLOCK CHANNEL
9.7b CHANNEL EROSION AND OVERBURDEN SLOUGHING
FIGURE 9.7 CHANNEL CONSTRUCTION IN MODERATELY AND STEEPLY SLOPING TERRAIN
FIGURE 9.7 CHANNEL CONSTRUCTION IN MODERATELY AND STEEPLY SLOPING TERRAIN
< PREVIOUS VIEW
9-18
Chapter 9
MAY 2009
can also occur when it becomes necessary to tie an embankment into an excavated rock face, as
opposed to the original soil cover on the natural slope. Tying the embankment material into the steep
and broken rock increases the potential for future problems related to seepage, leakage and embank-
ment stability.
For cases where multiple auxiliary spillways are required with construction of succeeding stages,
the designer may wish to consider a series of cascading spillways. A new embankment stage with its
associated spillway channel can be confgured to extend a sufcient distance downstream to allow
the outfow to drop into the spillway channel of the preceding stage with the addition of a plunge
pool. For such confgurations, the hydraulic design of the channels and plunge pool and the erosion
resistance of the rock must be carefully evaluated.
In some cases, the topography may permit an open-channel spillway to be located away from the
embankment, such as through a saddle in a ridgeline, so that the fow is discharged into an adja-
cent watershed, as shown in Figure 9.6. This arrangement can be benefcial in that potential issues
associated with fow escaping from the spillway channel and adversely afecting the downstream
face of the embankment are avoided.
9.8a EXCAVATION VOLUME VS. CHANNEL WIDTH
V
O
L
U
M
E
O
F
E
X
C
A
V
A
T
I
O
N
CHANNEL WIDTH
W
1
W
2
FIGURE 9.8 CHANNEL CONSTRUCTION IN GENTLY SLOPING TERRAIN
9.8b CHANNEL EROSION
ORIGINAL CHANNEL
BOTTOM
ERODED CHANNEL
BOTTOM
W
1
W
2
FIGURE 9.8 CHANNEL CONSTRUCTION IN GENTLY SLOPING TERRAIN
FIGURE 9.8 CHANNEL CONSTRUCTION IN GENTLY SLOPING TERRAIN
< PREVIOUS VIEW
9-19
Hydrology and Hydraulics
MAY 2009
9.3.2.2 Weather and Climate
Weather and climatic conditions should be considered as part of the planning associated with the
design and construction of hydraulic structures. Specifc examples include:
In most coal regions of the United States, construction of channels, ditches, con-
crete spillways and decant systems should be scheduled to the extent possible
during normal construction seasons and should be avoided in winter, when
freezing conditions and snowfall may interrupt construction. Accordingly, the
staging of any disposal facility should be planned so that there is adequate fex-
ibility to allow extensions, replacement or modifcation to hydraulic systems
during favorable weather even though coal refuse is handled and disposed on a
year-round basis.
Many western coal felds are in arid or semi-arid climates where the growth of veg-
etation is a very slow process. In these areas, using vegetation as a means of erosion
protection in excavated channels may not be practical.
9.3.2.3 Geology
Normally, geologic conditions do not change drastically within a small geographic area, and thus
they generally do not directly afect disposal facility site selection alternatives. However, soil and
rock conditions at a site are always important to the design of hydraulic structures and ofen are the
deciding factor in choosing among several hydraulic system alternatives with similar cost and utility
characteristics. The following are important geologic and geotechnical factors that must be consid-
ered in design:
If excavated channels in steep slopes are being considered, the designer should eval-
uate the stability of the cut slopes. If excessive costs will be required to achieve sta-
bility, either by benching or by constructing retaining systems, an alternative system
may be more cost efective.
If an excavated channel is to be located along a hillside, it should have a sufcient
capacity that overtopping or discharge that could cause cascading water to fow
onto a critical portion of the embankment does not occur. If a bend in the channel is
required, the efects of fow, erosion, and water superelevation caused by the change
of direction should be carefully evaluated. The outlet end of a spillway channel
should be located sufciently far downstream that the discharge will not erode the
downstream face of the embankment.
Channels should be designed to resist potential erosion efects, so that post-construc-
tion stabilization is not required.
If hydraulic structures are to be constructed in or over sof soils or sof coal refuse,
the amount of setlement that could occur should be estimated in order to determine
whether special construction will be required. Similar considerations may arise in
situations where diferential setlement may occur, such as where hydraulic struc-
tures are constructed across rock abutments and onto fll materials. This is especially
important for conduits through an embankment. Where possible, such conduits
should be founded on and properly bedded in frm materials that will not setle
signifcantly. Where setlement is unavoidable, the initial slope, camber, joints and
conduit material should be selected such that anticipated setlements can be accom-
modated without damage to the system.
The efect of geologic conditions on runof during storms is discussed in Section 9.6.1.
< PREVIOUS VIEW
9-20
Chapter 9
MAY 2009
9.3.3 Construction Materials
The selection of construction materials for hydraulic structures should account for the following:
The potential for corrosion of construction material is high at many coal refuse
disposal facilities because of the chemical characteristics of water seeping through
refuse materials. Choosing corrosion-resistant materials with higher initial cost
may be far less expensive over the long term than repairing a deteriorated structure
several years afer its original installation, especially if the structure will be buried
under many feet of refuse.
Any conduit or structure beneath or within an embankment should be designed for
the external pressure of the maximum potential height of the embankment above it
and for deformations that may result from embankment construction.
Channel lining material should be selected to be resistant to the maximum antici-
pated fow velocities with provisions for drainage and resistance to uplif pressures.
Filter criteria for all materials used in the embankment and appurtenant structure
construction should be evaluated so that the potential for erosion and piping within
the embankment or loss of structural support and/or failure of the hydraulic convey-
ance structures is minimized.
9.4 DESIGN CONSIDERATIONS FOR DISPOSAL FACILITY EMBANKMENT TYPES
In addition to the general design considerations discussed in Chapter 5, there are specifc hydrologic
and hydraulic design considerations for each type of coal refuse disposal facility embankment. The
following discussion of facility-dependent hydrologic and hydraulic design considerations is a gen-
eral summary of the most common considerations for each type of disposal facility.
Some of the primary hydraulic system functions common to all refuse disposal facilities are listed
below. The type and confguration of the coal refuse disposal facility determines the signifcance of
each function.
Collection of runof from the watershed above the embankment and from the surface
of the embankment.
Control, conveyance and discharge of collected water to a downstream location.
Control of the embankment slope utilizing benches at 50-foot or lower vertical inter-
vals to reduce potential erosion.
Erosion protection of the embankment surface during initial, interim and reclama-
tion stages, especially along the embankment face.
Protection of streams or wetlands from encroachment or other potential environmen-
tal impacts that may require mitigation.
Protection of downstream water quality from sediment-laden runof, leachate from
internal drain collection systems, or collected seepage.
The specifc impact of the above hydrologic and hydraulic design considerations is discussed in the
following sections. While typical fgures are presented to assist in recognizing specifc conditions,
they do not depict all design situations.
9.4.1 Non-Impounding Embankments
Non-impounding embankments are used for the disposal of coarse, combined, and dewatered fne
coal refuse. A non-impounding coal refuse disposal facility is designed such that no fne coal refuse
slurry, process water or direct or indirect runof can accumulate within or upstream of the disposal
< PREVIOUS VIEW
9-21
Hydrology and Hydraulics
MAY 2009
facility limits. General types of non-impounding coal refuse embankments include valley, ridge, side-
hill and heaped flls.
9.4.1.1 Valley-Fill Embankments
Valley-fll refuse embankments, as illustrated in Figure 9.9, are ofen constructed by starting dis-
posal at the upper end of a valley and extending the embankment in stages down the valley in such
a manner that an impoundment is never created. Ofen these types of embankments are located in
large valleys so that large refuse disposal volumes can be placed. The potential runof in such valleys
during a large storm event can be high, and to prevent excessive erosion large diversion channels
may be needed. A key design objective associated with the collection of watershed and embankment
surface runof and the discharge of the collected water at a downstream point is to provide the opti-
mum balance between channel cross section and slope, thereby minimizing the cost associated with
channel erosion protection.
FIGURE 9.9 DRAINAGE CONTROL FOR VALLEY-FILL, NON-IMPOUNDING EMBANKMENT
The most difcult portion of the channel design is along the embankment face at the interface of the
coal refuse and the natural ground surface where the steep slope typically results in high velocities. If
practical, the channels should be extended along the valley wall, within natural soil and rock, beyond
the limits of the coal refuse embankment to discharge beyond the embankment toe. If such an exten-
sion is not practical, it is normally necessary to construct a lined or otherwise protected channel at
the interface of the refuse embankment and valley wall to carry the runof safely to the valley foor.
The long diversion/collection ditches along the crest of the disposal facility should be designed with
a base width and slope that allows use of grass-lined channel sections, if possible.
FIGURE 9.9 VALLEY-FILL, NON-IMPOUNDING EMBANKMENT
DRAINAGE CONTROL
S
L
O
P
E
T
O
D
R
A
I
N
DIVERSION CHANNEL
PERIMETER COLLECTION
CHANNEL
SEDIMENT POND
FIGURE 9.9 DRAINAGE CONTROL FOR VALLEY-FILL, NON-IMPOUNDING EMBANKMENT
< PREVIOUS VIEW
9-22
Chapter 9
MAY 2009
The mixture of runof, leachate and seepage may require treatment prior to discharge to the receiv-
ing waterway. Such treatment could entail construction of sedimentation ponds and also ponds for
chemical treatment. Sufcient area for construction of the sedimentation/treatment ponds should
be allocated. However, these facilities should be located above the level of the 100-year foodplain
associated with the receiving stream and not in a position where they could be afected by normal
stream fows.
9.4.1.2 Side-Hill, Ridge and Heaped Embankments
Side-hill, ridge and heaped non-impounding embankments have design confguration considerations
similar to those for valley-fll embankments. The upstream and perimeter watersheds are generally
smaller than for a valley-fll embankment, but the steepness of the fnal embankment slopes and the
water quality of the runof and seepage result in similar hydrologic and hydraulic design consider-
ations as for a non-impounding, valley-fll embankment. Figure 9.10 shows drainage control for side-
hill and heaped embankments.
Side-hill embankments are usually constructed in stages that extend progressively higher on a natu-
ral slope. Therefore temporary diversion ditches are needed for collecting and diverting runof at
intermediate stages when the embankment is not at full elevation. The channel dimensions, slope
and required erosion protection should be designed to meet the fnal conveyance requirements and
to provide economical erosion protection. The location of the toe of the embankment should lie out-
side the 100-year foodplain limits of nearby streams to minimize any potential encroachments, and
sufcient area should be available for sediment/treatment pond construction.
Ridge embankments are generally in the upper reaches of a watershed and may resemble a side-hill
embankment extending above and over a ridge line. The collection and conveyance of precipitation
falling directly onto the embankment is the primary issue since there is typically litle if any upstream
watershed. This type of facility generally has a limited downstream area available for sedimentation
control and chemical treatment, and the natural ground surface may slope away from the disposal
facility in several directions and potentially into other watersheds. Therefore, multiple sedimentation
ponds and pumping to a common point for treatment may be required.
Heaped embankments are generally located on fat terrain. The collection and conveyance of runof is
primarily related to conveying precipitation that falls directly onto the facility and diverting adjacent
area runof away from the facility. Collection ditches on benches and at the crest are typically gently
sloping, and grass-lining can normally be used as the channel erosion control. Ditches conveying
runof from the crest or benches to the toe of the embankment are steeper than the collection channels
and typically require a more durable lining material such as riprap, concrete or manufactured erosion
protection material. Also, the outlet structure must be sufciently oriented and properly designed to
prevent erosion of the embankment toe. For high embankments, special consideration is required
at the discharge points so that the energy of the high velocity fow is dissipated and/or the fow is
directed away from the embankment in a manner that prevents erosion of the toe.
9.4.2 Slurry Cell Embankments
The hydrologic and hydraulic aspects of slurry cell embankment design must accommodate the volu-
metric sequencing of the slurry cells as well as the collection and conveyance of both runof around
the cells and direct runof that accumulates within the slurry cells. Individual slurry cell design must
meet structural and hydraulic design requirements, and construction must be controlled in such a
manner that the slurry cells do not become a large interconnected impoundment. The slurry cell con-
cept is based on limiting the total capacity of all open cells (and fowable material if present in closed
cells) to a level that is consistent with a low-hazard-potential classifcation for the facility or does not
meet the criteria for a regulated impoundment provided in 30 CFR 77.216.
< PREVIOUS VIEW
9-23
Hydrology and Hydraulics
MAY 2009
FIGURE 9.10 DRAINAGE CONTROL FOR SIDE-HILL AND HEAPED EMBANKMENTS
DIVERSION CHANNEL
PERIMETER COLLECTION
CHANNEL
SEDIMENT POND
9.10a SIDE-HILL EMBANKMENT DRAINAGE CONTROL
9.10b HEAPED EMBANKMENT DRAINAGE CONTROL
PERIMETER COLLECTION
CHANNEL
SEDIMENT POND
SEDIMENT POND
SLO
PE TO
D
R
AIN
CREST/SLOPE
COLLECTION
CHANNELS
FIGURE 9.10 DRAINAGE CONTROL FOR SIDE-HILL AND HEAPED EMBANKMENTS
FIGURE 9.10 DRAINAGE CONTROL FOR SIDE-HILL AND HEAPED EMBANKMENTS
< PREVIOUS VIEW
9-24
Chapter 9
MAY 2009
In order for a slurry cell embankment with multiple cells to not require an approved impoundment
plan in accordance with the criteria in 30 CFR 77.216, each individual cell must not exceed the 20-
acre-feet size criterion. Furthermore, where the failure of one cell can result in the failure of another,
or where slope failure can result in the release of water or slurry from multiple cells, the cumulative
storage capacity of the afected cells must not exceed 20 acre-feet. In situations where multiple cells
are operated or arranged such that they may interact and exceed the 20-acre-feet limit, the embank-
ment should be classifed as impounding and should be designed for the appropriate design storm
based on its hazard classifcation. A critical consideration in determining the hazard classifcation for
an impounding embankment is the fowability of the fne coal refuse. Generally, slurry cells work
most efectively when the depth of fnes in the cells is kept relatively shallow, preferably to fve feet
or less, such that afer dewatering and capping the material is unlikely to be fowable. In instances
where there is concern for draining of the fne coal refuse, the following guidance for assessing the
fowability of fne coal refuse is suggested:
Fine refuse is generally considered fowable for: (1) operating cells with active fne
refuse disposal, (2) non-operating cells containing predominantly saturated, fne
refuse deposits that have not been covered, and (3) covered cells with predomi-
nantly saturated fne refuse deposits characterized as very loose sand or very sof
silt or clay.
Fine refuse is generally considered non-fowable for: (1) non-operating cells with
predominantly unsaturated, fne refuse deposits that have been covered and (2)
covered cells with predominantly saturated fne refuse deposits characterized as
medium dense sand or medium stif silt or clay.
Fine refuse should generally be considered fowable, unless additional testing and
analysis demonstrates that it is non-fowable, for non-operating cells with predom-
inantly saturated fne refuse deposits characterized as loose sand or sof silt or clay.
Michael et al. (2005) in an OSM report prepared a review of the fowability of impounded fne coal
refuse that discusses recent work and ideas in the engineering profession.
The major hydrologic and hydraulic considerations for slurry cells are the collection, conveyance
and discharge of runof within the main diversion and perimeter ditches plus the discharge of
direct runof from individual slurry cells. As ditches are relocated and new cells are constructed
at higher elevations, care should be taken so that the embankment is not advanced vertically to
the extent that its impounding capacity exceeds the disposal plan criteria and afects hazard clas-
sifcation. Special consideration is required at the discharge points to control fow and prevent
erosion of the embankment. The location of the toe of the embankment should lie above the 100-
year foodplain limits of nearby streams in order to minimize the potential for encroachments, and
sufcient area should be available for sediment/treatment pond construction. Figure 9.11 shows
drainage control measures for a typical slurry cell facility.
9.4.3 Slurry Impoundments
The primary hydrologic/hydraulic issue associated with slurry impoundment design is the continu-
ous balancing of coarse coal refuse disposal, fne coal refuse slurry disposal and maintenance of storm
water runof storage/routing capacity. Direct runof at a slurry impoundment is typically controlled
by a decant system or principal spillway, although some disposal facilities also employ an auxiliary
(or emergency) spillway. The operation and performance of these outlet works is integral to fne and
coarse coal refuse disposal and the safe operation of the impoundment. To protect the impounding
embankment from erosion, perimeter runof control structures must also be incorporated into the
design. The location of the toe of the embankment should lie outside the 100-year foodplain limits of
< PREVIOUS VIEW
9-25
Hydrology and Hydraulics
MAY 2009
nearby streams to minimize any potential encroachments, and sufcient area should be available for
sediment/treatment pond construction.
The type of coal refuse disposal facility confguration (e.g., cross-valley, diked or incised impound-
ment) is typically a function of topographic conditions in the vicinity of the coal mine. Frequently,
a decant system and storage are used to control runof and thus minimize costs associated with
other types of outlet structures. However, this requires sufcient embankment materials to achieve
the required storage and may not be feasible for large watersheds. Therefore, some impoundments
with large watersheds have auxiliary (or emergency) spillways in combination with planned storage
capacity and a decant system to control runof from the design storm.
Regardless of the outlet structures chosen for various impoundment development stages, special
consideration must also be given to the conditions that will exist when the site is no longer main-
tained as an impoundment. At that point, the impounding capacity must be eliminated by: (1)
backflling the impoundment (typically with coarse coal refuse), (2) excavating a channel through
the embankment to the level of the backflled stabilized fnes, or (3) a combination of these meth-
ods, which is typically the most efective approach. The approach taken must include measures to
prevent signifcant erosion.
9.4.3.1 Cross-Valley Impoundment
A cross-valley impoundment typically consists of an embankment constructed primarily of coarse
coal refuse that functions as a dam to impound a mixture of setled fne coal refuse, slurry, clarifed
water and runof. The impoundment storage and outfow capacity determine the hydraulic struc-
tures needed for controlling runof.
FIGURE 9.11 SLURRY CELL FACILITY DRAINAGE CONTROL
DIVERSION CHANNEL
PERIMETER COLLECTION
CHANNEL
SEDIMENT POND
S
L
O
P
E
T
O
D
R
A
I
N
FIGURE 9.11 DRAINAGE CONTROL FOR SLURRY-CELL FACILITY
FIGURE 9.11 DRAINAGE CONTROL FOR SLURRY-CELL FACILITY
< PREVIOUS VIEW
9-26
Chapter 9
MAY 2009
The most appropriate method for minimizing the spillway construction efort is to provide a very
large surcharge capacity between the initial pond elevation and the initial embankment crest. A spill-
way can then be constructed at a signifcant height above the initial pond level, providing adequate
surcharge capacity for a long operational period before the hydraulic system must be expanded.
Coarse coal refuse typically provides the material for economically constructing this surcharge capac-
ity. An extension of this approach would be to initially provide for total storage with no requirement
for a spillway (although with this approach there must be provisions for drawing down the reservoir
in response to consecutive or repeated storms).
Regardless of the percentage of runof to be handled through reservoir storage, the design confgura-
tion must always accommodate the continual rise in the normal pool level due to the disposal of fne
coal refuse slurry. Reduction in reservoir storage capacity due to upstream construction pushouts
and stages must also be taken into account. A decant system allows the controlled discharge of sur-
charge runof. It may also be used to evacuate clarifed slurry water. Depending upon the confgura-
tion of the impoundment, an open-channel spillway may be needed to discharge runof from larger
storm events.
To protect the downstream face of the coal refuse embankment from erosion, perimeter runof that
is intercepted by embankment bench guters, road guters and collection and diversion ditches must
be controlled and routed to a sediment/treatment pond. The conveyance structure confguration and
erosion protection should be designed to be appropriate for all stages of development, including
reclamation. Some typical drainage control measures for a cross-valley impoundment are illustrated
in Figure 9.12.
9.4.3.2 Diked Impoundment
Diked impoundments have design constraints similar to those for cross-valley impoundments. If a
facility is completely diked such that there is no upstream watershed, the required impoundment
surcharge capacity is minimized, and the primary factor afecting the impoundment storage capacity
is the production of fne coal refuse and clarifcation of slurry. Typically, a decant system and/or prin-
cipal spillway are adequate for control of runof. If an auxiliary spillway is employed, the channel
section through the embankment requires erosion-resistant linings.
Perimeter ditches and bench guters tend to be of substantial length and should be designed with
sufcient slope to adequately convey runof to sedimentation ponds and to drain efectively without
low areas. Where ditches traverse embankment slopes, they should be provided with erosion-resis-
tant linings. Figure 9.13 shows drainage control measures implemented for a typical diked impound-
ment.
9.4.3.3 Incised Impoundment
Incised impoundments, or ponds, are used for the disposal of fne coal refuse. They are typically small
and ofen used for temporary or emergency disposal. The hydrologic and hydraulic considerations
associated with cross-valley impoundments and diked impoundments are generally not major issues
for incised ponds because of the reduced risk of catastrophic failure. There are three principal design
considerations: (1) an outlet structure to decant or control the release of clarifed process water, (2)
diversion to convey adjacent area runof around the incised pond, and (3) fooding potential, if the
incised pond is located close to or within foodplain limits.
9.4.4 Other Impounding Structures
Coal mining operations generally include sedimentation, treatment and fresh water ponds. The
capacity of each of these structures is a function of the intended use. Sedimentation or treatment
< PREVIOUS VIEW
9-27
Hydrology and Hydraulics
MAY 2009
FIGURE 9.12 CROSS-VALLEY IMPOUNDMENT DRAINAGE CONTROL
9.12b CROSS-VALLEY IMPOUNDMENT WITH DESIGN STORM STORAGE
AND WITHOUT OPEN-CHANNEL SPILLWAY
CROSS SECTION
SURCHARGE CAPACITY
FOR DESIGN STORM
SEDIMENT POND
DECANT INLET
DECANT
OUTLET
PLAN
DECANT INLET DECANT
OUTLET
SEDIMENT
POND
9.12a CROSS-VALLEY IMPOUNDMENT WITH OPEN-CHANNEL SPILLWAY
CROSS SECTION
SEDIMENT POND
DECANT INLET
OPEN-CHANNEL
SPILLWAY
DECANT
OUTLET
PLAN
DECANT
INLET
DECANT
OUTLET
SEDIMENT
POND
OPEN-CHANNEL
SPILLWAY
FIGURE 9.12 DRAINAGE CONTROL FOR CROSS-VALLEY IMPOUNDMENT
FIGURE 9.12 DRAINAGE CONTROL FOR CROSS-VALLEY IMPOUNDMENT
< PREVIOUS VIEW
9-28
Chapter 9
MAY 2009
pond capacity is related to the ability of the structure to remove constituents such as suspended
solids or metals that exceed efuent limitations. Fresh water ponds must have the reservoir capacity
to meet the coal processing and other mining requirements. The size (height and reservoir storage
capacity) and downstream impacts of failure of these structures determines the hazard potential and,
as a consequence, the design criteria.
9.4.4.1 Sedimentation and Treatment Ponds
Sedimentation ponds and treatment ponds are typically located beyond the toe of coal refuse dis-
posal facilities or below mining-disturbed land, so that they can receive gravity infow. The sediment
and setling capacity of these structures is typically specifed in state erosion and sedimentation con-
trol guidelines and efuent limitations. Similarly, treatment pond size is dependent on the ponds
ability to treat/remove and discharge acceptable water quality. Pond principal and auxiliary spillway
structures should be designed to discharge water at a rate consistent with design storm criteria and
state regulatory requirements. A primary consideration is the maximum anticipated runof asso-
ciated with the embankment staging based on watershed size, hydrologic considerations, and the
surcharge storage capacity, which is signifcantly less than the gross impoundment capacity. For sedi-
ment ponds, as storage capacity drops, the principal and auxiliary spillways must be able to handle
increased discharges. For ponds located below coal refuse disposal facilities, pond size is a function
of the size of and outfow from the upstream structure. The infow may be only surface runof from
the face of a coal refuse embankment, but it more typically includes decant water discharges, internal
drain system discharges from the coal refuse disposal facility, and other adjacent area runof.
FIGURE 9.13 DIKED IMPOUNDMENT DRAINAGE CONTROL
SECTION A - A
SURCHARGE CAPACITY
FOR DESIGN STORM
DECANT
PIPE
PLAN
A A
FIGURE 9.13 DRAINAGE CONTROL FOR DIKED IMPOUNDMENT
FIGURE 9.13 DRAINAGE CONTROL FOR DIKED IMPOUNDMENT
< PREVIOUS VIEW
9-29
Hydrology and Hydraulics
MAY 2009
9.4.4.2 Fresh Water Impoundments
Fresh water impoundment capacity is determined by the mine and mine processing plant require-
ments. Fresh water impoundment capacities are generally large, and these impoundments are ofen
regulated as high-hazard-potential structures. Fresh water impoundments should be designed
and constructed according to accepted criteria for conventional dams. Outlet structures for these
impoundments generally include both principal and auxiliary spillways.
9.5 DESIGN STORM CRITERIA
The quantity and distribution of runof during a design storm for a coal refuse disposal facility site
largely controls the design of hydraulic appurtenant structures. This section discusses design storm
criteria in terms of the recurrence interval of the precipitation and the magnitude of precipitation
measured in inches of rainfall. Section 9.6 discusses methods for converting design precipitation to
design runof volume and peak fow rates.
The appropriate design storm for a coal refuse disposal facility depends primarily on the conse-
quences of the uncontrolled release of impounded material due to failure or faulty operation of the
facility. Other factors that may afect the design storm include the facility confguration and size, type
of hydraulic systems and operational period. Portions of the total hydraulic system, such as drainage
culverts, ditches and some diversion channels will not generally create potentially hazardous condi-
tions, so other design criteria can be selected for these structures. This situation is most likely to occur
at non-impounding disposal facilities and at the perimeter of and appurtenant structures associated
with impounding facilities.
Criteria for selecting a design storm for the operational period of an impounding facility are presented
in Section 9.5.1. Design storms that are applicable for short-term conditions are discussed in Section 9.5.2.
Design storm criteria for minor site drainage conveyance structures are presented in Section 9.5.3.
9.5.1 Design Storms for Impoundments
9.5.1.1 General Considerations
Numerous design storm criteria are employed in hydrologic analyses for water retention and food
control dams. The common factor associated with practically all of these criteria is that diferentiations
are made based on the projected maximum size of the impoundment and the magnitude of potential
downstream hazard in the event of failure. MSHA has developed guidelines for design storms for the
impoundments and embankments that they regulate; however, state and local criteria must also be
considered. For any impoundment, the most conservative of applicable criteria should be used.
As part of the identifcation of the design storm, the size of the dam and reservoir and the associated
hazard potential is typically determined either by inspection or analysis. Table 9.4 indicates appropriate
design storms as related to impoundment size and hazard potential. Coal refuse impoundments should
be designed for the Probable Maximum Flood (PMF) event, unless a lesser criterion can be justifed
consistent with Table 9.4. For determining the impoundment size, the impoundment volume and depth
should include all water, sediment, and slurry that can be impounded. For determining the hazard
potential, both the water and fowable materials retained in the impoundment should be considered.
The PMF is defned as the maximum runof condition resulting from the most severe combination of
hydrologic and meteorological conditions that is considered reasonably possible for the watershed. A
PMF consists of an antecedent storm, a principal storm and a subsequent storm. The current assumed
conditions for a PMF design storm in the MSHA guidelines are the following (MSHA, 2007):
1. Antecedent storm 100-year precipitation event, with antecedent moisture condition
II (AMC II) occurring 5 days prior to principal storm.
< PREVIOUS VIEW
9-30 REV. AUG.
Chapter 9
2010
TABLE 9.4 RECOMMENDED MINIMUM DESIGN STORM CRITERIA FOR
COAL REFUSE DISPOSAL IMPOUNDMENTS
A. Impoundment Size Classifcation
Category
Impoundment Size
Maximum Volume of Stored Water
During Design Storm
(acre-ft)
Maximum Depth of Water
During Design Storm
(ft)
Small to Intermediate < 1,000 and < 40
Large 1,000 or 40
B. Hazard Potential Classifcation
Category Description
Low Hazard
Potential
Facilities where failure results in no probable loss of human life and low economic and/or environmental
losses. Such facilities would be located in rural or agricultural areas where losses would be limited
principally to the owners property, or failure would cause only slight damage, such as to farm
buildings, forest, and agricultural land, or minor roads.
Signifcant
Hazard
Potential
Facilities where failure results in no probable loss of human life but can cause economic loss,
environmental damage, or disruption of lifeline facilities. Such facilities would often be located in
predominantly rural areas, but could be located in areas with population and signifcant infrastructures,
and where failure may damage isolated homes, main highways, minor railroads or disrupt the use of
service of public utilities.
High Hazard
Potential
Facilities where failure will probably cause loss of life. Such facilities would be located where
failure could be reasonably expected to cause loss of life, serious damage to homes, industrial and
commercial buildings, important utilities, highways and railroads.
C. Recommended Design Storm for Long-Term and Short-Term Conditions
(1)
Impoundment
Size
Hazard
Potential
Minimum
Design Storm
for Long Term
Minimum Design
Storm for Short
Term
(1)
Additional Criterion
Small to
Intermediate
Low
Signifcant
High
100-Year
-PMF
PMF
100-Year
100-Year
-PMF
The indicated storm is appropriate only
if the combination of spillways and
decants for the facility can evacuate 90
percent of the incre mental volume of
stored storm water within 10 days.
Large
Low
Signifcant
High
-PMF
PMF
PMF
100 Year
-PMF
-PMF
Note: 1. Situations where short-term criteria may apply include:
a. Initial construction. A new impoundment should be capable of accommodating the runoff from
the short-term storm within one year and the long-term storm within two years.
b. Changing from an open-channel spillway to handle the design storm by storage. The time period
when the long-term design storm cannot be accommodated should be kept as short as possible
with detailed planning of the process.
c. Abandonment by elimination of impounding capacity. The impounding capability should be
eliminated within two years after the impoundment can no longer accommodate the long-term
design storm, and the work should be phased so that the facility is capable of accommodating
less than the short-term storm for no more than one year.
< PREVIOUS VIEW
9-31
Hydrology and Hydraulics
MAY 2009
2. Principal storm Probable Maximum Precipitation (PMP) with AMC III. The princi-
pal storm rainfall must be distributed spatially and temporally to produce the most
severe conditions with respect to impoundment freeboard and spillway discharge.
3. Subsequent storm The subsequent storm criterion can be considered to be met if,
within 10 days of the peak impoundment level associated with the principal storm,
at least 90 percent of the volume of water stored above the normal operating level
can be discharged from the impoundment. Alternatively, for facilities designed with
sufcient storage but limited discharge capabilities that do not meet this criterion,
the subsequent storm may be a second PMP storm with the same hydrologic and
meteorological parameters as the principal storm, provided that the storage from
both storms is drawn down at a rate sufcient to evacuate 90 percent of one storm
from the impoundment within 30 days.
The antecedent storm precipitation can be obtained from National Weather Service publications. The
most current defnition of PMP (NWS, 1988) is theoretically, the greatest depth of precipitation for
a given duration that is physically possible over a given size storm area at a particular geographical
location at a certain time of the year. The PMP can be determined from the National Weather Service
publications discussed in the following paragraphs.
In the Western U.S., determination of the PMF may be based upon either: (1) the PMP and (2) the
Probable Maximum Thunderstorm (PMTS). The PMTS is a very high-intensity, short-duration storm
with intense precipitation occurring during a one-hour period. When designing a coal refuse dis-
posal facility in this region, the more critical of these two criteria should be used. In this Manual, the
term PMP represents the more severe of the PMP and PMTS for areas of the U.S. west of the 105
th
meridian.
Dams or impoundments used for fne coal refuse disposal, fresh water retention, erosion and sedi-
ment control or other mine-related operations may need to have PMF storage/routing capacity. Less
critical impoundments may have reduced design storm criteria based on embankment size and the
potential downstream hazard. For such structures, both the 6-hour and 24-hour precipitation inten-
sity (unless criteria are specifed by state regulations) should be evaluated and the more conservative
used for design.
As with water-impounding dams, basic design storm criteria apply to the long-term operation of coal
refuse disposal facilities. However, short-term criteria, as summarized in Table 9.4, may be used for
construction periods that typically extend from several months to two years for impounding struc-
tures subject to PMF design storm criteria. The designer of coal refuse disposal facilities must take
into account that the confguration of the impounding embankment will be continually changing
as additional refuse is placed and that the time associated with any one phase or the time between
phases may be quite short. This can be accounted for by additional or modifed design storm criteria
presented in Section 9.6. These modifed criteria should only be used for unavoidable situations
that occur: (1) during short-term operations associated with initial construction of a disposal facility,
(2) when a major modifcation is being made to an existing disposal facility, and (3) when a refuse
disposal facility is being prepared for abandonment.
For water-retaining impoundments, diferent design storms are sometimes used for individual por-
tions of the total hydraulic system such as the principal spillway and auxiliary spillway (NRCS,
2005b). This practice is generally not followed in the design of coal refuse disposal facilities provided
the overall hazard criteria are satisfed because of the operational characteristics of a disposal facility,
the dynamic nature of facility growth and the limited operational period. This practice may be appli-
cable to other impounding facilities that support the mining operations (e.g., fresh water impound-
ment, sediment ponds, etc.).
< PREVIOUS VIEW
9-32
Chapter 9
MAY 2009
In the design of coal refuse disposal facilities, it is important to diferentiate between the functions of
spillways and decants. The main function of a decant system is to discharge clarifed water from the
impoundment afer the fne refuse has setled. Under normal precipitation conditions, the elevation
of the decant inlet controls the normal operational water level in the impoundment. The capacity of
a decant is limited and is typically too small to signifcantly afect the peak outfow during a large
storm. Therefore, the storm runof is almost totally controlled by impoundment storage or a combi-
nation of impoundment storage and auxiliary spillway capacity.
Even though an impoundment decant system does not have a signifcant impact on the outfow during
the design storm, its capacity must be considered in other analyses related to storms. If the auxiliary
spillway level is above the normal impoundment level (the typical condition) or if the hydraulic
system design relies entirely on storage (no auxiliary spillway), the excess storm runof must either
be discharged totally through the decant system, or the decant system must serve as the primary
outlet until the spillway level is reached. As indicated in Table 9.4, within ten days, the combined
capacity of the spillway and decant systems must be capable of removing 90 percent of the maximum
volume of water stored above the allowable normal operating water level during the design storm.
The 10-day drawdown criterion begins at the time the water surface reaches the maximum elevation
associated with the design storm. Alternatively, if there is sufcient impoundment capacity to store
the runof from two design storms (specifcally, the antecedent storm and two principal storms), an
extension of the 10-day criterion is reasonable, provided that an efective means for discharging the
storage from both storms is available. Generally, an evacuation rate that will remove 90 percent of the
stored runof from one design storm within 30 days is considered to be reasonable.
9.5.1.2 Recommended Design Storm Criteria
Table 9.4 provides recommended minimum design storm criteria for coal-mining-related impound-
ing facilities for both long-term and short-term conditions. Selection of the appropriate storm for a
specifc impounding structure is based on the impoundment size and hazard-potential classifcation.
The selected criteria for the storage and routing of the design storm and hydraulic structure design
should also refect any other applicable regulatory reviewing agency criteria.
Dams and impoundments that are small to intermediate in size (less than 40 feet in height or 1,000
acre feet in storage volume) with low hazard potential should be designed for a long-term storm
event with no less than a 100-year recurrence interval. For coal refuse impoundments equal to or
greater than 40 feet in height or 1,000 acre feet in storage volume with low or signifcant hazard
potential, the minimum long-term design storm should be either the PMF or full PMF, respectively.
The PMF design storm should have one-half of the infow rate and runof volume of the full PMF.
For coal refuse impoundments with high hazard potential, the minimum long-term design storm
should be the full PMF. In cases where the design storm for long-term conditions is less than the full
PMF, it may be prudent to adopt minimum design storm criteria greater than those provided in Table
9.4 and thus achieve greater protection from food events and related damage.
The following paragraphs discuss the basis and/or justifcations for criteria and information pre-
sented in Table 9.4. Procedures for quantitatively determining the magnitude of precipitation to be
used in the calculation of runof are discussed in Section 9.6.
9.5.1.3 Size and Hazard-Potential Classifcation
The rationale for relating the design storm to the size and hazard potential of the disposal facility
impoundment is evident. Impoundment size is defned by the maximum depth and total volume
of retained water, sediment and slurry; however, determining the hazard-potential classifcation
requires judgment and, unless otherwise obvious, should be based upon hydraulic analyses. The
bases for the criteria listed in Table 9.4 are discussed in the following subsections.
< PREVIOUS VIEW
9-33
Hydrology and Hydraulics
MAY 2009
9.5.1.3.1 Impoundment-Size Classifcation
The size classifcation presented in Table 9.4 is based on the total volume and depth of all water, sedi-
ment and slurry impounded during the design storm. As indicated in the table, the recommended
design storms for small and intermediate size impoundments are the same.
9.5.1.3.2 Hazard-Potential Classifcation
The hazard-potential classifcation presented in Table 9.4 is the same as that presented in Chapter 3
and used in the overall classifcation system for coal refuse disposal facilities. Dams that are located
where loss of life is probable in the event of failure are classifed as having high hazard potential.
In applying these criteria, it is important to recognize the difculty of determining whether minor
or major damage or the loss of life will result from the failure of a refuse disposal facility. For most
coal refuse disposal facilities, this determination is based upon: (1) the confguration and location
of the facility and (2) the downstream conditions (both existing and planned) including popula-
tion, topography and the size of streams that would receive food fow resulting from an embank-
ment failure or a breakthrough-type release from the impoundment. Downstream conditions are
typically evaluated by reviewing USGS topographic quadrangle maps and by feld verifcation.
The manner that MSHA addresses the hazard associated with a breakthrough-type release is dis-
cussed in Section 3.1.
Generally, unless it is otherwise evident, the determination of hazard potential is based upon a dam
or impoundment breach analysis and inundation mapping. Section 9.9 presents dam-breach-analy-
sis methods. A dam-breach analysis should provide inundation levels for two conditions: (1) postu-
lated failure of the dam under design-storm conditions and (2) postulated failure of the dam during
normal operations (sunny day or fair weather breach failure). If doubt exists as to the possible efects
of an impoundment failure on downstream areas, the more conservative hazard classifcation should
be selected. However, it may also be useful to evaluate the downstream inundation and damage that
could result from a major storm in the refuse disposal facility watershed, but without failure of the
impoundment. This inundation level and related damage can then be compared to the incremental
inundation and damage that would be caused by failure of the disposal facility under design-storm
conditions. If the additional damage can be reasonably predicted as small, then a less conservative
design storm may be appropriate (FEMA, 2004a), or the hazard-potential classifcation may be gov-
erned by the fair weather breach.
For most large dams and impoundments where downstream residential, commercial or industrial
development is present adjacent to streams, a high-hazard-potential classifcation is selected based
on probable loss of human life. Other situations can arise where the threat is less evident or where
the distinction between signifcant and low hazard potential is important. FEMA (2004a) provides
guidance for interpreting the probable loss of life by clarifying that postulating every conceivable
circumstance that might remotely place a person in the inundation zone should not be the basis
for determining the appropriate classifcation level. In the defnition of high hazard potential, the
probable loss of human life is clarifed to exclude consideration of the casual user of downstream
or upstream areas. However, personnel who routinely or frequently work or occupy locations or
structures in the downstream area should be considered in the assessment of hazard-potential clas-
sifcation.
USBR (1988) provides guidance based upon the number of lives in jeopardy (all individuals within the
inundation boundaries who, if they took no action to evacuate, would be subject to danger) to aid in
assessing the potential for probable loss of life. In cases where a dam-breach analysis indicates limited
inundation at occupied structures in relatively undeveloped areas, such guidance in assigning hazard
potential may be useful. USBR (1988) provides guidelines for interpreting the signifcance of predicted
inundation depth and velocity at downstream residences, roadways, and pedestrian routes.
< PREVIOUS VIEW
9-34 REV. AUG.
Chapter 9
2010
There have been a limited number of mining situations, primarily in the western United States, where
high-hazard-potential dams have been designed using hydrologic design criteria associated with a
lower hazard-potential-classifcation and incorporating a warning system. An example is a dam con-
structed across a watercourse for prevention or mitigation of fooding damage to a surface mine pit.
To design a food-control structure to totally accommodate the design event would necessitate the
construction of a very large dam that would function only temporarily. Failure of this dam could pos-
sibly result in a higher hazard potential due to the additional storage. In such cases, some designers
have proposed dams using low- or signifcant-hazard-potential criteria and incorporating warning
systems. The warning systems are designed to notify the mining operation when the water behind
the dam reaches a specifed level. At that time, all potentially afected personnel are withdrawn from
the downstream area. Allowance for warning time must not be a substitute for appropriate dam
design and construction. MSHA (Fredland, 2008) has indicated that this approach may be acceptable
on a case-by-case basis for temporary mining operations. Conditions associated with warning sys-
tems for this approach are discussed in Section 3.7.
Hazard-potential classifcation is also dependent on the potential for economic, environmental or
lifeline losses. If a dam or impoundment is not classifed as having high hazard potential because
there is no probable loss of human life, generally it refects a situation where there are few down-
stream structures and thus limited potential for associated economic damages. FEMA (2004a) clari-
fes that for classifcation of a dam as having low hazard potential (as opposed to signifcant hazard
potential), the economic, environmental or lifeline losses must be low and generally limited to the
owner of the structure. While economic damages to downstream development may be determined
to be low and thus could support classifcation of a dam as having low hazard potential, the possi-
bility of environmental damages may warrant consideration of higher hazard classifcation levels.
9.5.1.4 Determination of Design Storm Precipitation
Once the size and hazard-potential classifcation of a disposal facility impoundment are established,
the recommended design storm can be determined from Table 9.4. The procedure for determining the
magnitude of the precipitation for the design storm is discussed in the following paragraphs, while
the procedure for computing the resulting runof is presented in Section 9.6.
9.5.1.4.1 Prediction of the PMP and PMTS
Predictions of the PMP (inches of rainfall) for a watershed of 10 square miles and durations of 6 to
72 hours are presented in reports prepared by the National Weather Services Hydrometeorological
Design Studies Center. Figure 9.14 identifes applicable Hydrometeorological Reports (HMRs) for
various regions of the U.S. For areas east of the 105
th
meridian, HMR 51 (NWS, 1978) should be used
for determining the PMP magnitude and for extending the PMP to longer durations. The only excep-
tion is an area in the Tennessee River Valley that is addressed in HMR 56 (NWS, 1986). Procedures for
determining critical rainfall spatial and temporal distribution for areas east of the 105
th
meridian are
provided in HMR 52 (NWS, 1982); however, the document may not be applicable to all watersheds,
particularly watersheds with areas less than 10 square miles. Seasonal variation of PMP for areas east
of the 105
th
meridian is addressed in HMR 53 (NWS, 1980). For the region between the 103
rd
and 105
th
meridian, HMR 55A (NWS, 1988) and HMR 52 should be used. For the area between the 103
rd
merid-
ian and the continental divide, HMR 55A is applicable. For areas west of the continental divide, HMR
49 (NWS, 1977), HMR 57 (NWS, 1994), HMR 58 (NWS, 1998), or HMR 59 (NWS, 1999) should be used,
as indicated by the shaded areas in Figure 9.14.
As indicated above, HMR 56 (NWS, 1986) was developed for the Tennessee Valley. While HMR 56 is
recommended for projects in that region, the study indicates that in non-orographic areas numerous
comparisons were made between the results from HMR 56 and the results from HMR 51 and HMR
52, indicating that minor diferences in results can be expected depending upon the size of the study
< PREVIOUS VIEW
9-35
Hydrology and Hydraulics
MAY 2009
region. In more mountainous orographic areas, HMR 56 provides guidance for determining the areal
distribution of storm-averaged depths with reference to HMR 52.
The extension of the PMP for watersheds exceeding 10 square miles and other durations are dis-
cussed in relation to analyses for determining runof in Section 9.6. For coal refuse disposal impound-
ments, the applicable watershed is typically much smaller than 10 square miles, resulting in no or
only limited adjustments for spatial distribution using HMR 52 (NWS, 1982). Because impoundments
are designed with considerable storage capacity, and in many cases the ability to store the runof from
the entire design storm, determination of adjustments that afect the peak infow rate may not be nec-
FIGURE 9.14 U.S. REGIONS COVERED BY GENERALIZED PMP STUDIES
essary. However, in such cases the PMP must generally be extended to 72 hours. For impoundments
with watershed areas as small as one square mile that rely on open-channel spillways for routing
the PMF, HMR 52 (NWS, 1982) provides a means for estimating the adjusted PMP distribution using
depth-duration ratios and 1-hour PMP values.
9.5.1.4.2 Prediction of the 100-year and Lesser Design Storms
The 100-year-recurrence-interval design storm and lesser design storm precipitation data can be obtained
from several sources. NOAA Atlas 14 Volume 1 (NWS, 2006a) and Volume 2 (NWS, 2006b) provide rain-
FIGURE 9.14 U.S. REGIONS COVERED BY GENERALIZED PMP STUDIES
(ADAPTED FROM FERC, 2001)
NOTE: 1. HMR 57 REPLACED HMR 43 IN 1994.
2. HMR 58 AND 59 REPLACED HMR 36 AND 49 FOR CALIFORNIA IN 1999.
3. HMR 51, 52, 53 MAY APPLY FOR BASINS UNDER 24,000 SQUARE MILES.
4. HMR 33 (SEASONAL PMP VARIATIONS FOR AREAS NOT SUPERCEDED
BY SITE-SPECIFIC SEASONAL PMP REPORTS)
HMR 57
(1)
HMR 49
HMR 55A
HMR 51, 52, 53
(4)
103RD
MERIDIAN
105TH
MERIDIAN
HMR 48 (SNOWMELT CRITERIA)
HMR 40
(3)
HMR 42, 54
TP 47
HMR 58, 59
(2)
CONTINENTAL
DIVIDE
EPRI, 1993
PMP WI/MI
HMR 56
TENNESSEE
VALLEY REGION
HMR 52, 55A
FIGURE 9.14 U.S. REGIONS COVERED BY GENERALIZED PMP STUDIES
< PREVIOUS VIEW
9-36
Chapter 9
MAY 2009
fall frequency values for much of the U.S., and NOAA Atlas 2 (NWS, 1973) provides data for some areas
of the western U.S not covered in Atlas 14 Volume 2. Other areas of the country are addressed in various
technical publications available from NOAA or in other sources listed in Table 9.5. Access to precipitation,
frequency, and intensity data for specifc locations is available from the NOAA web site.
The current practice of precipitation frequency analysis is based upon the implicit assumption that
past experience can be used to predict future events and that the climate will not change. In its cur-
rent studies, the NWS is assuming that the full period of the available historical record is suitable
for use, as current climate change forecasts do not reliably defne future changes in precipitation
frequency distribution.
9.5.2 Special Considerations for Short-Term Conditions
Although coal refuse disposal facilities are typically dynamic or constantly changing entities, careful
consideration of growth characteristics and proper planning of modifcations will result in compli-
ance with long-term design storm criteria over the facilitys entire service life. Occasionally, however,
it may be impossible to meet long-term requirements during brief periods when signifcant physical
changes to the facility are occurring.
Appropriate design storms for short-term conditions are provided in Table 9.4. Short-term conditions for
periods of signifcant physical change are more related to general construction practices, and therefore the
criteria for temporary (stream) diversions are generally dictated by state guidelines. The upper limit for a
short-term condition is two years. The short-term design storms provided in Table 9.4 are more conserva-
tive (higher) than those normally used for dam construction (USBR 1987a). The more stringent criteria are
recommended because planning and implementation of modifcations at coal refuse disposal facilities are
dependent upon day-to-day coal and refuse generation unlike other types of embankments.
It is stressed that these short-term criteria are not intended as less costly design alternatives based on
the rationale that a short-term condition is always appropriate for a given site because it is continu-
ally changing in confguration. If such an approach is followed, it should be expected that regulatory
acceptance of a lesser storm will not be granted. The temporary use of design storms of lesser magni-
tude than those required for long-term facility operation will likely be accepted only if the following
conditions are met:
The facility will be designed to satisfactorily meet the requirements for such interim
use, including, but not limited, to safe control of the short-term design storm.
As part of the overall design and planning process, interim periods of short-term
use are unavoidable and are identifed and their duration realistically scheduled. As
these periods are approached during construction, the scheduling of these transi-
tional periods should be adjusted as required and thereafer strictly followed. Such
preplanning and scheduling should be done in a manner that minimizes the dura-
tion of the short-term condition and facilitates the speedy transition to either a long-
term operating status or abandonment.
Periods during the service life of a refuse disposal facility when even careful planning may occasion-
ally be insufcient to achieve compliance with long-term design criteria include:
Initial construction of a new impounding structure. The impoundment should be
capable of accommodating the runof from the short-term storm within one year and
the long-term storm within two years.
Transitioning from a lower open-channel spillway to a higher open-channel spill-way
as part of raising the embankment stage crest or changing from an open-channel spill-
< PREVIOUS VIEW
9-37
Hydrology and Hydraulics
REV. AUG. 2010
way to handling the design storm by storage. The time period when the long-term
design storm cannot be accommodated should be kept as short as possible, and a com-
prehensive plan and schedule for the sequence of the change should be provided.
Abandonment by elimination of impounding capability. The impounding capability
of the facility should be eliminated within 2 years afer the time that the impound-
ment can no longer accom modate the long-term design storm. Additionally, aban-
donment should be phased such that the time period when the facility is capable of
handling less than the short-term storm is no more than one year.
TABLE 9.5 NWS PRECIPITATION FREQUENCY PUBLICATIONS
Location
Design Storm Duration
5 to 60 min 1 to 24 hrs 2 to 10 days
DE, IL, IN, KY, MD, NJ,
NC, OH, PA, SC, TN, VA,
DC, WV
NOAA Atlas 14
Volume 2 (NWS, 2006b)
NOAA Atlas 14
Volume 2 (NWS, 2006b)
NOAA Atlas 14
Volume 2 (NWS, 2006b)
Remainder of Eastern
United States and TX
Technical Memorandum
NWS HYDRO-35
(Frederick et al.,1977)
Technical Paper 40
(Hershfeld, 1961)
Technical Paper 49
(NWS, 1964)
AZ, NV, NM, UT,
Southeast CA
NOAA Atlas 14
Volume 1 (NWS, 2006a)
NOAA Atlas 14
Volume 1 (NWS, 2006a)
NOAA Atlas 14
Volume 1 (NWS, 2006a)
Remainder of Western
United States
Arkell and Richards (1986);
Frederick and Miller (1979)
NOAA Atlas 2
(NWS, 1973)
Technical Paper 49
(NWS, 1964)
Alaska
Technical Paper 47
(NWS, 1963)
Technical Paper 47
(NWS, 1963)
Technical Paper 52
(NWS, 1965)
(ADAPTED FROM NRCS, 1986)
9.5.3 Hydraulic Design Criteria for Drainage Conveyance Installations
Hydraulic structures for both non-impounding and impounding coal refuse disposal facilities fall
into three general categories:
1. Those structures that by failure, overtopping and/or blockage could threaten the
overall stability of the disposal facility.
2. Those structures that by failure, overtopping and/or blockage would not threaten the
overall stability of the disposal facility, but could lead to localized instability.
3. Those structures that, even if non-functional, would not endanger the overall stabil-
ity of the facility and would not greatly afect day-to-day operation of the facility.
Hydraulic structures that are critical to the overall safety of coal refuse disposal facilities must be
designed to adequately control the facility design storm. Although this most commonly applies
to the hydraulic structures associated with impounding facilities, the requirement applies equally
to hydraulic structures at non-impounding facilities. Permanent hydraulic structures (other than
impoundment spillways) at coal refuse disposal embankments should be designed to handle the
100-year storm.
The purpose of many refuse disposal facility permanent hydraulic structures is to limit erosion or
other types of localized instability rather than to provide total hydraulic control during major storms.
Whether or not a facility is impounding, non-impounding, active or abandoned, these structures are
< PREVIOUS VIEW
9-38
Chapter 9
MAY 2009
important to the development and operation of a disposal facility and should generally be designed
for the 100-year storm. Design criteria for these structures may also be governed by state or local
regulations. Table 9.6 provides a summary of typical design criteria for minor hydraulic structures at
locations that are not part of the coal refuse disposal facility. These structures typically include storm
sewers, culverts, drainage ditches and guters.
9.6 DETERMINATION OF RUNOFF QUANTITIES
The most important aspects of hydrologic analyses related to refuse disposal facility performance
during and afer storm rainfalls are the determination of the peak runof rate and the total runof
volume at the point of interest. Four methods for determining these parameters that are available to
the designer are presented in Table 9.7. The frst three methods presented in the table are discussed
in this section following a general discussion of basic hydrology parameters.
9.6.1 Basic Hydrology Parameters
There are three basic factors that must be considered when predicting runof rates and quantities.
These are: (1) precipitation (intensity and duration), (2) watershed (size and time of concentration),
and (3) soil types and land use conditions. These factors are further explained in the following
subsections.
9.6.1.1 Precipitation Intensity-Duration and Distribution
Storms are defned by their precipitation intensity-duration relationships. Storms can range from
high-intensity, short-duration thunderstorms to low-intensity, long-duration rainfalls lasting several
days. The intensity-duration relationship that should be used for hydrologic analyses and channel
design is that which produces the maximum peak runof rate. This is particularly true for the small
watersheds common to coal refuse facilities where the time of concentration (time required for rain-
fall to travel from the most hydrologically distant point in the watershed to the point of interest) is
TABLE 9.6 TYPICAL DESIGN CRITERIA FOR MINOR HYDRAULIC STRUCTURES
(1)
Structure Type and Condition Design Criteria
Storm Sewers 10-year rainfall
Diversion Systems
Temporary (1-year life or less and watershed > 5 acres)
Construction areas, roads, pipelines 2-year rainfall
Permanent
Sediment Retention Structures (watershed <100 acres and
height < 15 feet):
10-year rainfall
Emergency spillway capacity 25-year, 24-hour rainfall
Principal spillway capacity 10-year, 24-hour rainfall
Culverts:
Access Roads and Drainage Swales 10-year, 24-hour rainfall
Local and Urban Roads 25-year, 24-hour rainfall
Highways and Streams 100-year, 24-hour rainfall
Drainage Ditches and Gutters 10-year rainfall
Note: 1. These criteria do not apply to minor structures on coal refuse disposal facilities. Permanent
perimeter ditches and bench gutters on coal refuse disposal facilities should be designed for
the 100-year storm.
< PREVIOUS VIEW
9-39
Hydrology and Hydraulics
MAY 2009
small. In larger watersheds, the averaging efects of short and long times of concentration tend to
compensate for small errors in the predicted intensity-duration relationship.
Figure 9.15 presents the dimensionless design storm distribution frequently used to evaluate the
intensity-duration relationship (NRCS, 2005b) based upon a 6-hour-duration storm. The 24-hour
storm can be constructed by critically stacking incremental rainfall amounts for successive 6-, 12-,
and 24-hour durations, as discussed in HMR 52. Runof determinations for mining facilities are typi-
cally based upon either 6-hour- or 24-hour-duration precipitation events (with extension to periods
up to 72 hours for impoundments that rely on storage for food routing) except for states west of the
105
th
Meridian where PMTS runof must also be considered.
As indicated in Table 9.4, development of specifc-frequency food hydrographs may be required for
the design of structures with low- or signifcant-hazard-potential classifcation. High-hazard-poten-
tial structures require design for the PMF, which for coal refuse disposal facilities is typically derived
from the PMP for the watershed. The PMP is a 6- to 72-hour duration precipitation distribution that
results in a peak intensity occurring during the third quadrant of precipitation. This distribution
curve is recommended for most coal refuse impoundment hydrologic design and analysis applica-
tions. For smaller impounding structures, it is recommended that both short-duration (6-hour) and
long-duration (24-hour) storms be utilized to determine the peak runof for sizing of the outfow
structures.
In addition to knowing how the intensity of precipitation may be distributed within a six-hour storm,
it is also important to recognize that storms may continue for longer periods of time at decreasing
intensities. Such storms may be critical for disposal facilities that rely primarily on reservoir storage
to control runof, since the amount of runof occurring afer the frst six hours may represent a sig-
nifcant portion of the runof volume of the total storm. HMR 51 and HMR 52 can be used to extend
the predicted six-hour PMP.
Figure 9.14 shows applicable HMRs for determining magnitudes and temporal distributions for
probable maximum storms based upon regionalized criteria. Charts for PMP values are presented
in HMR 51 for most areas east of the 105
th
Meridian, and procedures are provided in HMR 52
TABLE 9.7 METHODS FOR DETERMINING RUNOFF RATE AND VOLUME
Method Applicable Conditions
Hydrograph Method (Section 9.6.2) Applicable to all runoff analyses, but normally used when a time-
related runoff distribution is required or when less exact methods
for estimating runoff are not suffciently accurate for design of an
economical drainage system.
Peak Runoff Determination (Section 9.6.3) For (1) determining estimates of peak runoff rate and runoff
volume for system sizing when time-related runoff distribution is
not required for fnal design or (2) preliminary system sizing prior
to generating a runoff hydrograph for food routing.
Rational Method (Section 9.6.4) For designing drainage conveyance structures such as diversion
and collection ditches and road culverts for small watersheds.
Stream Gage Data Analysis (USDA, 1972;
Chow, 1964)
For predicting runoff by statistical analysis of measured stream
fow records when a long history of data is available for the stream
or for a nearby similar stream and watershed. Since these data are
not generally available for the types of streams passing through or
adjacent to coal refuse disposal facilities, methods using stream
fow records are not presented herein.
< PREVIOUS VIEW
9-40
Chapter 9
MAY 2009
that translate these values to a spatially and temporally distributed estimate of the site PMP. The
computer program HMR 52 developed by the U.S. Army Corps of Engineers (USACE, 1984b)
determines the most severe storm conditions considering basin characteristics and regional condi-
tions that are critical for watersheds with areas greater than 10 square miles. For many coal refuse
impoundments, watersheds are small (typically less than 1 square mile) and the procedures in the
computer program HMR 52 may need to be adjusted for these smaller watersheds using methods
presented in NWS (1982).
HMR 56 is applicable in the Tennessee Valley region. For the region between the 105
th
Meridian and
the Continental Divide, HMR 55A should be used. Probable maximum storm estimates for areas west
of the Continental Divide may be developed using HMR 49, HMR 57, and HMR 59, which account for
orographic efects and include procedures for evaluating local (thunderstorm) PMP storms.
Short-term design storm criteria and low-hazard-potential dam design criteria require precipitation
frequency information that is available from NOAA, as indicated in Table 9.5.
9.6.1.2 Unit Hydrographs and Time of Concentration
Unit hydrograph theory is the basis for computing infow hydrographs for design storms. A unit
hydrograph can be derived from observed hydrographs recorded on gauged streams, although for
most coal refuse disposal facilities located in small watersheds, they are synthesized using relation-
ships between rainfall and runof that are dependent on watershed conditions. Empirical equations
are typically employed to estimate parameters for synthetic unit hydrographs, although some gov-
ernment agencies can provide parameters for ungaged stream basins, including:
FIGURE 9.15 DIMENSIONLESS DESIGN STORM DISTRIBUTION
FIGURE 9.15 DIMENSIONLESS DESIGN STORM DISTRIBUTION
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
DURATION
A
C
C
U
M
U
L
A
T
I
O
N
(NRCS, 2005b)
FIGURE 9.15 DIMENSIONLESS DESIGN STORM DISTRIBUTION
< PREVIOUS VIEW
9-41
Hydrology and Hydraulics
MAY 2009
USACE has developed coefcients for use in computing Snyder and Clark unit
hydrographs for some areas of the U.S. USACE district ofces can provide informa-
tion on the results of studies in their districts.
The USBR has developed a set of lag-time equations, dimensionless unit hydro-
graphs, and S-graphs for diferent parts of the western U.S. (Cudworth, 1989).
The USGS has performed regional studies for development of unit hydrographs
in cooperation with state departments of transportation. These are published as
USGS water resources investigation reports. Some of these are applicable to the
states of Illinois, Tennessee, and Alabama (Graf et al., 1982; Robbins, 1986; Olin
and Akins, 1988).
Before applying published parameters for watersheds in a region, the possible efects of diferences
in drainage area, cover, soil type, orientation, or geology should be evaluated. Additionally, the ter-
minology used to defne the various hydrographs and basin parameters in a regional study should be
carefully reviewed so that application to ungaged watersheds is consistent (e.g., lag time and channel
slope may be defned diferently in the various methodologies).
If published parameters for watersheds in the region are not available, and the drainage basin is
larger than about 100 square miles, a regional analysis may be prepared by analyzing rainfall and
streamfow records at gauged watersheds to relate the peak fow rate and lag time to the drainage
area. Procedures are described in FERC (2001).
The most common method available to designers for the small watersheds typically associated with
coal refuse disposal facilities is based on empirically derived coefcients for synthetic unit hydro-
graphs. The common methods for developing parameters from empirical equations include the Clark,
Snyder and SCS unit hydrograph procedures. These methods are incorporated into the widely used
computer programs for development of infow design foods (e.g., HEC-1 and HEC-HMS developed
by the U.S. Army Corps of Engineers Hydraulic Engineering Center (HEC) and similar privately
marketed programs).
9.6.1.2.1 Snyder Unit Hydrograph
The equations used for the Snyder unit hydrograph are (USACE, 1990b):
t
p
= C
t
( L L
ca
)
0.3
(9-1)
C
p
= ( Q
p
t
p
) / ( 640A) (9-2)
t
p
= time lag measured from the centroid of precipitation excess to the time of peak
fow at the point of interest (hr)
L
ca
= length along the main watercourse measured from the outlet upstream to a
point nearest the basin centroid (mi)
L = length of the main watercourse (mi)
Q
p
= peak fow rate of the unit hydrograph (cfs)
A
= drainage area (mi
2
)
The coefcients C
t
and C
p
are empirical values applicable to specifc regions that account for
watershed storage and slope and food-wave velocity and channel storage, respectively. These
parameters are obtained from regional studies and, if they are representative of conditions of the
< PREVIOUS VIEW
9-42
Chapter 9
MAY 2009
watershed being analyzed, are entered into HEC hydrologic sofware for a Snyder unit-hydro-
graph analysis.
9.6.1.2.2 Clark Unit Hydrograph
The Clark unit hydrograph uses a time-area curve to represent the watershed and uses a com-
puted time of concentration (T
c
) that can be calculated based on SCS procedures unless more
reliable regional data are available. Additionally, the Clark unit hydrograph also uses a coefcient
that refects the efect of storage within the watershed. HEC hydrologic sofware (e.g., HEC-1) can
be used to calculate the value of this coefcient through its optimization routine, but the result
obtained should be evaluated and compared to published or available regional data. The Clark
method is usually not employed for the small watersheds that are typically associated with coal
refuse disposal facilities.
9.6.1.2.3 SCS Dimensionless Unit Hydrograph
The SCS method is the most commonly used approach for small watersheds and is frequently used
for coal refuse disposal facility design. The primary analytical requirement for this method to be
applied in a HEC-1 analysis is the estimation of the lag time for the basin, which is generally assumed
to be equal to 0.6 T
c
.
Time of Concentration and Lag Time
The time of concentration T
c
is the time required for runof to travel from the most hydrologically
remote point in the watershed to the point of interest (Figure 9.2). The hydrologically most distant
path within a watershed may not necessarily be along the longest water course; therefore, various
watershed length and slope combinations should be evaluated.
Empirical equations have been developed by the SCS, the USBR, and others for estimation of T
c
as a
function of the length, surface texture and vegetation, and watershed slopes. Additionally, T
c
may be
computed by analysis of the overland and channel fow travel time using surface drainage sofware.
A common method is to use the computer program TR-55 to determine fow velocity and associated
time of concentration for subbasins within a watershed and thus estimate T
c
. Empirical equations for
determination of T
c
are presented below.
USBR Method
The USBR (1973) determined T
c
from
the following equation that has historically been applied to
small watersheds for design of small dams and coal refuse disposal facilities:
T
c
= [ (11.9L
3
) / H]
0.385
(9-3)
where:
T
c
= time of concentration (hr)
L
= length of longest watercourse in watershed (mi)
H
= elevation diference between the highest and lowest points in the watershed (f)
For watersheds west of the 105
th
meridian and forested mountain watersheds east of the 105
th
merid-
ian, Table 9.8 lists correction factors that should be applied to T
c,
as predicted by Equation 9-3. The
Modifed Snyder Method developed by USBR (1987a), which is discussed in subsequent paragraphs,
is now more commonly used.
< PREVIOUS VIEW
9-43
Hydrology and Hydraulics
MAY 2009
TABLE 9.8 RECOMMENDED CORRECTION FACTORS FOR T
C
FOR WATERSHEDS
WEST OF THE 105
TH
MERIDIAN
CN T
c
/
T
c
80 1.0
70 1.4
60 1.8
50 2.2
(USBR, 1973)
SCS Methods
The lag method developed by the Natural Resources Conservation Service (NRCS, 1986), also referred
to as the curve-number method, applies to areas less than 2,000 acres:
T
c
=
5
3
[(L
0.8
(S+1)
0.7
) / 1900Y
0.5
] (9-4)
where:
S = ( 1000/ CN) 10
L
= hydraulic length of watershed (f)
Y
= average watershed land slope (percent)
CN
= curve number (dim)
The NRCS (1986) also developed an approach to determining time of concentration by computing the
travel time (T
t
) for runof to traverse the watershed, considering three components of fow: (1) sheet
fow in upland areas (generally applied to distances of 300 feet or less), (2) shallow concentrated fow
as runof concentrates beyond the sheet fow areas, and (3) open-channel fow as runof is conveyed
downstream. By summing the travel times, an estimate of T
c
can be determined from the following
relationship:
T
c
= T
t1
+ T
t2
+ T
t3
(9-5)
Sheet Flow
T
t1
= [ 0.007 ( nL
1
)
0.8
] / (P
2
0.5
s
0.4
) (9-6)
where:
T
t1
= sheet fow travel time (hr)
n
= Mannings roughness coefcient for sheet fow (Table 9.9)
L
1
= fow length (f)
P
2
= 2-year, 24-hour rainfall (in)
s
= average watershed land slope (f/f)
< PREVIOUS VIEW
9-44
Chapter 9
MAY 2009
TABLE 9.9 ROUGHNESS COEFFICIENTS (MANNINGS N) FOR SHEET FLOW
Surface Description n
Smooth surfaces (concrete, asphalt, gravel, or bare soil) 0.011
Fallow (no residue) 0.05
Cultivated soils
Residue cover 20% 0.06
Residue cover 20% 0.17
Grass
Short grass prairie 0.15
Dense grasses
(1)
0.24
Bermuda grass 0.41
Range (natural) 0.13
Woods
(2)
Light underbrush 0.40
Dense underbrush 0.80
Note: 1. Includes species such as weeping lovegrass, bluegrass, buffalo grass, blue grama
grass and native grass mixtures.
2. When selecting n, cover should be assumed to have a height of 0.1 foot. This is
the only portion of the plant cover that will obstruct fow.
(NRCS, 1986)
Shallow Concentrated Flow
T
t2
= L
2
/ ( 3600V
2
) (9-7)
where:
T
t2
= shallow concentrated fow travel time (hr)
L
2
= fow length (f)
V
2
= average velocity (f/sec) from Figure 9.16
Open-Channel Flow
T
t3
= L
3
/ ( 3600V
3
) (9-8)
V
3
= ( 1.49 R
0.67
s
0.5
)/ n
where:
T
t3
= open-channel-fow travel time (hr)
L
3
= channel fow length (f)
R
= hydraulic radius = cross-sectional-fow area/weted perimeter (f)
s
= slope of the channel (f/f)
n
= Mannings roughness coefcient (Section 9.7.2.2)
V
3
= average channel fow velocity (f/sec)
< PREVIOUS VIEW
9-45
Hydrology and Hydraulics
MAY 2009
FIGURE 9.16 AVERAGE VELOCITIES FOR ESTIMATING
TRAVEL TIME FOR
SHALLOW, CONCENTRATED FLOW
FIGURE 9.16 AVERAGE VELOCITIES FOR ESTIMATING TRAVEL
TIME FOR SHALLOW CONCENTRATED FLOW
U
N
P
A
V
E
D
P
A
V
E
D
AVERAGE VELOCITY (FT/SEC)
1 2 4 6 10 20
0.005
0.01
0.02
0.04
0.06
0.10
0.20
0.50
W
A
T
E
R
C
O
U
R
S
E
S
L
O
P
E
(
F
T
/
F
T
)
(NRCS, 1986)
FIGURE 9.16 AVERAGE VELOCITIES FOR ESTIMATING TRAVEL TIME FOR
SHALLOW, CONCENTRATED FLOW
< PREVIOUS VIEW
9-46
Chapter 9
MAY 2009
USBR (Modifed Snyder) Method
The USBR (1987a) provides guidance for unit hydrograph development based on lag time using the Modi-
fed Snyder Equation, including charts for regions of the country to assist in estimating the lag time:
L
g
= C[ ( L L
ca
) / S
0.5
) ]
0.33
(9-9)
where:
L
g
= unit hydrograph lag time (hr)
C
= constant (estimated as 26 times the average Mannings n value for in-channel
fows with higher values for overbank fow conditions)
L
= length of main watercourse (mi)
L
ca
= length along the main water course measured from the outlet upstream to a
point nearest the basin centroid (mi)
S
= overall slope for longest water course (f/mi)
9.6.1.3 Precipitation-Runof Relationship
Generally, not all of the precipitation that falls on a watershed during a design storm becomes runof;
a portion is retained in the soil and on vegetation. Chow (1964), the USDA (1972), the NRCS (2004a),
and the USBR (1973, 1987a) and other references on hydrology discuss the watershed characteristics
that determine the amount of precipitation that becomes runof. These characteristics include: (1) the
types of soil and their efect on the amount of water seeping into the ground; (2) the conditions (wet,
dry, snow covered or frozen) of the ground surface immediately prior to the precipitation (known as
the antecedent moisture condition or AMC); (3) the type and density of vegetation; (4) the types of
development, such as paved surfaces, channeling and storm sewers; and (5) dams, lakes, ponds, or
swamps upstream of the site that could store water on a permanent basis, release water at a slow rate,
or fail, thereby suddenly releasing large volumes of water.
9.6.1.3.1 General Rainfall Conditions
The NRCS (updating previous work when it was known as the SCS) has quantifed precipitation
runof conditions for a wide range of soil, moisture and soil-cover conditions. The procedure utilizes
a runof curve number (CN) and the following equations:
Q = ( P I
a
)
2
/ ( P + 0.8S) (9-10)
S = ( 1000/ CN) 10 (9-11)
where:
Q = direct runof (in)
P
= rainfall (in)
I
a
= initial abstraction = 0.2 S
S
= maximum potential diference between P and Q at beginning of storm
Figures 9.17 and 9.18 present charts for estimating runof from precipitation based on the above equa-
tions. Discussion of the derivation and use of these charts is presented in the NRCS (2004b) National
Engineering Handbook (NEH) updating previous work by the USDA (1972). To determine the char-
< PREVIOUS VIEW
9-47
Hydrology and Hydraulics
MAY 2009
acteristic CN, it is necessary to know or to estimate four watershed conditions: (1) the hydrologic soils
classifcation, (2) the land use and surface status or treatment, (3) the hydrologic efect of the land use
and status, and (4) the antecedent moisture condition (AMC).
Traditionally, the hydrologic soil classifcation has been determined from county soil surveys. The soils
identifed in the soil surveys are categorized into four hydrologic soil groups (HSGs), as described in
NEH Chapter 7 (NRCS, 2007a). The HSGs for soils of the United States are presented in Appendix A
of TR55 (NRCS, 1986; Appendix A updated 1999). Groups A through D for natural soils are described
below. The disturbed area soil profle for each group for watershed areas that have been afected by
mining related impacts and/or urbanization is also provided.
A. Low runof potential Soils having high infltration rates even when thoroughly
weted consisting chiefy of well to excessively drained deep sands or gravels, typi-
cally with less than 10 percent clay. These soils have a high rate of water transmis-
sion.
Soil description: Sand, loamy sand, or sandy loam.
B. Moderately low runof potential Soils having moderately low infltration rates when
thoroughly weted, consisting chiefy of moderately deep to deep soils with moder-
FIGURE 9.17 DIRECT RUNOFF FOR RAINFALL LESS THAN OR EQUAL TO 8 INCHES
FIGURE 9.17 DIRECT RUNOFF FOR RAINFALL LESS
THAN OR EQUAL TO 8 INCHES
RAINFALL, P (IN)
0
1
2
3
4
5
6
7
8
D
I
R
E
C
T
R
U
N
O
F
F
,
Q
(
I
N
)
(USDA, 1972)
1
0
0
9
5
9
0
8
5 8
0
7
5
7
0
6
5
6
0
5
5
5
0
4
5
4
0
3
5
3
0
1 2 3 4 5 6 7 8 0
NOTE: S = POTENTIAL MAXIMUM RETENTION
AFTER RAINFALL BEGINS
C
U
R
V
E
N
U
M
B
E
R
(
C
N
)
=
1
0
0
0
1
0
+
S
Q =
(P - 0.2S)
P + 0.8S
2
FIGURE 9.17 DIRECT RUNOFF FOR RAINFALL LESS THAN OR EQUAL TO 8 INCHES
< PREVIOUS VIEW
9-48
Chapter 9
MAY 2009
ately fne to moderately coarse textures, moderately well to well drained, typically
with 10 to 20 percent clay. These soils have a moderate rate of water transmission.
Soil Description: Silt loam or loam.
C. Moderately high runof potential Soils having slow infltration rates when thor-
oughly weted, consisting chiefy of soils with fne texture, or of soils with a layer that
impedes downward movement of water, typically with 20 to 40 percent clay. These
soils have a slow rate of water transmission.
Soil Description: Sandy clay loam.
D. High runof potential Soils having very slow infltration rates when thoroughly
weted. These are typically clay soils with high swelling potential, soils with a clay-
pan or clay layer at or near the surface, shallow soils over nearly impervious mate-
rial, and soils with a permanent high water table. These soils typically have greater
than 40 percent clay and a very slow rate of water transmission.
Soil Description: Clay loam, silty clay loam, sandy clay, silty clay, or clay
FIGURE 9.18 DIRECT RUNOFF FOR RAINFALL GREATER THAN OR EQUAL TO 8 INCHES
FIGURE 9.18 DIRECT RUNOFF FOR RAINFALL GREATER
THAN OR EQUAL TO 8 INCHES
0 4 8 12 16 20 24 28 32 36 40
0
4
8
12
16
20
24
28
32
36
40
RAINFALL, P (IN)
D
I
R
E
C
T
R
U
N
O
F
F
,
Q
(
I
N
)
(USDA,1972)
Q =
(P - 0.2S)
P + 0.8S
C
U
R
V
E
N
U
M
B
E
R
C
N
=
1
0
0
0
1
0
+
S
3
0
2
0
1
5
2
5
3
5
1
0
0
1
0
4
0
4
5
5
0
8
0
6
0
7
0
9
0
2
NOTE: S = POTENTIAL MAXIMUM RETENTION
AFTER RAINFALL BEGINS
FIGURE 9.18 DIRECT RUNOFF FOR RAINFALL GREATER THAN OR EQUAL TO 8 INCHES
< PREVIOUS VIEW
9-49
Hydrology and Hydraulics
MAY 2009
The second and third watershed conditions are simply descriptions of watershed land use and
surface status or treatment, such as woods or straight row, small grain crops, and the hydrologic
density or impact of the land use. A general description of the hydrologic grading for soil-cover
conditions typically encountered is described in conjunction with selection of the curve number
(Table 9.10).
The fourth watershed condition is an index of watershed wetness referred to as the antecedent mois-
ture condition (AMC). Antecedent moisture conditions are categorized into three groups:
AMC-I Optimum soil conditions soils are dry but not to the point of wilting
vegetation (not recommended for design).
AMC-II Average conditions and average value for annual foods.
AMC-III Wet or saturated conditions associated with heavy rainfall or light rainfall
and low temperatures within 5 days prior to the given storm.
The antecedent moisture condition that should be used for design of impoundment structures is
either AMC-II or AMC-III. The NRCS provides procedures for estimating the antecedent moisture
condition based on the 5 days of antecedent rainfall. Typically AMC-II is used for the design of drain-
age channels, and AMC-III is used when determining the PMF for impoundment design.
The curve number can be estimated based on the hydrologic soils group, land use, and soil-cover
conditions, as presented in Table 9.10 for AMC-II. Adjustments in the estimated CN value can be
made for AMC-I and AMC-III using Table 9.11. The NRCS (2004b) provides methods for developing
composite CN values when there are multiple antecedent moisture conditions.
Typically, basin or subbasin averaging is performed (as in the HEC-1 model) to represent watershed
areas and compute runof hydrographs using SCS or Snyder parameters. However, distributed cal-
culations based on mapped soil and land use conditions yield more representative runof estimates
than basin-averaged parameters, and the availability of programs such as HEC-GeoHMS that utilize
GIS terrain and spatial information will efciently facilitate this approach when digitized soils sur-
veys are available. An overview of procedures that can be employed is presented in FERC (2001).
When calculating runof, it is important to consider actual conditions that may be present in the
watershed, including the potential efects of any impoundments. The following guidelines are appli-
cable to refuse disposal facilities:
For watersheds having varying runof characteristics, subbasins should be developed
or a weighted average CN should be computed.
The actual impoundment is generally considered impervious and contributes 100
percent to the runof.
Upstream impoundments require special consideration when determining runof.
Runof hydrographs should be routed through such impoundments to confrm their
operation during the design storm and the potential impact should the structures
overtop and fail.
9.6.1.3.2 Thunderstorm Rainfall Conditions
Hydrometeorological reports for the western regions of the U.S. provide guidance for analysis of
thunderstorm rainfall. For the Rocky Mountain region, the USBR (1987a) identifes a local, high-
intensity thunderstorm event that should be considered along with the general storm event. Charts
provide guidance for selection of lag time and development of unit hydrographs.
< PREVIOUS VIEW
9-50
Chapter 9
MAY 2009
TABLE 9.10 RUNOFF CURVE NUMBERS FOR WATERSHED COMPLEXES AND AMC-II
Cover Description
Curve Numbers for
Hydrologic Soil Group
Cover Type
Hydrologic
Condition
A B C D
Cultivated Agricultural Lands
(1, 2, 3, 4)
Fallow
Bare soil 77 86 91 94
Crop residue cover (CR) Poor 76 85 90 93
Good 74 83 88 90
Row crops
Straight row (SR) Poor 72 81 88 91
Good 67 78 85 89
SR + CR Poor 71 80 87 90
Good 64 75 82 85
Contoured (C) Poor 70 79 84 88
Good 65 75 82 86
C + CR Poor 69 78 83 87
Good 64 74 81 85
Contoured & terraced (C&T) Poor 66 74 80 82
Good 62 71 78 81
C&T + CR Poor 65 73 79 81
Good 61 70 77 80
Small grain
SR
Poor 65 76 84 88
Good 63 75 83 87
SR + CR
Poor 64 75 83 86
Good 60 72 80 84
C
Poor 63 74 82 85
Good 61 73 81 84
C + CR
Poor 62 73 81 84
Good 60 72 80 83
C&T
Poor 61 72 79 82
Good 59 70 78 81
C&T + CR
Poor 60 71 78 81
Good 58 69 77 80
Close-seeded or
broadcast legumes
or rotation meadow
SR
Poor 66 77 85 89
Good 58 72 81 85
C
Poor 64 75 83 85
Good 55 69 78 83
C&T
Poor 63 73 80 83
Good 51 67 76 80
Pasture, grassland, or range continuous
forage for grazing
(5)
Poor
68 79 86 89
Fair 49 69 79 84
Good 39 61 74 80
Other Agricultural Lands
(1)
Meadow continuous grass, protected from
grazing and generally mowed for hay
30 58 71 78
Brush brush-weed-grass mixture with brush
the major element
(6)
Poor 48 67 77 83
Fair 35 56 70 77
Good 30
(7)
48 65 73
Woods grass combination (orchard or tree
farm)
(8)
Poor 57 73 82 86
Fair 43 65 76 82
Good 32 58 72 79
< PREVIOUS VIEW
9-51
Hydrology and Hydraulics
MAY 2009
Cover Description
Curve Numbers for
Hydrologic Soil Group
Cover Type
Hydrologic
Condition
A B C D
Woods
(9)
Poor 45 66 77 83
Fair 36 60 73 79
Good 30
(7)
55 70 77
Farmsteads buildings, lanes, driveways, and
surrounding lots
59 74 82 86
Arid and Semiarid Range Lands
(1, 10)
Herbaceous mixture of grass, weeds, and low-
growing brush, with brush the minor element
Poor 80 87 93
Fair 71 81 89
Good 62 74 85
Oak-aspen mountain brush mixture of oak
brush, aspen, mountain mahogany, bitter brush,
maple, and other brush
Poor 66 74 79
Fair 48 57 63
Good 30 41 48
Pinyon-juniper pinyon, juniper, or both; grass
understory
Poor 75 85 89
Fair 58 73 80
Good 41 61 71
Sagebrush with grass understory
Poor 67 80 85
Fair 51 63 70
Good 35 47 55
Desert shrub major plants include saltbush,
greasewood, creosotebush, blackbrush,
bursage, palo verde, mesquite, and cactus
Poor 63 77 85 88
Fair 55 72 81 86
Good 49 68 79 84
Note: 1. Average runoff conditions and I
a
= 0.2S.
2. Crop residue cover applies only if residue is on at least 5 percent of the surface throughout the year.
3. Hydrologic condition is based on a combination of factors that affect infltration and runoff, including:
(a) density and canopy of vegetative areas, (b) amount of year-round cover, (c) amount of grass or
close-seeded legumes, (d) percent of residue cover on the land surface (good 20 percent), and
(e) degree of surface roughness.
4. Poor: Factors impair infltration and tend to increase runoff.
Good: Factors encourage average and better than average infltration and tend to decrease runoff.
5. Poor: < 50 percent ground cover or heavily grazed with no mulch.
Fair: 50 to 75 percent ground cover and not heavily grazed.
Good: > 75 percent ground cover and lightly or only occasionally grazed
6. Poor: < 50 percent ground cover
Fair: 50 to 75 percent ground cover
Good: > 75 percent ground cover
7. Actual curve number is less than 30; use CN = 30 for runoff computations.
8. CNs shown were computed for areas with 50 percent woods and 50 percent grass (pasture) cover.
Other combinations may be computed from the CNs for woods and pasture.
9. Poor: Forest litter, small trees, and brush are destroyed by heavy grazing or regular burning.
Fair: Woods are grazed, but not burned, and some forest litter covers the soil.
Good: Woods are protected from grazing, and litter and brush adequately cover the soil.
10. Poor: < 30 percent ground cover (litter, grass, and brush overstory)
Fair: 30 to 70 percent ground cover
Good: > 70 percent ground cover
(ADAPTED FROM NRCS, 1986)
TABLE 9.10 RUNOFF CURVE NUMBERS FOR WATERSHED COMPLEXES AND AMC-II
(Continued)
< PREVIOUS VIEW
9-52
Chapter 9
MAY 2009
9.6.1.4 Channel and Impoundment Storage Characteristics
For non-impounding coal refuse disposal facilities, storage normally does not enter into the hydro-
logic and hydraulic analyses. Channels are generally designed for the peak runof rate, and storage
beyond the channel capacity is not a consideration. Channel and food plain storage can be a consid-
eration when evaluating some natural drainage systems, and hydraulic analysis sofware can incor-
porate the associated storage into the model through cross sections determined from topographic
maps. For impounding disposal facilities, reservoir storage provisions are an important component
of the hydraulic design and can vary between:
The condition where the impoundment does not have reservoir storage capacity to
handle any signifcant portion of the runof during a storm, requiring the outfow
to essentially equal the infow rate. This condition usually requires a relatively large
spillway to pass the storm infow.
The condition where the impoundment does have storage capacity to temporarily
handle all runof during a storm, allowing the immediate outfow discharge to be
essentially zero. This condition requires an embankment high enough to safely store
the storm infow.
Most impounding disposal facilities during some period of their life fall between these extremes, and
part of the runof becomes reservoir storage and the remainder is passed as outfow. An important
factor that diferentiates coal refuse disposal facility impoundments from other types of impound-
ments is that the embankment and impoundment confgurations continually change with both the
TABLE 9.11 RUNOFF CURVE NUMBERS (CN) FOR ANTECEDENT MOISTURE CONDITIONS
AMC II
AMC I
(Not Recommended for
Hydrologic Design)
AMC III
100 100 100
95 87 98
90 78 96
85 70 94
80 63 91
75 57 88
70 51 85
65 45 82
60 40 78
55 35 74
50 31 70
40 22 60
30 15 50
20 9 37
10 4 22
0 0 0
(USDA, 1972)
< PREVIOUS VIEW
9-53
Hydrology and Hydraulics
MAY 2009
disposal of coarse refuse on the embankment and the disposal of fne refuse slurry into the impound-
ment. This variation must be taken into account in the hydraulic design of a disposal facility.
Several terms commonly used to describe conditions associated with an impoundment are defned below:
Normal pool elevation The surface elevation of water or slurry impounded by an
embank ment during normal disposal operations. The normal pool elevation is usually
established by the decant inlet level or pump control level. With slurry impoundments,
the normal pool elevation changes with time as fne refuse slurry is disposed and the
decant level or pump control level is raised to successively higher elevations.
Minimum pool elevation The lowest surface elevation that can normally be
atained. This is ofen the same as the normal pool elevation, but can be lower if
drainage is provided by siphoning, pumping, or modifying the outlet system.
Useful storage The storage capacity between the normal and minimum pool elevations.
Dead storage The storage capacity below the minimum pool elevation that is usu-
ally flled primarily by setled slurry.
Surcharge storage The storage capacity between the normal pool elevation and the
maximum permissible pool elevation. This capacity is intended primarily for tempo-
rary storage of runof during storms.
Freeboard The diference in elevation between the dike or embankment crest (i.e.,
lowest portion of impoundment perimeter except for an open-channel spillway)
and the impoundment water surface. Normal freeboard is the distance between the
minimum embankment crest elevation and the normal pool level, as established by
the lowest outlet structure used for food routing purposes. Design-storm freeboard
is the distance between the embankment crest and the maximum pool level during
the design storm. The design-storm freeboard should be such that the embankment
is not overtopped.
Reservoir storage volumes can be determined in several ways, all based on the topographic confgu-
ration of the impoundment area. Topographic data, generally obtained by aerial photography with
2- to 5-foot contour intervals, are suitable for most analyses. USGS topographic quadrangle maps
may be used for preliminary calculations for evaluation of initial feasibility.
The normally preferred procedure for calculating the elevation-volume relationship for an impound-
ment is described in Section 9.2.3 and is illustrated in Figure 9.4. The impoundment surface area for
each successive elevation contour is frst determined, from which an area-elevation curve can be plot-
ted. The storage is then computed as the area beneath the area-elevation curve at any elevation, from
which the volume-elevation curve can be ploted. CADD sofware can be used to generate area- and
volume-elevation data. Typically, both area and volume curves are presented as part of an impound-
ment design report or plan.
Design of impoundments for upstream or centerline construction involves placement of subsequent
embankment stages within the impoundment area, which impacts reservoir storage associated with
that stage of construction. Accordingly, the impoundment area-elevation and volume-elevation data
for each stage of construction must allow for the efect of upstream or centerline construction.
Surcharge storage and freeboard to accommodate the design-storm runof is required for most
impoundments. In determining the surcharge storage, the volume of the setled fnes above the pool
level (delta deposits) must be taken into account, particularly for diked confgurations where slurry is
discharged over a substantial perimeter of the reservoir. The volume of the delta deposits is generally
estimated based upon the position and elevation of the slurry discharge and an assumed slope of the
deposit (typically between 1 and 3 percent).
< PREVIOUS VIEW
9-54
Chapter 9
MAY 2009
Freeboard at an impoundment is provided in order to account for such factors as uncertainties
in the hydrologic analyses, setlement of the embankment crest, and extreme wind efects such
as wave runup. The design-storm freeboard for any impounding embankment is a function of
the wave height and the wave runup conditions at the upstream face of the embankment. Guid-
ance for the evaluation of freeboard for reservoirs is provided in USBR (1987a). Other factors that
should be considered in determining freeboard requirements include: (1) frequency of the design
storm, (2) duration of high water level, (3) ability to resist erosion, (4) and potential for setle-
ment or mine subsidence. A reference for wave runup analysis is the Coastal Engineering Manual
developed by the USACE (2002). Coal refuse impounding embankments are typically required
to have a design-storm freeboard of 3 feet, which is consistent with a wind fetch of generally less
than one mile.
9.6.2 Runof Determination: Hydrograph Method
When a time history of runof is required for fnal reservoir routing (Section 9.8), a runof hydrograph
must be developed. As illustrated in Figure 9.3, a hydrograph is a plot of fow rate at the point of
interest versus time following storm initiation. Development of a runof hydrograph requires the
watershed runof data discussed in the preceding section. The data are used to create unit value
runof hydrographs for small time increments within the total storm duration. A unit hydrograph
models the time history of runof fowing from the watershed at the point of interest for one time
increment of precipitation. A composite runof hydrograph can then be constructed by superimpos-
ing unit hydrographs for all increments of precipitation.
Computer programs developed by the USACE Hydrologic Engineering Center (HEC) have ofen
been used for the hydrologic and hydraulic design of coal refuse disposal facilities. The HEC-1 Flood
Hydrograph package calculates runof hydrographs and has several optional capabilities. Use of
HEC-1 for precipitation-runof modeling requires subbasin boundary delineation, precipitation data,
and runof and routing parameters. Typically, synthetic unit hydrographs based on the SCS, Snyder
or Clark methods are employed. The HEC-Hydrologic Modeling System (HEC-HMS) computer pro-
gram provides advancements in several areas over the HEC-1 program, including the optional use of
distributed analysis of runof through interfacing with GIS terrain models.
The following capabilities of HEC-1 and HEC-HMS are frequently utilized during the design of coal
refuse disposal facilities:
Distributed runof analysis using kinematic wave and Muskingum-Cunge routing,
which can provide a more refned analysis of peak fows for impoundment designs
incorporating open-channel spillways.
Modeling of base fow, which may be important for large watersheds and inundation
analysis.
Channel routing using a variety of methods (e.g., Muskingum, Modifed Puls, etc.) to
perform inundation analysis and evaluation of impoundments in series.
Reservoir routing (level pool routing) based on storage, elevation, and outfow rating
parameters, which is a basic requirement for impounding facility design.
Dam breach and inundation analysis for determination of the hazard-potential
classifcation for an impounding facility and for preparation of an Emergency
Action Plan.
Users Manuals for HEC-1 (USACE, 1998b) and HEC-HMS (USACE, 2000) provide documenta-
tion of their capabilities and applications. Several private companies market these programs with
enhanced input and output features and also have developed similar programs with enhanced
capabilities.
< PREVIOUS VIEW
9-55
Hydrology and Hydraulics
MAY 2009
Regardless of the method of hydrograph development and analysis, the following computation and
data validation issues should be considered:
The computational time increment should be selected based on the lag time and
precipitation data such that a smooth hydrograph is obtained. The maximum time
increment of rainfall to be used in the hydrograph analysis should be the lag time
divided by 5 ( L
g
/ 5) rounded to the next lower even number (FERC, 2001). The
USACE (1998b) indicates that the time increment should be no larger than 0.29 L
g
.
It can be problematic to apply this limitation to small watersheds with small lag
times because some sofware programs limit the time increment; however HEC-
HMS does not have this limitation. For impoundments that are designed to store
the design storm with release through a decant pipe, the volume of runof and
peak impoundment pool level is relatively insensitive to the computation time
increment.
Determination of the lag time and time of concentration using multiple methods is
a useful check of the parameter validity prior to use in hydrograph simulation. If a
regional analysis is performed to estimate these parameters, performing a check on
the time of concentration using the TR-55 computer program should be considered.
The applicability of any hydrograph development method is subject to uncertainty,
and verifcation is sometimes accomplished by investigating multiple methods (and
sources of watershed parameters such as the USACE) or by reproducing a large
historical food of record from a watershed with similar characteristics. This consid-
eration is most important for impoundments where the food hydrograph is routed
through an open-channel spillway.
9.6.3 Peak Runof Determination Key Parameters Method
When a complete time history of runof is not required, estimates of the key runof parameters can be
used to determine the total runof volume and peak runof rate developed during a design storm. The
key parameters method is useful for fnal design when: (1) the entire volume of runof will be tempo-
rarily stored so that the rate of runof does not need to be calculated or (2) the storage capacity is so
small that the peak outfow rate is essentially the same as the peak runof rate and reservoir routing
is not necessary. The key parameters method is also useful for estimating approximate storage and
outfow requirements during feasibility planning for various facility confgurations and preliminary
sizing for various hydraulic structures prior to undertaking more detailed procedures using hydro-
graphs and reservoir routing analyses.
The USDA (1973b) published charts for the determination of total runof volume and peak runof
rates for small volumes of runof that can be used to estimate key parameters for design purposes.
These charts provide runof values based on watershed size, slope, and runof curve number (CN).
Figures 9.19 and 9.20
present total runof volume as a function of total precipitation and watershed
size for runof CN = 70 and CN = 80, respectively. The curves were derived directly from Figures 9.17
and 9.18 and do not include the minimum retentions due to soil infltration discussed in Section
9.6.1.3. Through interpolation and extrapolation, these curves may be used to estimate total runof
volume for other CN values.
Figures 9.21 and 9.22
may be used for estimating the peak runof rate for a range of six-hour design
storms for CN = 70 and CN = 80. These fgures were prepared from hydrograph analyses utiliz-
ing the recommended SCS rainfall distribution presented in Figure 9.15. Through interpolation
or extrapolation, these curves can be used to estimate the peak runof rate for other CN values
between 70 and 80.
< PREVIOUS VIEW
9-56
Chapter 9
MAY 2009
FIGURE 9.19 ESTIMATED TOTAL RUNOFF VOLUME FOR CN = 70
FIGURE 9.19 ESTIMATED TOTAL RUNOFF VOLUME FOR CN = 70
T
O
T
A
L
V
O
L
U
M
E
O
F
R
U
N
O
F
F
(
A
C
R
E
-
F
E
E
T
)
1
10
100
1,000
10,000
50,000
TOTAL RAINFALL (IN)
WATERSHED (ACRES)
60
40
50
35
30
25
20
15
10
5
5 10 100 1,000 10,000
FIGURE 9.19 ESTIMATED TOTAL RUNOFF VOLUME FOR CN = 70
< PREVIOUS VIEW
9-57
Hydrology and Hydraulics
MAY 2009
FIGURE 9.20 ESTIMATED TOTAL RUNOFF VOLUME FOR CN = 80
FIGURE 9.20 ESTIMATED TOTAL RUNOFF VOLUME FOR CN = 80
T
O
T
A
L
V
O
L
U
M
E
O
F
R
U
N
O
F
F
(
A
C
R
E
-
F
E
E
T
)
1
10
100
1,000
10,000
50,000
TOTAL RAINFALL (IN)
60
40
50
35
30
25
20
15
10
5
WATERSHED (ACRES)
10,000 1,000 100 10 5
FIGURE 9.20 ESTIMATED TOTAL RUNOFF VOLUME FOR CN = 80
< PREVIOUS VIEW
9-58
Chapter 9
MAY 2009
9.6.4 Runof Determination Rational Method
The rational method is the simplest procedure for estimating peak runof rates and is typically used
for developing design fows for minor drainage features. Originally developed by the U.S. Bureau
of Public Roads and subsequently updated by the FHWA (2001), the method is usually restricted to
watersheds of less than 200 acres and to storm recurrence intervals of less than 100 years. The USGS
(2005) presents a review and comparison of the method with observed runof events. Using the ratio-
nal method, the peak rate of runof is determined from the following relationship:
Q = Ci A (9-12)
Q = peak rate of runof (cfs)
C = weighted average runof coefcient
i = average precipitation intensity for a duration equal to the watershed time of
concentration and the selected storm recurrence interval (in/hr)
A
= watershed area tributary to the point of interest (acres)
FIGURE 9.21 ESTIMATED PEAK RUNOFF RATE FOR 6-HOUR DESIGN STORM AND CN = 70
TIME OF CONCENTRATION (HR)
FIGURE 9.21 ESTIMATED PEAK RUNOFF RATE FOR
6-HOUR DESIGN STORM AND CN = 70
P
E
A
K
R
A
T
E
O
F
R
U
N
O
F
F
(
C
F
S
/
A
C
R
E
)
6
-
H
O
U
R
R
A
I
N
F
A
L
L
(
I
N
)
40.0
20.0
0.05
0.1
0.2
0.5
1.0
2.0
5.0
10.0
4.0 0.1 0.2 0.5 1.0 2.0
34
2
3
4
6
10
14
18
22
26
30
FIGURE 9.21 ESTIMATED PEAK RUNOFF RATE FOR 6-HOUR DESIGN STORM AND CN = 70
< PREVIOUS VIEW
9-59
Hydrology and Hydraulics
MAY 2009
The runof coefcient C can be further defned as the ratio of the rate of runof to the rate of precipita-
tion during the storm when all the drainage area is contributing to the runof. The value of C can be
estimated from the data provided in Table 9.12. The range of the coefcients for rural areas permits
some allowance for diferences in slope and ground cover conditions. The lower values in the table
should only be used when the watershed is fat and the surface is permeable. When the watershed
exhibits multiple slope and ground cover conditions, the runof coefcient should be determined as
a weighted average based on the relative area of each of the conditions present.
The time period is equal to the time of concentration. This is the time required for the entire drainage
area to be contributing to the runof. The average precipitation intensity (i) can be determined from
precipitation frequency tables presented by the National Weather Service Precipitation Frequency
Data Server (web site) as an update for several states covered by Technical Paper 40. Table 9.5 pres-
ents publications for precipitation frequency data.
It is emphasized that use of the rational method for determining runof should be limited to the
design of minor drainage appurtenances at a coal refuse disposal facility. An example of application
of the rational method is presented in Figure 9.23.
FIGURE 9.22 ESTIMATED PEAK RUNOFF RATE FOR 6-HOUR DESIGN STORM AND CN = 80
TIME OF CONCENTRATION (HR)
FIGURE 9.22 ESTIMATED PEAK RUNOFF RATE FOR
6-HOUR DESIGN STORM AND CN = 80
P
E
A
K
R
A
T
E
O
F
R
U
N
O
F
F
(
C
F
S
/
A
C
R
E
)
6
-
H
O
U
R
R
A
I
N
F
A
L
L
(
I
N
)
2
3
4
6
10
14
18
22
34
30
26
0.1 0.2 0.5 1.0 2.0 4.0
0.05
0.1
0.2
0.5
1.0
2.0
5.0
10.0
20.0
40.0
FIGURE 9.22 ESTIMATED PEAK RUNOFF RATE FOR 6-HOUR DESIGN STORM AND CN = 80
< PREVIOUS VIEW
9-60
Chapter 9
MAY 2009
9.7 DESIGN OF OUTFLOW SYSTEMS
This section presents the planning and design requirements for hydraulic systems to safely transport
watershed runof through, around and beyond a coal refuse disposal facility. Because the discussion
is general in scope, frequent references for specifc applications are made to texts and publications on
hydraulic design and engineering, including: Chow (1959), USBR (1987a), Brater et al. (1996), Hen-
derson (1966), and the USACE (1990b).
The analytical steps necessary for hydraulic design are discussed with emphasis on the important
relationship of each system component to the overall requirements of the disposal facility. The dis-
cussion is presented in the following order:
The basic considerations that determine the types of analyses to be performed.
Introduction to basic hydraulic system components including the inlet, transport and
outlet components for both open-channel and closed-conduit fow.
TABLE 9.12 RUNOFF COEFFICIENTS FOR USE IN THE RATIONAL METHOD
Rural Areas (Haan and Barfeld, 1978)
Cover Type Terrain
Soil Texture Runoff Coeffcient
Open Sandy
Loam
Clay and Silt
Loam
Tight Clay
Woodland
Flat (0-5% slope) 0.10 0.30 0.40
Rolling (5-10% slope) 0.25 0.35 0.50
Hilly (10-30% slope) 0.30 0.50 0.60
Pasture
Flat 0.10 0.30 0.40
Rolling 0.16 0.36 0.55
Hilly 0.22 0.42 0.60
Cultivated
Flat 0.30 0.50 0.60
Rolling 0.40 0.60 0.70
Hilly 0.52 0.72 0.82
Rural Areas (FHWA, 2001)
Cover Description Runoff Coeffcient
(1)
Concrete or asphalt pavement 0.8 to 0.9
Asphalt macadam pavement 0.6 to 0.8
Gravel roadways or shoulders 0.4 to 0.6
Bare earth 0.2 to 0.9
Steep, grassed areas (2:1 slope) 0.5 to 0.7
Turf meadows 0.1 to 0.4
Forested Areas 0.1 to 0.3
Cultivated felds 0.2 to 0.4
Note: 1. For fat slopes or soils with high hydraulic conductivity, the lower values should be used; for steep slopes or low-
hydraulic-conductivity soils, the higher values should be used.
< PREVIOUS VIEW
9-61
Hydrology and Hydraulics
MAY 2009
Identification of special design considerations for hydraulic structures includ-
ing erosion protection, effects of direction changes, cavitation and materials
selection.
Discussion of the types of principal hydraulic systems commonly encountered at
coal refuse disposal facilities including spillways, decants, diversion ditches, culverts
and natural streams.
9.7.1 Basic Considerations
Geometric design constraints, the coarse and fne coal production rates and the estimated runof rate
and volume infuence the selection of the outfow structure(s) that can be utilized for a coal refuse
disposal facility embankment or impoundment. Basic considerations in the design of hydraulic sys-
tems are noted for the following four situations:
1. No runof storage In this case, the hydraulic system must be designed to handle the
maximum rate of runof (i.e., outfow capacity is equal to the peak runof rate).
2. Maximum reservoir storage Reservoir storage is maximized with limited spillway
outfow during the design storm event (Section 9.8). In this case, the designer must
provide outfow capacity for all runof in excess of the available storage. Under some
conditions the optimal solution may involve total storage of the design storm runof
with delayed release through a decant system.
3. Defned surcharge storage capacity Surcharge reservoir storage is available,
but the available volume is generally limited. In this case, an iterative procedure
must be employed to establish the optimum balance between reservoir storage
and outfow, afer which the calculated maximum outfow is used to design the
hydraulic system.
4. Fixed hydraulic structure capacity For this case, the hydraulic system capacities
are fxed and the runof that will cause failure must be determined. This situa-
tion is most ofen encountered when evaluating an existing facility or the fow in a
natural stream.
The types of impounding and non-impounding disposal facilities and their unique characteristics
have been discussed in Sections 9.3 and 9.4. The major points of these previous discussions are sum-
marized below:
To accommodate the progressive development and operation of coal refuse disposal
facilities, intermediate-stage hydraulic drainage systems may be needed.
All operational periods of the impoundment hydraulic systems will need to be eval-
uated if the progressive development of a coal refuse disposal facility impacts avail-
able reservoir storage and/or outfow capacity.
Coal operators may be performing site development construction and, as a con-
sequence, have certain types of equipment always available on site; the design of
hydraulic system components should refect the equipment available and operator
capabilities.
The reclaimed confguration of a coal refuse disposal facility may impact hydraulic
structure design at intermediate facility stages. The fnal facility abandonment con-
fguration must be taken into account in the design of intermediate stage hydraulic
structures.
The design of hydraulic structures should take into account corrosion, abrasion, ero-
sion, weathering and maintenance requirements.
< PREVIOUS VIEW
9-62
Chapter 9
MAY 2009
FIGURE 9.23 EXAMPLE OF RATIONAL METHOD OF INFLOW CALCULATION
FIGURE 9.23 EXAMPLE OF RATIONAL METHOD OF INFLOW CALCULATION
STORM DURATION (MIN)
0
2
4
6
8
10
S
T
O
R
M
I
N
T
E
N
S
I
T
Y
(
I
N
/
H
R
)
1 2 5 10 20 50 100
PROBLEM DETERMINE THE RUNOFF FROM A 100-YEAR RECURRENCE INTERVAL STORM FOR A STEEPLY-SLOPED, 85-ACRE
WATERSHED IN ALLEGHENY COUNTY, PENNSYLVANIA. THE WATERSHED HAS THE FOLLOWING CHARACTERISTICS:
LONGEST WATERCOURSE = 1900 FEET
ELEVATION DIFFERENCE = 100 FEET
TOPOGRAPHY = 60% WOODED, 30% PASTURE, 10% BARE EARTH
SOLUTION 1. DETERMINE A WEIGHTED-AVERAGE RUNOFF COEFFICIENT FROM TABLE 9.12
TOPOGRAPHY RUNOFF COEFFICIENT
WOODED-STEEP 0.30
PASTURE-STEEP 0.70
BARE EARTH-STEEP 0.90
THEN C = 0.60 (0.30) + 0.30 (0.70) + 0.10 (0.90) = 0.48
2. DETERMINE 100-YEAR INTENSITY-DURATION RELATIONSHIP FROM TP-40 (HERSHFIELD, 1961)
DURATION
PORTION OF 30-MIN
RAINFALL
RAINFALL INTENSITY
HR MIN IN IN/HR
1 60 2.35 2.35
30 1.00 1.85 3.70
15
(1)
0.72 1.33 5.32
10
(1)
0.57 1.05 6.30
5
(1)
0.37 0.68 8.16
NOTE: 1. FROM PAGE 5 OF TP-40.
3. DETERMINE 100-YEAR RAINFALL INTENSITY (i ) FOR SITE
FROM FIGURE 30 OF USBR (1973), FOR L = 1,900 FEET AND H = 100 FEET, T
C
= 8 MINUTES
FROM INTENSITY-DURATION PLOT ABOVE, i = 6.9 INCHES/HOUR
4. CALCULATE 100-YEAR PEAK RUNOFF FOR SITE:
Q = Ci A = 0.48 ( 6.9) ( 85) = 281.5 CFS
FIGURE 9.23 EXAMPLE OF RATIONAL METHOD OF INFLOW CALCULATION
< PREVIOUS VIEW
9-63
Hydrology and Hydraulics
MAY 2009
9.7.2 Hydraulic System Components
Hydraulic systems that transport runof through, around and beyond coal refuse disposal facilities
have three basic components:
1. The inlet, where fow enters the system.
2. The transport or conveyance section that carries fow between the inlet and outlet.
3. The outlet, where water is discharged in an acceptable manner.
The rate of fow can be controlled by any portion of the hydraulic system by varying the size, loca-
tion, elevation, slope, shape or confguration of these three components. Interrelationships between
the inlet, the transport section and the outlet must be evaluated as part of hydraulic system design.
An important design consideration when optimizing the balance between storage and outfow (Sec-
tion 9.8) is that the maximum head of water needed to develop the desired rate of outfow does not
exceed the maximum permissible impoundment pool elevation at any time during the life of the
structure.
The following sections detail important aspects of each hydraulic system component, identifying
appropriate design procedures and discussing the role of each component in various types of hydrau-
lic systems.
9.7.2.1 Inlets
The inlet to a hydraulic system can: (1) simply direct fow to the transport section or (2) regulate the
rate of fow (volume and/or velocity) to the transport section.
In the frst case, the inlet must have a larger fow capacity than the transport section so that the inlet
does not restrict the fow, and the inlet must be arranged such that water cannot bypass the transport
section. Ofen the design requirement for regulating the amount of fow passing into the transport
section is associated with the entrance to a decant system (Section 9.7.4.1) or to a spillway (Section
9.7.4.2). In these cases, a primary design criterion is the relationship between the height of water
above the controlling elevation of the inlet, referred to as head H and the discharge Q. Typical head-
discharge curves for decant-conduit and spillway systems are presented in Figure 9.24.
As shown in Figures 9.25 and 9.26, most inlets have a form of weir control, and the basic equation for
the associated fow is:
Q = CLH
3/2
(9-13)
where:
Q = discharge or rate of fow (cfs)
C = a coefcient that depends on the shape of the channel entrance and
the head on the structure
L = the efective length of the entrance crest (f)
H
= the total head on the entrance crest (f)
Equation 9-13 can be used to determine fow rates over both broad- and sharp-crested weirs. Broad-
crested weirs are characterized by small H/B ratios, where H and B are as defned in Figure 9.25a. An
unrestricted channel entrance, such as shown in Figure 9.25c, can sometimes be described as a broad-
crested weir. The discharge coefcient C for broad-crested weirs is a function of the weir geometry and
< PREVIOUS VIEW
9-64
Chapter 9
MAY 2009
approach conditions, as discussed in texts on hydraulics such as Brater et al. (1996), Chow (1959), and
Henderson (1969). The value of the weir coefcient is generally about 3.0, but varies based on the weir
confguration. Hydraulic analysis of the discharge and depth of fow through a broad-crested weir,
including charts for determining fow profles, is presented in USDA Technical Release 39 (1968).
Discharge coefcients for sharp-crested weirs, such as the decant inlets shown in Figure 9.26, are
related to the weir height and the hydraulic head. The value of C can be determined from the revised
Rehbock formula (Brater et al., 1996), which has been verifed in tests performed by the USBR:
C = 3.22 + 0.44 ( H/ P) (9-14)
where:
H = hydraulic head (f)
P = height of weir (f)
A range of weir coefcients is presented in Figure 9.25b for an ogee crest. The designer is referred
to the USBR (1987a) for a more detailed presentation of discharge coefcients for ogee crests. The
USACE (1990b) presents design methods and discharge coefcients for elliptical (ogee) spillways and
spillways with free outfall.
With increasing head on a decant or spillway conduit, the weir will become submerged and orifce
fow conditions may govern the head-discharge relationship, or pressure fow may prevail depend-
ing on the transport section. Under orifce fow, the control section is located within the conduit or
throat of the transition between the inlet and transport section, below the crest of the vertical intake
shown in Figure 9.26a. USBR (1987a) presents procedures for determining the head-discharge rela-
FIGURE 9.24 TYPICAL HEAD-DISCHARGE CURVES FOR HYDRAULIC SYSTEMS
FIGURE 9.24 TYPICAL HEAD-DISCHARGE CURVES FOR HYDRAULIC SYSTEMS
DECANT - CONDUIT
SYSTEM
SPILLWAY SYSTEM
DISCHARGE OR RATE OF FLOW, Q
H
E
A
D
O
R
H
E
I
G
H
T
O
F
W
A
T
E
R
A
B
O
V
E
C
O
N
T
R
O
L
L
I
N
G
E
L
E
V
A
T
I
O
N
,
H
FIGURE 9.24 TYPICAL HEAD-DISCHARGE CURVES FOR HYDRAULIC SYSTEMS
< PREVIOUS VIEW
9-65
Hydrology and Hydraulics
MAY 2009
FIGURE 9.25 SPILLWAY APPROACH CHANNELS
FIGURE 9.25 SPILLWAY APPROACH CHANNELS
B
H
A
A
9.25c UNRESTRICTED CHANNEL ENTRANCE
FOR STEEP CHANNEL SLOPE Q = C L H
1.5
WHERE C = 3.0
FOR MILD CHANNEL SLOPE Q DEPENDS ON CHANNEL GEOMETRY AND
MAY BE CONTROLLED BY DOWNSTREAM
CONDITIONS
9.25a RECTANGULAR WEIR
RECTANGULAR FLOW SECTION
SECTION A - A
L
TRAPEZOIDAL FLOW SECTION
L ~
SECTION A - A
9.25b OGEE CREST DAM
H
Q = C L H
1.5
C = WEIR COEFFICIENT
L = LENGTH OF CREST
B = BREADTH OF CREST
P = HEIGHT OF WEIR
H = HEAD ON CREST
Q = C L H
1.5
C = 3.80 TO 3.95
L = LENGTH OF CREST
H = HEAD ON CREST
H
P
B
FIGURE 9.25 SPILLWAY APPROACH CHANNELS
< PREVIOUS VIEW
9-66
Chapter 9
MAY 2009 FIGURE 9.26 DROP-INLET DECANT SYSTEMS
9.26a CIRCULAR DROP INLET PIPE
Q = CDH
1.5
C = WEIR COEFFICIENT
D = DIAMETER OF DROP-INLET PIPE
H = HEAD AT TOP OF DROP-INLET PIPE
9.26b RECTANGULAR DROP INLET
W
L
Q = 2C(W+L)H
1.5
C = WEIR COEFFICIENT
W = WIDTH OF DROP INLET
L = LENGTH OF DROP INLET
H = HEAD AT TOP OF DROP INLET
NOTE: 1. FLOW IS SHOWN ON FOUR SIDES OF DROP
INLET. IF PARTIALLY OBSTRUCTED, THE
ACTUAL FLOW LENGTH SHOULD BE USED.
2. THE INLET CONTROL FLOWS ARE ONLY
APPLICABLE PRIOR TO PRESSURIZED
PIPE FLOW.
D
TOP OF DROP INLET PIPE
H
TWO-DIMENSIONAL CROSS SECTION
OF CIRCULAR DROP-INLET PIPE
FIGURE 9.26 DROP-INLET DECANT SYSTEMS
tionship for conduit spillways. For small-diameter conduits, as typically found with many decant
structures, the head-discharge relationship under orifce fow control can be estimated as follows:
Q = CA( 2g H)
0.5
(9-15)
where:
C = coefcient of discharge (dependent on confguration of conduit opening)
A = cross-sectional area of conduit (f
2
)
g = gravitational acceleration (f/sec
2
)
H = total head on the conduit entrance (f)
FIGURE 9.26 DROP-INLET DECANT SYSTEMS
< PREVIOUS VIEW
9-67
Hydrology and Hydraulics
MAY 2009
For orifce fow, the coefcient of discharge C for vertical, circular decant systems is typically
about 0.6 for square-edged conditions. Brater et al. (1996) and Haan and Barfeld (1978) present
the results of studies of varying orifce conditions and associated coefcient of discharges for
various inlet shapes.
9.7.2.2 Transport Sections
The transport or conveyance section of a hydraulic system conveys fow from the inlet to the outlet.
The two basic types of fow in transport sections are:
1. Open-channel fow Gravity fow of water through an open channel where the
fow depth and velocity depend upon the cross section, surface material and chan-
nel slope, and possibly also upon interrelationships with the inlet and outlet. Open-
channel fow is typical for spillway channels, diversion ditches and natural streams,
but may also apply to a less-than-full culvert or a decant conduit.
2. Pressure fow Pressure fow through a closed conduit or pipe where the fow
capacity depends upon the inlet and outlet conditions, the pressure head on the con-
duit and the conduit size and material. This type of transport section is common to
decant systems.
Flow within a spillway system, as illustrated in Figure 9.25a and 9.25b, where the transport section
is simply the free-fall fow over a weir or high-velocity fow over an ogee crest, are not generally
encountered at coal refuse disposal facilities. The designer is referred to the USBR (1987a) and the
USACE (1990b) for this type of hydraulic design
9.7.2.2.1 Open-Channel Flow
Figure 9.27 illustrates several types of open-channel cross sections. The fow rate Q for each channel
is the product of the average fow velocity V and fow area A:
Q = VA (9-16)
where:
Q = fow rate (cfs)
V = average velocity of fow (f/sec)
A = cross-sectional area of fow (f
2
)
The fow area is a function of the fow depth which, along with the required freeboard, establishes
the required height of the channel. Velocity is important because it is directly related to the potential
for erosion, cavitation and energy dissipation. Normally, the required fow area is determined frst,
because the channel geometry is ofen controlled by the refuse disposal facility and site confgura-
tions and the desire to have uniform channel cross sections to simplify construction.
Methods for the analysis of open-channel fow and procedures for calculating depth of fow at any
point along a channel are available in many texts, including Brater et al. (1996), Henderson (1969),
and Chow (1959). The FHWA (1961) presents tables and charts for determination of open-channel
fow parameters in prismatic channels. Computer sofware for determining steady-state and tran-
sient fow conditions in open channels is also readily available. Accordingly, the following discussion
of design and analysis considerations relates to fow regulation and the identifcation of specifc con-
ditions or features encountered at coal refuse disposal facilities.
< PREVIOUS VIEW
9-68
Chapter 9
MAY 2009
Four possible fow (and depth) conditions must be considered in the design of an open channel.
These are: (1) critical depth, (2) depth at subcritical fow, (3) depth at supercritical fow, and (4) normal
depth. The point at which open-channel fow passes through the critical depth is a regulating condi-
tion of fow. Hereafer, the point where this occurs is referred to as a control point or location regulat-
ing the fow rate. The depth of fow at a control point may also be referred to as the control depth.
Several examples illustrating these basic fow conditions are presented in Figure 9.28. Additional
explanation of these conditions is provided in the following:
1. Critical depth The critical depth of fow in an open channel is the depth at which
fow occurs with the minimum specifc energy, defned as the minimum depth y
plus the velocity head (V
2
/2g) for the channel. At the critical state of fow, the veloc-
ity head is equal to half the hydraulic depth D, as defned in Figure 9.27. Chow
(1959) and FHWA (1961) provide detailed discussion of the critical depth and mini-
mum specifc energy. For design, the critical depth is normally calculated for com-
parison to the actual depth to determine if the actual fow condition is subcritical
or supercritical. Figure 9.29 presents curves for determining the critical depth for
trapezoidal and circular channels. Except at fow control points, design at or close to
critical depth should be avoided in order to prevent the occurrence of unstable fow
regimes.
2. Depth at subcritical fow For subcritical fow the channel slope is milder than that
associated with critical fow, and the fow velocity is relatively low. Channel slopes
with subcritical fow are called subcritical slopes. For subcritical fow, the control
point will be downstream.
FIGURE 9.27 OPEN-CHANNEL CROSS SECTIONS
A
WP
CONDUIT
A
WP
TRIANGULAR
A
WP
RECTANGULAR
A
WP
TRAPEZOIDAL
HYDRAULIC RADIUS (R) =
AREA (A)
WETTED PERIMETER (WP)
HYDRAULIC DEPTH (D) =
AREA (A)
TOPWIDTH (T)
T T
T
T
FIGURE 9.27 OPEN-CHANNEL CROSS SECTIONS
FIGURE 9.27 OPEN-CHANNEL CROSS SECTIONS
< PREVIOUS VIEW
9-69
Hydrology and Hydraulics
MAY 2009
3. Depth at supercritical fow For supercritical fow the channel slope is steeper than
that associated with critical fow and the fow velocity is relatively high. Channel
slopes with supercritical fow are called supercritical slopes. For supercritical fow,
the control point will be upstream. If it is important that water be carried away
from the inlet section without a possible back up, it may be desirable to design for
a supercritical fow condition for a portion or all of the length of the channel. How-
ever, this may result in high fow velocities, and special atention will have to be
given to factors such as erosion, fow at direction changes and energy dissipation at
the discharge point. Most coal refuse disposal facilities have a large elevation difer-
ence between the embankment crest and downstream toe, and this generally results
in some section of the channel between these two points having a supercritical fow
condition.
4. Normal depth Normal fow depth occurs when the energy increase during eleva-
tion drop is exactly balanced by friction losses along a channel. When this occurs, the
fow depth and velocity in a channel of constant cross section and slope will remain
constant. Open-channel fow naturally tends toward the normal depth condition,
but for short channel lengths it ofen does not reach the normal condition. Because
normal depth is easily determined, it is the fow condition most ofen used during
preliminary hydraulic analyses to establish whether or not the selected channel slope
is feasible (as limited by geotechnical or structural conditions).
The Manning equations are used to estimate the velocity and depth of fow for a given fow rate and
channel confguration. These equations are:
Q =
1.49
AR
0.67
S
0.5
(9-17)
n
V =
1.49
R
0.67
S
0.5
(9-18)
n
where:
Q = fow rate (cfs)
V = average velocity of fow (f/sec)
A = cross-sectional area of fow (f
2
)
R = hydraulic radius (f)
S = slope of the channel (f/f)
n = Mannings coefcient of channel roughness
The depth determined from the Manning formula is normal depth. The term R is the hydraulic radius
of the fow and is equal to the area of fow A divided by the weted perimeter WP. The weted perime-
ter for several types of cross sections is shown in Figure 9.27. Values of Mannings coefcient of chan-
nel roughness n are listed in Table 9.13. An extensive presentation of n values for a wide variety of
channels is provided in Chow (1959) and FHWA (1961). Depending upon the efective roughness of
the surface, n can range from as high as 0.2 for channels with dense brush growth and many obstruc-
tions to as low as 0.01 for smooth-fnish, concrete-lined channels.
Since the cross-sectional area A and hydraulic radius R are both functions of fow depth, an iterative
procedure is required to determine a normal depth that satisfes the Manning equations. Commer-
cially available sofware can be used to determine fow depth, velocity, critical depth and slope, and
various open-channel design parameters (e.g., HEC-RAS and privately marketed similar programs).
< PREVIOUS VIEW
9-70
Chapter 9
MAY 2009
FIGURE 9.28 CHANNEL FLOW CONDITIONS
9.28a STEEP SLOPE (SUPERCRITICAL FLOW)
D2
D
1
FIGURE 9.28 CHANNEL FLOW CONDITIONS
9.28b MILD SLOPE (SUBCRITICAL FLOW)
1. D2 IS CONTROLLED BY DOWNSTREAM CONDITIONS
2. D1 IS CONTROLLED BY D2
3. D1 APPROACHES NORMAL DEPTH
D1
D2
CRITICAL DEPTH
D1
D2
9.28c MILD SLOPE WITH STEEP DROP
1. CONTROL IS AT THE POINT WHERE CRITICAL DEPTH (Dc) OCCURS
2. Dc IS A FUNCTION OF CHANNEL SHAPE AND FLOW (Q)
3. D2 AND D1 ARE CONTROLLED BY Dc
Dc
1. D1 IS CONTROLLED BY INLET CONDITIONS
2. D2 IS CONTROLLED BY D1
3. D2 APPROACHES NORMAL DEPTH
FIGURE 9.28 CHANNEL FLOW CONDITIONS
< PREVIOUS VIEW
9-71
Hydrology and Hydraulics
MAY 2009
Many commercially manufactured erosion control products have been developed to resist potential
erosion. These products include vegetation control mats, interlocking concrete blocks, concrete mat
systems, cellular confnement mats, etc. The Mannings n value for these erosion control products is
typically provided by the manufacturer.
An experienced designer can determine if the normal depth of fow from the Manning equations is suf-
fcient for determining fow conditions for fnal design. Detailed analyses of water surface profles may
be necessary, particularly when depths associated with changes in channel confguration, erosion pro-
tection and slope may be required. In these situations, the change in water surface elevation (water sur-
face profle) due to velocity head and channel losses should be determined. The normal depth does not
adequately describe conditions at the ends or transitional zones in the transport section. For example,
where a channel with supercritical slope changes to a short length of channel with subcritical slope (e.g.,
benches on an embankment face), the designer must determine if the depth in the subcritical section
will create a hydraulic jump, and calculation of the fow profle along the length of the open channel
ofen must be made. Another instance where fow profles must be determined is when establishing a
spillway rating curve (stage-discharge curve) for an approach channel between the impoundment and
an open-channel spillway weir (control section) or when subcritical fow is present downstream of or
within the spillway channel. Procedures for analyzing such channel sections are presented in many
references on open-channel fow, including Brater et al. (1996), Henderson (1969) and Chow (1959).
The FHWA (2006), USBR (1987a) and the USDA (1956) present methods and charts for determining the
depths and other parameters associated with hydraulic jumps in open channels. If a hydraulic jump
occurs, the channel depth must be sufcient to contain the sequent fow depth that will develop.
FIGURE 9.29 CRITICAL DEPTH FOR TRAPEZOIDAL AND CIRCULAR SECTIONS
C
I
R
C
U
L
A
R
0.0001 0.001 0.01 0.1 1 10
0.001 0.01 0.1 1 10 100
0.01
0.02
0.03
0.04
0.06
0.1
0.2
0.3
0.4
0.6
1
2
3
4
6
10
1
z
b
y
y
d
o
z = 0.5
z = 1.0
z = 1.5
z = 2.0
z = 2.5
z = 3.0
z = 4.0
FOR CIRCULAR SECTIONS
z
=
0
(
R
E
C
T
A
N
G
U
L
A
R
)
C
IR
C
U
L
A
R
y
/
b
A
N
D
y
/
d
Q
d
o
2.5
g
FOR TRAPEZOIDAL SECTIONS
Q
b g
2.5
o
FIGURE 9.29 CRITICAL DEPTH FOR TRAPEZOIDAL AND CIRCULAR SECTIONS
FIGURE 9.29 CRITICAL DEPTH FOR TRAPEZOIDAL AND CIRCULAR SECTIONS
< PREVIOUS VIEW
9-72
Chapter 9
MAY 2009
TABLE 9.13 MANNINGS COEFFICIENTS OF CHANNEL ROUGHNESS
Constructed Channel Condition
Values of n
Minimum Maximum Average
Earth channels, straight and uniform 0.017 0.025 0.022
Dredged earth channels 0.025 0.033 0.028
Rock channels, straight and uniform 0.025 0.035 0.033
Rock channels, jagged and irregular 0.035 0.045 0.045
Concrete lined, regular fnish 0.012 0.018 0.014
Neat cement lined 0.010 0.013 0.011
Grouted rubble paving 0.017 0.030 0.025
Corrugated metal 0.023 0.025 0.024
Natural Channel Condition Value of n
Smoothest natural earth channels, free from growth with straight alignment 0.017
Smooth natural earth channels, free from growth, little curvature 0.020
Average, well-constructed, moderate-sized earth channels 0.0225
Small earth channels in good condition or large earth channels with some growth on banks or
scattered cobbles in bed
0.025
Earth channels with considerable growth, natural streams with good alignment and fairly constant
section, or large foodway channels well maintained
0.030
Earth channels considerably covered with small growth or cleared but not continuously maintained
foodways
0.035
Mountain streams in clean loose cobbles, rivers with variable cross section and some vegetation
growing in banks, or earth channels with thick aquatic growths
0.050
Rivers with fairly straight alignment and cross section, badly obstructed by small trees, very little
underbrush or aquatic growth
0.075
Rivers with irregular alignment and cross section, moderately obstructed by small trees and
underbrush
0.100
Rivers with fairly regular alignment and cross section, heavily obstructed by small trees and
underbrush
0.100
Rivers with irregular alignment and cross section, covered with growth of virgin timber and
occasional dense patches of bushes and small trees, some logs and dead fallen trees
0.125
Rivers with very irregular alignment and cross section, many roots, trees, large logs, and other
drift on bottom, trees continually falling into channel due to bank caving
0.200
(USBR, 1987a)
9.7.2.2.2 Pressure Flow
Closed conduits are used as the transport section for two types of hydraulic systems. When the fow
enters the conduit with a nearly horizontal approach, the system is called a culvert. When the fow
enters the conduit through a steeply inclined or vertical drop inlet, the system is called a decant.
These two types of hydraulic systems are illustrated in Figure 9.30.
When fowing full, both types of closed-conduit systems behave in essentially the same manner,
and their capacity can be determined from pipe fow analyses. However, the systems difer sig-
nifcantly in behavior during conditions leading up to the full-fow condition. The following
paragraphs discuss the hydraulics of culvert systems in detail. The hydraulic behavior of decant
systems is then presented with particular emphasis on the diferences between decant and cul-
vert systems.
< PREVIOUS VIEW
9-73
Hydrology and Hydraulics
MAY 2009
Culvert Conduits
A culvert is a pipe placed beneath an embankment with an entrance that is horizontal or slightly
inclined to the direction of fow. A culvert can be prefabricated or cast-in-place and the cross section
can be round, square, rectangular, oval or arched. Factors that control the fow in a culvert are slope,
size, shape, length and roughness (material) and inlet and outlet confguration. The combined efects
of these factors determine the hydraulic control location, which in turn determines whether the cul-
vert fows partly or completely full and establishes the head-discharge relationship. As with open-
channel fow, the slope of the culvert may be mild or steep, resulting in subcritical or supercritical
fow conditions. For either slope condition, the control point may be either at the inlet or the outlet,
depending on the system geometry and the upstream and downstream head conditions. Figures 9.31
and 9.32 illustrate the various factors that afect fow in culverts.
The control point for a culvert on a mild slope fowing partly full will usually be at the outlet if the
inlet is not submerged. The fow ordinarily will be subcritical, and the discharge may be predicted
according to the open-channel fow procedures discussed above. If the outlet discharges freely, the
fow at that point will pass through critical depth (Figure 9.31a). As the inlet becomes submerged,
the control point moves downstream within the culvert. The fow could be supercritical just inside
the culvert if limited submergence is sustained (H/ D ~ 1.2). As submergence of the entrance increases
(H/ D > 1.2, Figure 9.31b), or the culvert is sufciently long, or the elevation of the downstream back-
water is sufciently high, full pipe fow occurs with control at the downstream end of the conduit.
FIGURE 9.30 CLOSED-CONDUIT HYDRAULIC SYSTEMS
FIGURE 9.30 CLOSED-CONDUIT HYDRAULIC SYSTEMS
SEEPAGE DIAPHRAGM (OUTLET NOT SHOWN)
CONDUIT
INLET WITH NEAR
HORIZONTAL FLOW
9.30a CULVERT SYSTEM
CONDUIT
SEEPAGE DIAPHRAGM (OUTLET NOT SHOWN)
DROP INLET WITH
NEAR VERTICAL FLOW
9.30b DECANT SYSTEM
FIGURE 9.30 CLOSED-CONDUIT HYDRAULIC SYSTEMS
< PREVIOUS VIEW
9-74
Chapter 9
MAY 2009
Most coal refuse disposal facility decant systems are less that 36 inches in diameter with a length of a
few hundred to more than 1,000 feet and are thus sufciently long (Chow, 1959) that there is full pipe
fow in the conduit and free discharge from the outlet. If the elevation of the downstream backwater
is above the critical depth elevation, the depth associated with the backwater may control the fow in
the culvert. One of these conditions typically occurs, such that culverts fow full for their entire length
(Figure 9.31b).
When a culvert is on a steep slope and the inlet is not submerged, the fow is controlled by critical
depth at the inlet (Figure 9.32a), and open-channel fow at supercritical velocity will occur through-
out the culvert. Afer the inlet has been submerged (H exceeds about 1.2D), it is still possible to have
open-channel fow at supercritical velocities in the culvert if the control point remains at the inlet
(Figure 9.32b). In this case, discharge is governed by orifce fow at the inlet, leading to formation of
a fow contraction at the top of the culvert entrance and aeration over the remaining culvert length.
As the head at a submerged inlet increases, friction or local disturbances may reduce the fow veloc-
ity, causing the culvert to fow full near the outlet. This may seal the downstream end of the pipe,
even though an orifce fow contraction tends to occur at the inlet. The associated high-velocity fow
will tend to carry away the air trapped at the top of the culvert, thus reducing the internal pressure
to less than atmospheric pressure. This can lead to damage to the culvert from cavitation (Section
9.7.3.3). However, if the entrance is shaped to eliminate the inlet fow contraction, the culvert will
start to fow full near the inlet, afer which the full fow zone will extend rapidly toward the outlet.
The efect of the full fow condition will be a draf tube action (similar to siphonic action) that will
FIGURE 9.31 TYPICAL FLOW CONDITIONS CULVERT CONDUITS ON MILD SLOPES
FIGURE 9.31 TYPICAL FLOW CONDITIONS - CULVERT CONDUITS ON MILD SLOPES
PART FULL FLOW - INLET NOT SUBMERGED
9.31a MILD SLOPE - SUBCRITICAL FLOW: CONTROL AT
CRITICAL DEPTH AT OUTLET
H
D
< 1.2
S < CRITICAL
H
d
c
D
FULL FLOW - INLET SUBMERGED
9.31b MILD SLOPE - CONTROL AT OUTLET: EFFECTIVE HEAD = H
t
- LOSSES
S < CRITICAL
H
t
FOR DISCHARGE
AFFECTED BY DOWN-
STREAM BACKWATER
H
t
FOR FREE
DISCHARGE
SHARP-EDGED OR
ROUNDED INLET
H
D
> 1.2
H
D
FIGURE 9.31 TYPICAL FLOW CONDITIONS CULVERT CONDUITS ON MILD SLOPES
< PREVIOUS VIEW
9-75
Hydrology and Hydraulics
MAY 2009
FIGURE 9.32 TYPICAL FLOW CONDITIONS - CULVERT CONDUITS ON STEEP SLOPES
PART FULL FLOW-INLET NOT SUBMERGED
PART FULL FLOW-INLET SUBMERGED
FULL FLOW-INLET SUBMERGED
9.32b STEEP SLOPE - SUPERCRITICAL FLOW: ORIFICE FLOW CONTROL AT INLET
SHARP-EDGED INLET
H
D
~ 1.2
S > CRITICAL
D
d
c
9.32a STEEP SLOPE - SUPERCRITICAL FLOW: CONTROL AT CRITICAL DEPTH AT INLET
H
D
< 1.2
S > CRITICAL
D
d
c
H
ROUNDED INLET
H
D
= 1.2 TO 1.5
9.32c STEEP SLOPE - SUPERCRITICAL PULSATING SLUG FLOW: CONTROL
SWITCHING BETWEEN INLET AND SECTION WITHIN THE CONDUIT
S > CRITICAL
D
H
9.32d STEEP SLOPE CONTROL AT OUTLET EFFECTIVE HEAD = H
t
- LOSSES
H
D
> 1.5
S > CRITICAL
H
t
FOR DISCHARGE
AFFECTED BY
DOWNSTREAM
BACKWATER
H
t
ROUNDED INLET
D
H
H
FIGURE 9.32 TYPICAL FLOW CONDITIONS CULVERT CONDUITS ON STEEP SLOPES
FIGURE 9.32 TYPICAL FLOW CONDITIONS CULVERT CONDUITS ON STEEP SLOPES
< PREVIOUS VIEW
9-76
Chapter 9
MAY 2009
increase the discharge. The increased discharge will cause a deep drawdown just upstream from the
inlet, and a vortex will form, allowing air to enter the culvert, breaking the draf tube action. This
immediately reduces the discharge and causes a return to orifce control at the inlet. The full fow
action will begin again and the cycle will be repeated. This alternating action will cause a pulsating
slug fow phenomenon (Figure 9.32c). When the storage elevation and culvert dimensions are such
that the H/ D ratio exceeds about 1.5, the inlet drawdown will be insufcient to allow air to enter the
culvert and steady full fow will prevail (Figure 9.32d).
The fow velocity during the full-fow conditions illustrated by Figures 9.31b and 9.32d can be deter-
mined from the Bernoulli equation:
V
2
= H
t
losses
(9-19)
2g
where:
V = fow velocity (f/sec)
g = acceleration of gravity (f/sec
2
)
H
t
= head (elevation diference) between the impoundment surface and the
point of discharge (f)
The term losses encompasses all fow-reducing conditions associated with the inlet, culvert
geometry and culvert friction. These losses can be related to the velocity by:
losses
=
V
2
( f
L
+
K
L
) (9-20)
2g D
where:
f = friction factor for the culvert material and fow condition in the culvert
L = length of the culvert (f)
D = diameter of the culvert (f)
K
L
= sum of head losses associated with the inlet, valves, constrictions
and directional changes (f)
For most fow conditions and circular culvert materials, the friction factor ( f ) can be determined from
the following equation:
f = 185 n
2
/ d
1/3
(9-21)
where:
n = Mannings roughness coefcient for culvert
d = pipe diameter (f)
By substituting terms from Equation 9-20 into Equation 9-19, the relationship between H
t
and V becomes:
H
t
=
V
2
( f
L
+
K
L
+ 1) (9-22)
2g
D
< PREVIOUS VIEW
9-77
Hydrology and Hydraulics
MAY 2009
Typical head loss coefcients K
L
are tabulated in Table 9.14. Afer all of the loss parameters have been
determined, Equation 9-22 can easily be solved.
However, Equation 9-22 is only appropriate when dealing with water at normal temperatures. The
friction factor approach is superior when working with smooth pipes and/or large values of Reyn-
olds number. For materials and pipe sizes commonly used, the above equation will give acceptable
results. When unusual surfaces or very large pipe sizes are involved or when very long conduits with
large energy losses are used, an experienced hydraulic engineer should be consulted.
The geometry of an inlet is important to achieving discharge efciency. Until the headwater surface is
well above the culvert inlet, a square-edged inlet produces an inlet fow contraction without greatly
reducing the discharge capacity. Flow contractions can also be formed (but at reduced discharge
capacity) by a projecting inlet, a mitered inlet, an inlet orifce ring, or a top curtain wall closure. These
inlet confgurations are shown in Figure 9.33.
TABLE 9.14 HEAD LOSS COEFFICIENTS FOR CONDUITS
K
L
for Inlets (USBR, 1987a)
Inlet Type K
L
Fully rounded entrances (r / D 0.15) fush with vertical walls 0.10
Square edge entrances fush with vertical walls 0.50
Socket-ended concrete pipe fush with vertical walls 0.15
Projecting concrete pipe with socket ends 0.20
Projecting smooth-wall or corrugated pipe 0.85
Gate in thin wall unsuppressed contraction 1.50
Gate in thin wall corners rounded 0.50
K
L
for Directional Changes (Linsley and Franzini, 1972)
Radius of Bend/Pipe Diameter
Angle of Bend
90 45 22.5
1 K
L
0.50 0.37 0.25
2 K
L
0.30 0.22 0.15
4 K
L
0.25 0.19 0.12
6 K
L
0.15 0.11 0.08
8 K
L
0.15 0.11 0.08
K
L
for Valves and Fittings (Linsley and Franzini, 1972)
Type of Valve or Fitting K
L
Butterfy valve (wide open) 0.2
Gate valve (wide open) 0.2
Gate valve (half open) 5.6
Return bend 2.2
Standard tee 1.8
Standard 90 elbow 0.9
< PREVIOUS VIEW
9-78
Chapter 9
MAY 2009
If it is desired that a culvert fow full at increased headwater elevations, the control point during
maximum fow will be at the outlet and the geometry of the inlet will be less signifcant. For this case,
the inlet should be shaped so as to minimize the inlet fow contraction, thereby increasing the ten-
dency for full fow for all conditions except when the inlet is not submerged. Streamlining the shape
of the inlet will streamline the fow and reduce inlet head losses. The tendency to develop cavitation
pressures will also be reduced. Rounded entrances or a gradually tapering transition in size to the
basic culvert dimension will generally provide the desired streamlining.
9.33b PROJECTING INLET
9.33c MITERED INLET
FIGURE 9.33 CULVERT INLET CONFIGURATIONS
B
B
9.33e TOP CURTAIN WALL INLET
SECTION B - B
A
A
9.33d INLET ORIFICE RING
SECTION A - A
9.33a SQUARE-EDGED INLET
FIGURE 9.33 CULVERT INLET CONFIGURATIONS
FIGURE 9.33 CULVERT INLET CONFIGURATIONS
< PREVIOUS VIEW
9-79
Hydrology and Hydraulics
MAY 2009
It is obvious from the previous discussions and related tables and fgures that culvert inlets may
vary with respect to approach conditions, entrance arrangements, and cross-sectional shapes. For
example: (1) the approach to the inlet may or may not be a well defned channel with or without con-
structed wing walls, (2) the culvert may be fush with or protrude past the upstream embankment
surface or constructed headwall, (3) the face may be square, beveled or rounded, and (4) the cross
section may be round, square, rectangular or arched. All such variations have a marked infuence on
culvert performance as they afect weir or orifce discharge, inlet fow contractions and head losses
during fow entrance to the culvert.
For coal refuse disposal facilities, circular culverts with vertical square-edged headwalls and fared
wingwalls are sometimes encountered. The hydraulic design for these installations is discussed in
detail by the USBR (1987a). Also, procedures and charts for design of these types of culverts are pre-
sented in FHWA (2005b).
The purpose of hydraulic analysis of a closed-conduit system is similar to that for an open-channel
spillway, which is to determine the relationship between the elevation of the impounded storage
(headwater) and the rate of outfow or discharge from the system. The head-discharge relationship
for culvert-type conduit systems may be controlled by weir, orifce and pressure fow conditions
depending on the upstream head, as illustrated in Figures 9.31 and 9.32.
Drop Inlet Conditions
Closed-conduit fow in drop-inlet decant systems is similar to that for culvert systems, as discussed
above. The relationship between the elevation of the impounded storage and the rate of discharge
during these types of fow is indicated by the curves shown in Figure 9.34. For the drop-inlet systems
illustrated in Figure 9.34, curve A results from crest-controlled (also referred to as weir-controlled)
fow, as afected by the inlet geometry and water level conditions. The inlet geometry can range from
a sharp-edged, vertically placed circular pipe to an elaborately formed concrete ogee called a morn-
ing glory spillway.
Curve B results from orifce-controlled fow. The fow control point for curve A is at the crest, as illus-
trated in Figure 9.34a. The fow control point for curve B is at the entrance to the transport section of
the conduit, as illustrated in Figure 9.34b. The curve B fow condition occurs when the open-channel
capacity of the transport section and the discharge capacity of the outlet are both greater than the
capacity of the inlet orifce. This results in a backup of water in the drop tube and drowning of the
inlet weir.
The curve C fow condition is associated with full fow, generally controlled by the transport sec-
tion. Analysis for the curve C condition is the same as for a culvert system under full fow (Equation
9-22) except that an additional head loss term related to the drop inlet should be included. The loss
coefcient K
L
for the transition from the drop inlet to the transport section can range from 0.2 for an
unobstructed, specially-formed structure to 2.0 for a debris-clogged, square-edged turn. A value of
1.0 is normally satisfactory for design.
To maximize the discharge capacity of decant systems under low heads, conical-shaped inlets that
extend the crest control fow (curve A) have been employed. Other inlet shapes that incorporate an
orifce ring have been used to maintain orifce fow (curve B) and to prevent the transport pipe section
from fowing full or under pressure.
As a guide for assisting the designer in the evaluation of important secondary characteristics afect-
ing closed-conduit fow, a list of Manual sections and supplemental references for secondary hydrau-
lic design issues is provided in Table 9.15.
< PREVIOUS VIEW
9-80
Chapter 9
MAY 2009
FIGURE 9.34 FLOW AND DISCHARGE CHARACTERISTICS OF DECANT-INLET,
CLOSED-CONDUIT SYSTEM
H
WEIR CONTROL AT CREST
TRANSPORT SECTION OF CONDUIT
OPEN CHANNEL FLOW
9.34a CURVE A - CREST CONTROL
OUTFLOW, Q
A
B
C
POINT OF CHANGE FROM ORIFICE
TO FULL FLOW CONTROL
POINT OF CHANGE FROM
CREST TO ORIFICE CONTROL
H
E
A
D
O
N
I
N
L
E
T
,
H
H
CREST
FOR DOWNSTREAM
BACKWATER CONDITION
FOR FREE
DISCHARGE
CONDITION
9.34c CURVE C - FULL FLOW CONTROL
FULL FLOW
IN CONDUIT
H
ORIFICE CONTROL AT ENTRANCE
TO TRANSPORT SECTION
OPEN CHANNEL FLOW
9.34b CURVE B - ORIFICE CONTROL
CREST
FIGURE 9.34 FLOW AND DISCHARGE CHARACTERISTICS OF DECANT-INLET,
CLOSED-CONDUIT SYSTEM
FIGURE 9.34 FLOW AND DISCHARGE CHARACTERISTICS OF DECANT-INLET,
CLOSED-CONDUIT SYSTEM
< PREVIOUS VIEW
9-81
Hydrology and Hydraulics
MAY 2009
9.7.2.3 Outlets
There are two basic requirements that typically dictate the size, confguration and location of the
outlet section of the hydraulic system:
1. If the outlet is intended to control the relationship between headwater storage and
outfow for the entire hydraulic system, the outlet size and confguration must be
designed specifcally for this purpose.
2. The confguration and location of the outlet must be such that discharged water will
not create a hazardous condition (e.g., erosion of the channel or embankment toe)
and the energy of the outfow will be dissipated. The potential for localized damage
must be controlled to a level acceptable to the owner and to regulatory agencies.
Whether the frst requirement applies depends upon the design choices for the inlet and transport
sections of the system, as previously discussed. If applicable, the outlet size and confguration must
be designed to control the required discharge, generally by weir or orifce control. Evaluation of this
discharge was discussed in Section 9.7.2.1 and is not discussed further herein.
The second requirement has more general signifcance and necessarily requires considerable plan-
ning at the earliest stages of design so that the outlet of the hydraulic system will not be incompatible
with any other portion of the coal refuse disposal facility. Criteria to be considered in determining the
degree of efort required for construction of an outlet system normally include:
Under no circumstance should the outlet location or confguration allow fow to dis-
charge beyond the outlet in a manner that could contribute to embankment failure,
endangering lives or major downstream property.
Situations where signifcant localized damage could occur during moderate storms
should be avoided. Exceptions are when the cost of constructing a damage-safe
outlet far exceeds the cost of occasional maintenance. Ofen the 100-year storm (Sec-
tion 9.5) is considered as the basis for design when it is unlikely that lives or major
downstream property could be endangered.
For very large storms, such as those approaching the PMF (Section 9.5), it is generally
not required that the outlet totally prevent downstream damage due to discharge, if
the frst criterion above is satisfed. Under such severe storms, signifcant downstream
damage (such as erosion and roadway overtopping) can be expected to occur even if
the hydraulic system of the disposal facility satisfes all requirements of good design.
TABLE 9.15 REFERENCES FOR SECONDARY HYDRAULIC DESIGN ISSUES
Item Manual Section Supplemental References
External Embankment Pressure 6.6.6.2 FEMA (2005a)
Internal Water Pressure 9.7.2.2 Brater et al. (1996)
Thrust Forces 9.7.3.5 Brater et al. (1996)
Trash Racks 9.7.4.1 USBR (1987a)
Vortex Control 9.7.4.1 USBR (1987a)
Flow at Bends 9.7.3.4 Brater et al. (1996), USACE (1980)
Cavitation 9.7.3.3 USACE (1980), USBR (1990)
Materials Selection 6.6.6.1 FEMA (2005a)
< PREVIOUS VIEW
9-82
Chapter 9
MAY 2009
Several types of outlet conditions are illustrated in Figure 9.35. If the discharge point can be located
an adequate distance away from the overall facility, the most reliable and least expensive outlets are
those that allow the fow to discharge over erosion resistant rock or large riprap. This type of outlet
condition is most ofen used when the channel or spillway will function only infrequently because of
the limited recurrence of very large storms.
The energy dissipating structures shown in Figure 9.35 are commonly used for water-impounding earth
dams, but are normally not necessary for coal refuse disposal facilities if sufcient access is available
that periodic maintenance and repair can be performed. However, such outlets should be considered for
moderate but regular discharges or when future access could be difcult. Also, if extending the transport
section a signifcant distance beyond the downstream limits of the facility will be costly, coupling a short
transport section with an energy-dissipating structure may be a more cost-efective solution.
State and local agencies may prefer specifc methods or structures for energy dissipation at outlet
structures, and applicable regulatory guidance should be reviewed as part of design. Hydraulic
FIGURE 9.35 EXAMPLES OF CHANNEL AND CONDUIT OUTLETS
FIGURE 9.35 EXAMPLES OF CHANNEL AND CONDUIT OUTLETS
DISCHARGE WITHOUT ENERGY
DISSIPATING STRUCTURES
BOULDERS, RIPRAP OR
OTHER PROTECTION
SYSTEM
9.35b RIPRAP PROTECTION
DISCHARGE WITH ENERGY
DISSIPATING STRUCTURES
OUTFLOW CHANNEL
OR CONDUIT
9.35d PLUNGE POOL STILLING BASIN
IMPACT BLOCK OR
ROCK HILLSIDE
9.35e IMPACT BLOCK STILLING MECHANISM
OUTFLOW CHANNEL
OR CONDUIT
9.35c HYDRAULIC JUMP STILLING BASIN
SUBCRITICAL
FLOW
SUPERCRITICAL
FLOW
OUTFLOW CHANNEL
OR CONDUIT
COAL REFUSE EMBANKMENT
OUTFLOW CHUTE,
EXCAVATED SPILLWAY
CHANNEL OR CONDUIT
STREAM
9.35a FREE OUTFALL
EROSION
RESISTANT
ROCK
FIGURE 9.35 EXAMPLES OF CHANNEL AND CONDUIT OUTLETS
< PREVIOUS VIEW
9-83
Hydrology and Hydraulics
MAY 2009
design requirements for structures similar to those shown in Figure 9.35 are discussed by FHWA
(2006), USACE (1990b), USACE (1980), and USBR (1987a).
9.7.3 Special Design Considerations
The following discussion briefy emphasizes some important special considerations of hydraulic
system design and illustrates situations where special conditions must be evaluated.
9.7.3.1 Channel Freeboard
Freeboard is needed in open channels to account for minor channel irregularities, air entrain-
ment and wave action. The required freeboard is a function of the design fow velocity and fow
depth and typically varies from less than 5 percent to greater than 30 percent of the depth of
fow (Chow, 1959). Low freeboards may be appropriate for smooth, uniform channels with fow
velocities less than 8 feet per second. The following is an empirical relationship that provides the
desirable spillway channel freeboard based upon surface roughness, wave action, air bulking,
splash, and spray under supercritical conditions (USBR, 1987a):
Freeboard (f) = C
f
+ 0.025 V D
1/3
(9-23)
where:
C
f
= freeboard coefcient
V = design fow velocity (f/sec)
D = fow depth (f)
The USBR (1987a) has indicated that a value of C
f
equal to 2 is desirable for spillway channels
for dams. While Equation 9-23 has been applied to spillways with the cited fow conditions, this
formula has been used by some state agencies for the design of uniform channels ranging from
perimeter and diversion ditches to small spillway channels by adopting values for C
f
ranging from
0 (with not less than 0.3 feet of freeboard) to 1. MSHA (2007) has indicated that a value of C
f
equal
to 1 is prudent for design of perimeter ditches. Channel bends, convergence, and other geometric
changes may require greater freeboard to address variations in the fow profle. Additional discus-
sion is presented in Section 9.7.3.4.
9.7.3.2 Erosion Protection
Two methods are typically employed in the design of erosion protection linings for channels: (1) the
permissible velocity method or (2) the tractive (shear stress) force method. Under the permissible
velocity approach, the channel is assumed to be stable if the mean velocity is lower than the maxi-
mum permissible velocity for the channel materials. The tractive force approach focuses on stresses at
the channel boundary. The factors that normally cause erosion are velocity of fow (e.g., in steep chan-
nels and at intakes) and water impact at points where direction changes occur (e.g., at sharp bends in
channels, at the intersection of channels, and where free-falling water discharges from a channel or
conduit). If unchecked, erosion can result in degradation of water quality and frequent maintenance.
Excessive or uncontrolled erosion on or near an impounding embankment can result in catastrophic
failure during large storms.
The shear resistance of the soil/rock in an unlined channel or the shear resistance provided by the
channel lining will determine the stability of the channel relative to erosion. The shear force gener-
ated by water fowing through a channel section is given by:
=
R S (9-24)
< PREVIOUS VIEW
9-84
Chapter 9
MAY 2009
where:
= shear stress along the weted perimeter of fow (lb/f
2
)
d S (9-25)
where:
d
= shear stress in channel at maximum depth (lb/f
2
)
d
= maximum depth of fow in channel (f)
For botom width to depth ratios less than 4, the above equation conservatively overestimates the
shear stress for straight uniform channels. If a more refned analysis is desired, Equation 9-24 should
be applied.
The shear stress on the weted perimeter () is then compared with the permissible shear stress (
p
)
for the channel botom or lining. If the permissible shear stress is greater than or equal to the com-
puted shear stress, including a safety factor (SF), the channel and lining are considered stable. If this
condition is not met, a diferent channel confguration or lining with a higher permissible shear stress
is selected. The concept is expressed as:
p
SF (9-26)
The safety factor provides for a measure of uncertainty and conservatism for the designer. A safety
factor of 1.5 is generally recommended for the following conditions:
Critical or supercritical fows
Climatic regions where vegetation may be uneven or slow to establish
There is signifcant uncertainty regarding the design discharge
The consequences of failure are high
Permissible velocity and tractive shear force values for channel linings are presented in Table 9.16.
Determination of allowable tractive shear forces for commercially available erosion control products
should be obtained from the manufacturers. Determination of tractive shear forces for grass- and
riprap-lined channel sections is a function of the vegetation retardance class and the median riprap
size for grass- and riprap-lined channels, respectively. More detailed explanations of the design pro-
cedures for grass- and riprap-lined channels are presented below. It should be noted that the use of
grass and riprap erosion protection will generally be limited by the channel slope. For steep embank-
ments with slopes greater than 10 percent, grass or riprap channel lining has limited application.
Grouted riprap and concrete should be considered when tractive forces exceed the allowable limits
for grass or riprap channel linings. Also, commercially manufactured products have been developed
< PREVIOUS VIEW
9-85
Hydrology and Hydraulics
MAY 2009
TABLE 9.16 PERMISSIBLE VELOCITY AND TRACTIVE FORCE FOR CHANNEL LININGS
Channel Lining
Permissible Velocity
(1)
(ft/sec)
Permissible Tractive
Force
(2)
(lb/ft
2
)
Temporary Lining
Jute netting 0.45
Straw with net 1.45
Coir-double net 2.25
Coconut fber-double net 2.25
Curled wood mat 1.55
Curled wood-double net 1.75
Curled wood-high velocity 2.00
Synthetic net 2.00
Vegetative Lining
Class A 3.70
Class B 2.10
Class C 1.00
Class D 0.60
Class E 0.35
Kentucky Bluegrass, Tall Fescue
(3)
5-7
Grass Mixture, Red Canarygrass
(3)
4-5
Lespedeza Sericea, Weeping Lovegrass 2.5-3.5
Redtop, Red Fescue, Annuals
(3)
Riprap
(4)
R-1 2.5 0.25
R-2 4.5 0.50
R-3 6.5 1.00
R-4 9 2.00
R-5 11.5 3.00
R-6 13 4.00
R-7 14.5 5.00
R-8 17 8.00
Gabions 22 8.35
Reno Mattress
6-10 inches thick 12 8.35
10-12 inches thick 15 8.35
12-18 inches thick 18 8.35
Note: 1. USACE (1994), PADEP (2000); permissible velocity should only be applied in cases of straight, uniform
steady fow (e.g., bank protection)
2. FHWA (2005a), PADEP (2000)
3. Permissible velocity range for easily eroded to erosion resistant soils (clayey fne-grained soils and
coarse-grained soils) at slopes less than 5 percent. Use velocity exceeding 5 ft/sec only where
good cover and proper maintenance can be provided.
4. Based on rock with unit weight of 165 lb/ft
3
.
< PREVIOUS VIEW
9-86
Chapter 9
MAY 2009
to provide greater tractive shear resistance for steeper slopes, but product limitations and the impor-
tance of quality construction practices must be clearly understood, if these products are used.
9.7.3.2.1 Grass-Lined Channels
Design of stable grass-lined channels is related to the type of vegetation selected, as refected by the
vegetation retardance class and the tractive stress associated with peak fow. The FHWA (2005a) pres-
ents design procedures and data for the design of vegetative linings. Grass linings control erosion by
dissipating shear force within the grass stems before it reaches the soil surface, and the root and stem
stabilize the soil against turbulent water forces. As indicated in Table 9.17, vegetation retardance for
grasses is divided into fve classes A through E. In general, taller and denser grass species have a higher
resistance to fow (Class A), while short fexible grass has a low resistance to fow (Class E). The design
procedure presented in FHWA (2005a) includes initially estimating the fow depth and efective shear
stress on the grass lining based on the grass retardance class, roughness, stifness, and cover conditions.
Mannings n is determined from the efective shear stress, and the discharge is computed and compared
to the original estimate. Through a series of iterations, a solution is obtained that balances the fow depth,
shear stress, and discharge. The permissible shear stress for the soil type is then estimated from soil prop-
erties such as grain size, plasticity, and void ratio for comparison to the efective shear stress.
Haan and Barfeld (1978) present a simplifed procedure in which permissible velocities for vegetated
channels are used with the relationship between channel conveyance and Mannings n for the retar-
dance classes shown in Figure 9.36. Table 9.17 presents the vegetative cover and condition for various
retardance classes. The design procedure is to: (1) select the vegetation and an initial estimate of n,
(2) determine retardance class and permissible velocity based on Table 9.17 and an initial estimate
of VR (velocity V times hydraulic radius R) from the curves in Figure 9.26, (3) obtain the permissible
velocity for the vegetation based on Table 9.16 (or state regulatory guidance manual values) and (4)
compute the hydraulic radius R from Equation 9-18 using the permissible velocity. The product of the
permissile velocity and R from Step 4 is compared with the initial estimate of VR to determine a new
value of VR until convergence.
Commercially available erosion control products can be used to reinforce natural vegetation, includ-
ing non-degradable synthetic fbers, flaments, neting and/or wire mesh. These materials can be
integrated into the vegetation and soil lining or applied over the surface, afecting performance
and altering the design procedure. The FHWA (2005a) presents design methods for incorporating
such products into the design of channel linings, and manufacturers have developed sofware for
analysis of unreinforced and reinforced vegetated channel sections (e.g., North American Greens
Erosion Control Materials Design sofware). When such sofware is used, knowledge of the grass
retardance class, vegetation type, soil type and duration of peak fow are essential requirements in
the determination of the tractive shear resistance of the channel linings.
9.7.3.2.2 Riprap-Lined Channels
Riprap linings consist of a layer of rock or stone with a characteristic size, generally designated by
D
50
,
the median grain size of the lining material. As with grass linings, the fow conditions and chan-
nel shear stress are a function of the Mannings n, and an iterative procedure is required for evaluat-
ing lining stability. Values for permissible shear stress for riprap linings are based on laboratory and
feld research. More turbulent fow conditions are more likely to cause lining failure, and a higher
safety factor is recommended for such conditions. Typically, for situations where riprap is needed,
a safety factor of 1.5 should be used, although a lower value may be justifable with mild slopes and
low velocities. The FHWA (2005a) presents a design procedure for riprap-lined channels.
Barfeld et al. (1981) present a simplifed approach for riprap lining design based on USACE pro-
cedures and the critical tractive shear stress. The force on median-size riprap at the threshold of
< PREVIOUS VIEW
9-87
Hydrology and Hydraulics
MAY 2009
TABLE 9.17 VEGETAL RETARDANCE CLASSES
Retardance Class Cover Condition
A
Reed canarygrass Excellent stand, tall (average 36 inches)
Yellow bluestem ischaemum Excellent stand, tall (average 36 inches)
B
Smooth bromegrass
Good stand, mowed (average 12 to 15 inches)
Bermudagrass Good stand, tall (average 12 inches)
Native grass mixture (little bluestem, blue
grama, and other long and short mid-west
grasses)
Good stand, unmowed
Tall fescue
Good stand, unmowed (average 18 inches)
Sericea lespedeza
Good stand, not woody, tall (average 19 inches)
Grass-legume mixture Timothy, smooth
bromegrass, or orchardgrass
Good stand, uncut (average 20 inches)
Reed canarygrass
Good stand, mowed (average 12 to 15 inches)
Tall fescue, with birds foot trefoil or iodino
Good stand, uncut (average 18 inches)
Blue grama
Good stand, uncut (average 13 inches)
C
Bahia Good stand, uncut (6 to 8 inches)
Bermudagrass Good stand, mowed (average 6 inches)
Redtop Good stand, headed (15 to 20 inches)
Grass-legume mixture summer
(orchardgrass, redtop, Italian ryegrass,
and common lespedeza)
Good stand, uncut (6 to 8 inches)
Centipedegrass Very dense cover (average 6 inches)
Kentucky bluegrass Good stand, headed (6 to 12 inches)
D
Bermudagrass Good stand, cut to 2.5-inch height
Red fescue Good stand, headed (12 to 18 inches)
Buffalograss Good stand, uncut (3 to 6 inches)
Grass-legume mixture fall, spring
(orchardgrass, red-top, Italian ryegrass,
and common lespedeza)
Good stand, uncut (4 to 5 inches)
Sericea lespedeza After cutting to 2-inch height, very good stand
before cutting
E
Bermudagrass Good stand, cut to 1.5-inch height
Bermudagrass Burned stubble
(HAAN AND BARFIELD, 1978)
motion is considered to be the critical tractive resisting force per unit area (
c
) for riprap-lined
channel sections. Since riprap is a layer of discrete, individual rocks, the movement of an indi-
vidual rock can initiate further erosion, weakening the overall long-term erosion resistance of the
riprap layer. For most situations at coal refuse disposal facilities, the general practice is to consider
application of a safety factor for the channel base (SF
base
) of 1.5 for the design of riprap channel sec-
tions. The critical tractive resisting force at the channel base can be estimated from the following
equation:
SF
base
= (cos tan ) / ( sin + b tan ) (9-27)
< PREVIOUS VIEW
9-88
Chapter 9
MAY 2009
where:
b = /
c
= shear force per unit area on channel bed =
RS (lb/f
2
)
c
= critical tractive resisting force per unit area = 0.047
( SG 1) D
50
(lb/f
2
)
RS (9-28)
where:
K
3
= 4V
2
/ R
d
(dimensionless)
V = straight channel velocity (f/sec)
R
d
= channel radius of curvature measured at the outside channel bank (f)
The FHWA (2005a) presents details and discussion of design procedures at bends, including the
length over which increased shear stresses are experienced and the superelevation of the water sur-
face at and downstream of bends.
Geotextiles have been used increasingly as the flter medium between the riprap and underlying mate-
rial. However, riprap placed on a geotextile has a tendency to creep over time. Gravitational and water
forces have a tendency to re-orient riprap toward the base of the channel resulting in exposed geotextile
at the top of the channel slope. The use of an aggregate flter layer between the subbase soil and the
TABLE 9.18 RIPRAP SIZE DESIGNATION
Graded Riprap Stone
NSSGA No.
(1)
Sieve or Square-Opening Size (in)
Recommended Filter
Stone
NSSGA Size No
.(1)
Maximum D
50
Minimum
(2)
R-1 1 No. 8 FS-1
R-2 3 1 1 FS-1
R-3 6 3 2 FS-2
R-4 12 6 3 FS-2
R-5 18 9 5 FS-2
R-6 24 12 7 FS-3
R-7 30 15 12 FS-3
R-8 48 24 15 FS-3
Filter Stone
NSSGA No.
(1)
Sieve or Square-Opening Size (in)
Recommended
Placement Thickness
(in)
Maximum D
50
Minimum
(2)
FS-1 No. 30 No. 100 3
FS-2 2 No.4 No. 100 4 to 6
FS-3 6 2 No. 16 8 to 10
Note: 1. National Stone, Sand & Gravel Association (formerly National Stone Association) designation. The
table is based on a stone dry density of 165 lb/ft
3
.
2. Pieces smaller than the minimum size shown should not exceed 15 percent of the tonnage shipped.
(ADAPTED FROM NSSGA, 1989)
< PREVIOUS VIEW
9-90
Chapter 9
MAY 2009
riprap is preferred to geotextile due to the increased frictional resistance between the aggregate flter
material and the riprap. If geotextile is used, extra riprap should be placed on the channel slopes.
The recommended thickness and gradation of the flter stone in relation to the riprap size is pre-
sented in Table 9.18. The recommended riprap layer thickness is greater than or equal to 1.5 times the
D
50
stone size or greater than or equal to the D
100
size, whichever is greater (FHWA, 1989).
The publication Practical Riprap Design (Maynord, 1978) based on model testing provides riprap
size as a function of channel fow depth and Froude number (function of velocity and depth). Addi-
tionally, the USACE (1994) provides a generalized procedure for determining riprap size and dis-
tribution, including guidance for steep slope conditions (up to 20 percent). The model studies and
guidance have proved benefcial for determining riprap requirements in steep slope conditions where
the tractive force determined from Equation 9-27 may result in large riprap size relative to fow depth
such that uniform fow conditions may no longer be valid. In such steep slope conditions, grouted
riprap is ofen employed to reduce the riprap size. Sofware is available for riprap design that utilizes
the USACE and other procedures.
9.7.3.2.3 Grouted-Riprap Channels
Grouted-riprap channel lining consists of riprap with a cement grout flling the voids to create a
monolithic erosion protection mat. The use of grouted riprap can reduce the size and quantity of rock
required in a channel lining by creating a greater material mass to resist hydraulic forces. Because
grouted riprap linings are rigid, they are susceptible to cracking and damage due to subbase move-
ment or freeze-thaw action. The tractive force resistance of cracked, unconfned grout sections is
dependent on intimate surface contact between adjacent sections, section mass, and fow characteris-
tics. In situations of severe cracking with displacement, the primary resistance to tractive forces is the
grout and rock fragment pieces, which may be only slightly greater in size and mass than the riprap
alone. Thus, to limit the potential for displacement, subbase design is very important.
Design procedures for grouted-rock linings installed on channel banks are provided in FHWA (1989)
and address aspects such as rock size and lining thickness, rock grading, flter design, and pressure
relief. Rock size and lining thickness are a function of the fow velocity; the median rock size should
not exceed two-thirds of the lining thickness, and the largest rock size should not exceed the lining
thickness. Table 9.19 and Figure 9.37 can be used to estimate the riprap thickness based on velocity,
riprap size and the recommended depth of the grout penetration as follows:
Based on the average fow velocity, Figure 9.37 can be used to estimate the riprap thick-
ness. For the case where the ratio of botom width to depth is greater than 2, the fow veloc-
ity should be increased by 25 percent when determining the riprap thickness.
The gradation should be selected based on available riprap class (e.g., cobbles), con-
sidering that the median rock size should not exceed two-thirds of the lining thickness
and the largest rock size should not exceed the lining thickness. Table 9.19 presents
AASHTO class designations based on weight, which is approximately equivalent to
the efective diameter shown.
Grout penetration recommendations for each class or size of riprap are also presented
in Table 9.19.
This procedure is applicable to trapezoidal channels where the ratio of botom width to depth is less
than 2. For wider channels, the velocity and shear stress on the channel botom may be greater, and
a corresponding adjustment in velocity or shear stress may be necessary. Typically, the design veloc-
ity will increase by 25 percent because the tractive shear for channel side slopes is approximately 75
percent of the maximum shear at the base of the channel.
< PREVIOUS VIEW
9-91
Hydrology and Hydraulics
MAY 2009
0 5 10 15 20
0
1
2
3
4
5
VELOCITY IN VICINITY OF BANK (FT/SEC)
R
I
P
R
A
P
T
H
I
C
K
N
E
S
S
(
F
T
)
FIGURE 9.37 GROUTED RIPRAP LINING THICKNESS
AS A FUNCTION OF FLOW VELOCITY
25
(FHWA, 1989)
FIGURE 9.37 GROUTED RIPRAP LINING THICKNESS AS A FUNCTION OF FLOW VELOCITY
TABLE 9.19 RECOMMENDED GRADING OF GROUTED ROCK RIPRAP LINING
Rock Sizes Classes (Percent Larger Than Given Rock Size)
Equivalent Diameter
(ft)
2.75 2.25 1.75 1.25 1.0 0.5
3.5 0-5
2.75 50-100 0-5
2.25 50-100 0-5
1.75 95-100 50-100 0-5
1.25 95-100 50-100 0-5
1.0 95-100 95-100 50-100 0-5
0.5 95-100 95-100
Minimum Penetration of Grout
(ft)
2 1.5 1.3 0.8 0.67 0.5
(FHWA, 1989)
Similar to riprap channel lining, the use of an aggregate flter layer between the natural ground sur-
face and the riprap is recommended, but geotextiles can also be used in this application provided the
riprap does not shif on channel slopes during construction or under subsequent fow conditions.
FIGURE 9.37 GROUTED RIPRAP LINING THICKNESS AS A FUNCTION OF FLOW VELOCITY
< PREVIOUS VIEW
9-92
Chapter 9
MAY 2009
Foundation preparation is a critical factor in the performance of a grouted riprap lining. Grouted
riprap is not fexible and foundation conditions, the potential for development of hydrostatic pres-
sure beneath the rigid mat, and end treatment conditions need to be evaluated as part of the design
and construction process. Damage to any section of grouted riprap lining can result in complete
failure of the system.
The foundation for grouted-riprap channel lining should be a contoured, frm stable surface. Small
surface irregularities can be accommodated within the flter stone layer. The slopes should be graded
at no steeper than 1.5H:1V.
Weep holes for hydrostatic pressure relief should be installed in a grouted riprap channel lining. The
weep holes should extend below the grouted surface to the gravel zone between the grout and the
flter stone. It is recommended that 3-inch-diameter weep holes, constructed from PVC pipe and with
protective end screening, be installed at maximum 10-foot vertical spacing (FHWA, 1989).
The head, toe and terminal ends of grouted riprap linings require special treatment such as extension
of the lining to rock or to a depth below potential scour to prevent undermining. Additional riprap
can provide extra protection against undercuting at a bank toe. Figure 56 of Hydraulic Engineering
Circular 11 (FHWA, 1989) presents recommended construction details for these treatments. Guidance
is also provided for rock grading and quality, grout strength and penetration, and flter design and
pressure relief measures. The grout mix should typically be designed for a strength of at least 3,000
psi, with maximum aggregate size of inch and a slump between 3 and 4 inches. Sand mixes may
be used where roughness of the grout surface is undesirable. The fnished grout should leave the
face rock exposed for approximately one-quarter of their depth, and the surface of the grout should
expose the matrix of coarse aggregate. The following construction details should be addressed in the
specifcations for efective grouted riprap lining:
The prescribed method of riprap placement should prevent segregation of rock
sizes, and the riprap should be wet immediately prior to commencing grouting
operations.
The prescribed method of grout placement should control segregation and uniformity.
The prescribed method of grout placement should facilitate grout penetration by use
of vibrating, spading and/or rodding.
Quality control requirements during construction should be followed, including all
material placement methods, grout mix design and strength testing, and recording
of quantities of all materials. The volume of grout used should be compared to that
required to meet penetration requirements.
9.7.3.2.4 Concrete-Lined Channels
Concrete can provide a continuous, rigid channel lining. Similar to grouted riprap, foundation condi-
tions, hydrostatic pressure development and end treatments are design considerations that must be
addressed. Ofsets at joints may create additional hydrodynamic uplif forces. Reinforcing steel (No.
6 gage wire mesh or No. 4 reinforcing bars at 6-inch spacing are typical minimum recommendations
for 4-inch and 6-inch lining thicknesses, respectively) to reduce the development of thermal stresses,
shrinkage and fexural stresses within the Class A (AASHTO classifcation) concrete is normally pro-
vided. The FHWA (1989) provides guidance for design of concrete linings.
Filter layers should be placed below the concrete pavement, and weep holes should be employed
to prevent the development of hydrostatic uplif pressures. The weep-hole confguration described
above for grouted-riprap revetments should be the maximum spacing considered. Hydrodynamic
uplif pressure may also be a consideration where vertical ofsets or changes in channel slope occur,
< PREVIOUS VIEW
9-93
Hydrology and Hydraulics
MAY 2009
particularly at transitions from a steep slope to a fater slope. The USBR (2007b) presents results of
testing for a range of fow velocity, joint widths, and ofset dimensions and provides empirical esti-
mates of hydrodynamic uplif as a function of velocity head. The USACE (1990b) presents guidance
for analysis of vertical curve transitions and recommends that construction joints be excluded from
sections transitioning from steep to fater slopes. Other measures that may be considered include
anchor systems grouted into rock or secured to deadmen.
Edge treatment at the toe, head, and terminal ends should be utilized to prevent undermining. Stub
walls or cutof walls are recommended at expansion joints.
9.7.3.2.5 Commercially Available Composite Erosion Control Products
Commercially available composite erosion control products include permanent reinforced vegeta-
tion mats; interlocking concrete blocks; grout-flled nylon mats (unanchored or anchored); soil,
rock or concrete flled cellular confnement mats; gabion mats; cabled concrete or other similar
products. The allowable tractive shear for these products will generally be specifed by the manu-
facturer or can be determined by computer sofware developed by the manufacturer. The efective-
ness of a particular type of channel lining relative to the intended use should be verifed with the
manufacturer prior to design. The FHWA (1989, 2005a) provides guidance for some commercially
available products.
Applications of interlocking concrete blocks, grout-flled nylon mats, cellular confnement mats,
and cabled concrete at coal refuse disposal facilities have included emergency spillway linings,
groin ditches, principal spillway outlet channels, and diversion ditches where high velocities are
present and excavation into rock is not possible. Design procedures are generally available from
the manufacturers and these procedures should include methods for assessing tractive forces and
uplif. For grout and concrete systems, additives such as steel or polyester fbers are sometimes
employed to increase the strength of the concrete, decrease cracking, and to improve resistance to
hydraulic wear.
Some design and construction issues for concrete and grouted riprap linings may also apply to com-
posite systems, including:
The foundation should be graded smooth and compacted to maintain support and
prevent detrimental movement. Some composite systems require a flter and/or
drainage layer beneath the lining.
Side and end anchorages should follow manufacturers recommendations, and when
unusual site conditions are encountered, the manufacturer should be contacted for
input.
Weep holes must be provided where necessary, consistent with manufacturers rec-
ommendations.
Access to allow inspection and maintenance must be planned. If vehicles and equip-
ment have to cross the channel lining, a reinforced section should be designed.
9.7.3.3 Cavitation
Cavitation of a hydraulic system occurs when fow separates from a containment surface. The result
is the formation of a region of subatmospheric pressure that can lead to deterioration of the confning
surface. Cavitation can occur in either open-channel or closed-conduit fow. As a rule of thumb, cavi-
tation should be investigated whenever fow velocities exceed 35 feet per second (USACE, 1990b).
The potential for damage is a function of fow duration and geometry and the abrasion resistance of
the material.
< PREVIOUS VIEW
9-94
Chapter 9
MAY 2009
In open-channel fow, cavitation can occur at steepening grade changes. At constructed ogees and
paved channels, formation of a cavity of subatmospheric pressure can lead to damaging vibrations;
under extreme conditions, these vibrations may actually displace the structure. In unpaved channels,
cavitation can produce an upstream propagating erosion of the channel botom that will worsen and
possibly cause failure of the entire hydraulic system if uncorrected.
In closed conduits, cavitation most frequently occurs at sharp turns. For hydraulic systems common
to coal refuse disposal facilities, this is particularly likely to occur at the transition from a drop inlet
to the entrance of the transport conduit. Should the frequency and duration of high-velocity fow
and conduit susceptibility to cavitation damage warrant, methods to prevent cavitation in closed
conduits (USBR, 1987a) and USACE (1980, 1990b) may be needed.
9.7.3.4 Directional Changes in Open-Channel Flow
Methods for predicting the superelevation of the surface of curving fow are available, but the
superelevation is usually small for subcritical fow velocities. When the fow at a curve is super-
critical, a directional change is an exceptionally complicated problem because of the formation,
propagation and combination of channel cross waves. These factors must be considered in design
if channel overtopping is to be prevented. A simplifed determination of the additional fow depth
associated with superelevated fow at a rectangular channel bend is provided by the following
equation (Chow, 1959):
h = {V
max
2
/g} {(
20
3
)(r
c
/b) 16(r
c
3
/b
3
) + ((4r
c
2
/b
2
) 1)
2
ln[(2r
c
+ b)/(2r
c
b)]} (9-29)
where:
h = change in depth associated with the superelevation (f)
V
max
= velocity from Mannings equation in straight channel section
approaching bend (f/sec)
r
c
= radius of curvature measured to the centerline of the channel (f)
b = width of the channel (f)
g = gravitational acceleration = 32.2 f/sec
2
The FHWA (2005b) presents a method for estimating the superelevation of fow at bends in trap-
ezoidal channels based on the velocity head. Additional guidance is provided in USBR (1987a) and
USACE (1980, 1990b).
9.7.3.5 Directional Changes in Conduit Flow
Sharp directional changes and intersections in closed conduits cause static and dynamic thrusts
that must be considered in design. Depending on the conduit size, the internal pressure and the
fow velocity, thrust forces can cause the backfll surrounding the conduit to compress, and joints
in the conduit can open or the defection of the conduit can exceed the allowable pipe defection
due to the resulting movement. This is particularly a problem when the conduit location is shallow
(or the conduit is supported above ground) and soil resistance to movement is low. Where thrust
movements are identifed as a potential problem, a concrete thrust block is ofen poured around
the conduit to add mass and to distribute pressure over a larger area of soil. Discussions for evalu-
ating the necessity of special construction measures at directional changes or intersections and
procedures for analyzing requirements of thrust blocks are provided by Brater et al. (1996) and in
engineering manuals from conduit manufacturers. The need for thrust blocks should be evaluated
< PREVIOUS VIEW
9-95
Hydrology and Hydraulics
MAY 2009
for ftings such as tees and prefabricated bends and at directional changes where conduit joints
could separate.
9.7.3.6 Materials Selection
The construction of any type of hydraulic system, and particularly a closed-conduit system, should
not be undertaken without investigating the hydraulic and structural suitability of the proposed
construction materials. This topic is discussed in Section 9.3.3 and Section 6.6.6.1 of this Manual.
The discussion herein is for open-channel systems and culvert-conduit systems with emphasis
on the importance of selecting materials that will function as intended and avoid costly repair
or replacement. A detailed discussion of decant system material selection is presented in Section
9.7.4.1 with specifc atention to the structural integrity of the decant system due to embankment
loading conditions.
9.7.3.6.1 Open-Channel Lining Systems
As part of the hydraulic design process, the suitability of the selected channel lining material should
be determined. Material selection should be based on factors such as resistance to abrasion, the type
of fow in the channel (continuous or intermitent), the acidic nature of the coal refuse, the impact
that water quality characteristics may have on lining system integrity, the availability of the lining
material, foundation conditions (particularly for more rigid lining systems), the constructability of
the system (e.g., site topography and site access conditions), cost, and maintenance.
9.7.3.6.2 Conduit Materials
Culvert pipes can be corrugated metal (CMP); concrete; corrugated plastic (CPP), both smooth-wall
or corrugated interior; high density polyethylene (HDPE); polyvinyl chloride (PVC); aluminum; steel
or other materials. The type of material recommended is generally a function of intended use of the
culvert (temporary or permanent), loading conditions, foundation conditions, construction limita-
tions, and cost. Uncoated corrugated metal and steel are generally not recommended for long-term
use in a mining environment because of the corrosion potential of mine water. Limited usage or pro-
tective coatings can make these material alternatives more acceptable.
Culverts constructed of concrete, CMP and aluminum are manufactured in various shapes (box,
oval, arch, etc.) for installation in areas with limited height and clearance. Minimum and maxi-
mum cover height limitations are associated with all culvert installations. CPP and HDPE pipe
are generally more fexible and structurally stable in conditions where minor setlements may
occur.
In all applications, installation is critical to the structural integrity and hydraulic conveyance capa-
bilities of conduits. The thickness of and installation procedures for bedding and backfll materials
are critical to successful culvert construction. Joint connections in most culvert installations should
be minimal. If joints are present, they should be soil-tight and in most applications watertight.
Fusion-welded HDPE, gasketed joints for some concrete and CPP, glued PVC, and welded steel
pipe provide the most watertight applications. Pressure testing can be performed to verify that
joints are watertight.
9.7.4 Types of Hydraulic Systems
Sections 9.7.1, 9.7.2 and 9.7.3 have identifed: (1) basic considerations for planning hydraulic systems
at coal refuse disposal facilities, (2) the primary components of all hydraulic systems and techniques
appropriate for their analyses, and (3) special design considerations associated with hydraulic sys-
tems. This section integrates this basic information into the following discussions of specifc types of
hydraulic systems most common to coal refuse disposal facilities.
< PREVIOUS VIEW
9-96
Chapter 9
MAY 2009
9.7.4.1 Decant Systems
Decant systems at impounding disposal facilities serve one of the following three purposes or a com-
bination thereof:
1. To remove clarifed water during normal disposal of fne refuse.
2. To provide outfow during low-precipitation storms so that storage volume will be
available if a large storm occurs.
3. To drain (possibly in conjunction with a spillway) the stored volume of infow due to
a large storm within a reasonable period afer occurrence of the storm.
A decant system typically consists of: (1) an inlet section located in the impoundment at the elevation
required for controlling or limiting the normal water surface level, (2) a transport section consisting
of a closed conduit beneath or through the embankment, and (3) a discharge section located down-
stream from the embankment so as to minimize erosion of the embankment toe. It is the responsibil-
ity of the designer to select the optimum location for the transport section conduit and the discharge
point, based on foundation conditions, conduit size and material, and discharge rates.
The optimal selection of the decant inlet type, confguration and location is primarily a function of
the overall facility confguration, the required discharge capacity, the method for disposing of the fne
refuse, the rate at which the impoundment level will rise during operations, and eventual abandon-
ment or post-mining land use requirements. Design considerations include the height of the decant
inlet invert above the setled fne coal refuse, the method of evacuation of clarifed water between the
decant inlet invert and the setled fne coal refuse level, the trash rack system for preventing debris
from entering the decant pipeline, and the potential buoyancy of inlet and conduit.
Ofen the most challenging design consideration is selecting a conduit that will withstand the weight
of the overlying refuse embankment and that can deform as the foundation and embankment mate-
rials setle vertically or displace laterally. Generally, only specially designed concrete, high density
polyethylene (HDPE), or steel pipes are capable of withstanding the high pressures beneath a refuse
embankment. Section 6.6.6 provides guidance for the selection of decant materials and designing for
embankment loading, including design of the decant pipe bedding and backfll. The transition from
the riser to the transport section of the decant must be designed to handle the impact loads associ-
ated with directional change in fow. This is particularly important for rigid pipe systems, which are
sensitive to movement. Control of seepage along and adjacent to the transport section of the decant,
where it extends through the embankment, must also be addressed (Section 6.6.2.3).
To prevent erosion, the decant outlet must be able to accommodate the design fow rate and velocity.
The rate and velocity of outfow must also be considered in the design of downstream conveyance
and/or storage structures.
In the following sections, guidelines for the design of the hydraulic conveyance components of a
decant system are presented. The inlet, transport and outlet sections of a decant system will be dis-
cussed separately. The discussion is based on the use of concrete, welded steel, and HDPE, since these
are the most commonly used materials for impoundment decant systems.
9.7.4.1.1 Inlet
Location within the Impoundment
Location of the decant inlet within the impoundment is a function of the following factors: (1)
site terrain and foundation conditions, (2) limitations of the transport section of the decant such
as length and foundation conditions, (3) type of embankment construction such as upstream or
< PREVIOUS VIEW
9-97
Hydrology and Hydraulics
MAY 2009
downstream construction, and (4) limitations in positioning the slurry discharge. Ofen, the inlet
is positioned in the upstream portion of the impoundment, so that fne refuse can be deposited to
form a delta at the upstream slope of the embankment and clarifed water accumulates in the far-
thest upstream portion of the impoundment.
Trash Rack
The entrance to decant inlets should be protected by a trash rack. Even a partially obstructed pipe
will have a substantially reduced capacity, thus increasing the potential for dam overtopping during
a large storm event. Trash racks can become plugged if the openings are too small, and openings that
are too large can result in the obstruction of the pipe due to the intake of large debris. The connec-
tion of the trash rack to the decant structure must be strong enough to withstand the hydrostatic and
dynamic forces exerted on the trash rack during periods of high fow.
It is recommended that trash rack openings be sized so that they are a maximum of one-half the
nominal dimension of the outlet conduit. The minimum opening size should be 6 inches by 6 inches
or greater. The USBR (1987a) recommends that the area of trash rack openings be established based
on the fow velocity through the rack. Where trash racks are inaccessible for cleaning, this velocity
should not exceed 2 feet per second. A velocity of up to approximately 5 feet per second can be toler-
ated for racks that are accessible for cleaning. An anti-vortex device should be incorporated into the
trash rack design to prevent the formation of a fow-inhibiting vortex during periods of high fow.
The USBR (1987a) recommends that the anti-vortex device extend at least two diameters in front of
and to each side of the inlet. In practice, these devices are usually part of the trash rack assembly and
consist of a steel plate with width equal to the width (or diameter) of the rack.
Evacuation of Water below Riser Invert
As part of the design process, a sufcient minimum depth of water and associated height of riser above
the setled fne coal refuse should be provided in order to prevent short circuiting of the impound-
ment and release of fne coal refuse slurry. This depth is frequently estimated based on experience
and judgment, usually varying between 5 and 10 feet to accommodate setling of fne refuse in the
slurry. Setling tests can be performed in the laboratory to aid in determination of the rate of setle-
ment and the required impoundment retention time. The setling velocity can be determined from
Stokes Law based on the particle size and specifc gravity of the fne coal refuse:
V
s
= [ g ( S-1) D
2
] / 18 (9-30)
where:
V
s
= setling velocity (cm/sec)
g = acceleration of gravity (981 cm/sec
2
)
D = diameter of particle (cm)
= kinematic viscosity of fuid (cm
2
/sec)
S = specifc gravity of particle
The minimum height of the riser above the average setled fne coal refuse level can be calculated by
determining the approximate detention time based on the outfow rate (slurry discharge rate plus
watershed base fow) and impoundment geometry. To establish a balance between the detention time
and period for setling, an iterative process is required. Because of the broad range in refuse particle
sizes and varying impoundment geometry, experience and engineering judgment are generally used
for determining the depth of water to be retained over the fne refuse.
< PREVIOUS VIEW
9-98
Chapter 9
MAY 2009
Pumps are generally used to remove the clarifed water below the riser invert. Such pumps should be
capable of meeting the discharge requirement without increasing the fow rate through the setling
zone sufciently to cause removal of fne coal refuse.
Inlet Riser Pipe Buoyancy
Inlet riser pipe sections may be susceptible to the buoyancy forces sufcient to cause uplif, if the pipe
weight and backfll height is not adequate. This is particularly a concern for steel and HDPE decant
pipe inlets. Inlets to an HDPE decant pipe are generally installed in a trench extending up a natural
hillside with the inlets located at specifed elevation intervals. Sufcient fll or anchorage must be
provided to overcome buoyant forces, with a recommended factor of safety of 1.5.
Extension of Inlet
The decant riser inlet elevation is generally established based on storm routing such that adequate
impoundment freeboard and surcharge storage is provided. As a refuse embankment dam is raised
to increase the capacity of the impoundment, the riser inlet is correspondingly raised to provide
additional slurry disposal capacity. If the decant system has multiple risers, the lower riser is sealed
and abandoned and the next upper riser is put into service. Extension of an inlet will require that
the connection, extension, and new inlet section be designed to accommodate the fows and related
forces associated with continuing operation. If the decant system is designed with multiple risers,
sealing (such as by employing a bolt-on plate) and abandonment (such as by embedment in concrete)
of the riser must be part of the design. To assess the potential for additional stress and defection at
the base of the sealed riser, the loads associated with abandonment (along with future embankment
construction) must be evaluated.
In some instances where high embankments are required and there are concerns about pipe loading,
designers have limited the service life of a decant inlet and transport pipe section by abandoning the
entire system and installing a second system at a higher elevation. In such situations, the original inlet
and transport section should be sealed and abandoned upon completion of the replacement system.
9.7.4.1.2 Transport Section
The transport section of the decant pipe must typically be watertight and must include seepage con-
trol structures to intercept the fow of seepage along and adjacent to the conduit where it extends
through the refuse embankment. For these reasons, structural considerations based on the exter-
nal loading generally govern the design of this portion of the decant once the diameter has been
established. Determination of external loading conditions and structural design of a decant pipe are
addressed in Section 6.6.6.
Rigid pipe used within a refuse embankment is typically concrete pressure pipe because of the large
external loads and the requirement that the pipe be watertight. Concrete pipe is rigid and sensitive
to diferential setlements, particularly at the inlet-riser transition, resulting from directional fow
impacts, foundation conditions, or imposed embankment loading conditions.
Infltration and exfltration leakage problems can develop within the transport section of the decant
pipe. Irreparable damage of the pipe and possibly failure of the refuse embankment can occur if such
leakage is undetected. As part of the installation process, project specifcations should require that
pipes installed within the limits of the impounding refuse embankment be pressure tested prior to
backflling so that detected leaks can be immediately repaired.
Joint tightness is also a concern in non-pressure pipe installations, such as in the upstream inlet sec-
tion of the decant pipe or in a concrete riser extension. Infltration or exfltration at joints could impact
the pipe backfll if the material is erodible. Testing of non-pressure pipe joints is recommended.
< PREVIOUS VIEW
9-99
Hydrology and Hydraulics
MAY 2009
Flexibility, watertightness and relatively easy installation procedures have led to the use of
HDPE for decant pipes. Structural evaluation procedures associated with the fexible conduit
under large embankment loads are presented in Section 6.6.6. For fexible pipe, both structural
and hydraulic design considerations may control the decant pipe size. Thus, evaluation of the
embankment loading and required wall thickness of the pipe should be performed in parallel
with the hydraulic design.
Schematic examples of decant inlets that are most adaptable to coal refuse disposal facilities are
illustrated in Figure 9.38, while illustrations of actual decant systems that have been constructed
are shown in Figures 9.39 and 9.40. The primary advantage of the inlet types shown in Figure 9.38a
through d is their access for expansion as the level of setled fne refuse increases. The primary dis-
advantage is that the length of conduit required upstream of the embankment, beneath the setled
slurry, is relatively great. The length of conduit in the Figure 9.38a and c decant systems is particu-
larly signifcant because it must be as long as the entire impoundment. However, an advantage of
these systems is that the inlet is located at the point where the water depth is normally the greatest
(when the slurry is discharged at the embankment end of the impoundment), allowing the decant
system to drain the water without pumping.
The inlet shown in Figure 9.38d appears to be the simplest arrangement because it ofers the shortest
conduit length and the inlet can be extended in height by adding subsequent sections as the setled
fne refuse level rises. Major difculties with this type of decant system are: (1) its location within the
impoundment, which makes access very difcult, and (2) large impact forces at the point of direc-
tional change from the inlet to the transport section. With regard to the later, the efect of forces due
to falling water must be accounted for by providing a curved section at the base of the vertical riser
and/or constructing a concrete pad to distribute impact loads to the underlying soils.
Figure 9.39 presents an example of a decant system installed at a diked-impoundment facility. In
this example, a fexible pipe serves as the decant with the riser extending vertically up into the
impoundment. A water return line to the coal preparation plant is installed parallel to the decant
line, and the impoundment water level is controlled by a pumping system. The decant inlet is posi-
tioned so as to provide for storage of the design storm runof, which for a diked impoundment is
predominantly the precipitation falling upon the impoundment surface. Note that the water return
line, as would the case for any conduit passing through the embankment, must be designed with
seepage control measures.
Figure 9.40 shows an example of a decant system installed at a cross-valley impoundment, where
multiple risers, or inlets, extend from the transport section of the fexible pipe. In this example, a
pumping system is used to maintain the impoundment water level, and the decant serves to rout sig-
nifcant storm runof through the facility. As the setled slurry accumulates within the impoundment,
the lower riser inlet is sealed and the next upper riser inlet is fted with a trash rack for operation. In
the example system shown in Figure 9.40, a seepage interception zone has been installed within the
backfll for the decant system.
9.7.4.2 Overfow Spillway Systems
When the design storm is not stored in the impoundment at coal refuse disposal facilities, over-
fow spillways that protect downstream life and property by safely discharging the accumulated
runof from large storms are the most crucial hydraulic systems. They frequently must be designed
to convey substantial fows over steep terrain. The vital importance and design complexity of these
structures combined with the complications of a constantly changing embankment confguration,
dictates that spillways for coal refuse disposal facilities be viewed diferently than spillways for other
types of impoundments.
< PREVIOUS VIEW
9-100
Chapter 9
MAY 2009
FIGURE 9.38 EXAMPLES OF DECANTS
COAL REFUSE
EMBANKMENT
SETTLED
SLURRY
SETTLED
SLURRY COAL REFUSE
EMBANKMENT
BLOCKED EXTENSION FOR
FUTURE DECANT INLET
SEALED DECANT INLETS
FROM PREVIOUS EMBANK-
MENT STAGES
9.38a DECANT INLET ON SLOPE AT SIDE OF IMPOUNDMENT
COAL REFUSE
EMBANKMENT
SETTLED SLURRY
9.38c DECANT INLET AT UPSTREAM END OF IMPOUNDMENT
COAL REFUSE
EMBANKMENT
SETTLED
SLURRY
COAL REFUSE
EMBANKMENT
9.38d VERTICAL DECANT INLET
9.38b DECANT INLET ON EMBANKMENT SLOPE
FIGURE 9.38 EXAMPLES OF DECANTS
FIGURE 9.38 EXAMPLES OF DECANTS
< PREVIOUS VIEW
9-101
Hydrology and Hydraulics
MAY 2009FIGURE 9.39 EXAMPLE DECANT SYSTEM AT DIKED IMPOUNDMENT
28" DECANT
RISER (INLET EL. 650)
9.39a PLAN
S
T
A
G
E
V
I
I
I
A
C
R
E
S
T
E
L
.
6
5
8
PUMP FOR PLANT WATER
USE AND FLEXIBLE LINE
ON FLOATING PLATFORM
IMPOUNDMENT
FILTER/DRAINAGE
DIAPHRAGMS
DEPOSITED
FINE REFUSE
8" S
D
R
15.5 - C
LA
R
IFIE
D
W
ATE
R
D
IS
C
H
A
R
G
E
22" S
D
R
15.5 - S
TO
R
M
W
ATE
R
D
IS
C
H
A
R
G
E
DECANT OUTLET
INVERT EL. 585.5
DEPOSITED
FINE REFUSE
9.39c TRENCH DETAIL 9.39b DECANT RISER DETAIL
TRASH RACK
CONCRETE FOUNDATION
CONCRETE SAND
22" SDR 15.5 POLYETHYLENE PIPE
28" FLANGE ADAPTER (SDR 15.5)
24" TEE
(SDR 15.5)
B
d
6"
COMPACTED COARSE
COAL REFUSE
6"
CONCRETE SAND IN
IMPOUNDMENT AREA
MINIMUM B
d
= PIPE O.D. + 24
FIGURE 9.39 EXAMPLE DECANT SYSTEM AT DIKED IMPOUNDMENT
Most transport sections for open-channel spillway systems at impounding disposal facilities are chan-
nels excavated into an embankment abutment. Erosion protection should be provided if the abut-
ment materials in the excavated channel are not durable and could erode, creating stability concerns.
The spillway channel should extend to a point downstream of the embankment and be appropriately
lined and have sufcient freeboard to protect the embankment. Examples of spillways that have been
constructed at impounding disposal facilities are illustrated in Figures 9.41 through 9.43.
Figure 9.41 illustrates a condition where a spillway channel was constructed in an existing valley fll
and side hill embankment, necessitating a variety of channel confgurations and linings. Concrete
and grouted riprap materials were used in steep channel segments and at changes in fow direction.
FIGURE 9.39 EXAMPLE DECANT SYSTEM AT DIKED IMPOUNDMENT
< PREVIOUS VIEW
9-102
Chapter 9
MAY 2009
FIGURE 9.40 EXAMPLE DECANT SYSTEM AT CROSS-VALLEY IMPOUNDMENT
FIGURE 9.40 EXAMPLE DECANT SYSTEM AT CROSS-VALLEY IMPOUNDMENT
9.40a PROFILE
SEEPAGE
INTERCEPTION
ZONE
OPERATING
DECANT RISER
SEALED
DECANT RISER
INITIAL
STAGE
DECANT PIPE
PROTECTIVE COVER
COMPACTED TYPE A
CONCRETE SAND
AASHTO NO. 8
COARSE AGGREGATE
9.40c SECTION B - B
PROTECTIVE COVER
COMPACTED TYPE A
CONCRETE SAND
PERIODIC PIPE SUPPORT
FLOWABLE FILL
CRADLE
9.40c SECTION A - A
9.40b PLAN
C
C
A
A
4" PVC COLLECTION
PIPE (SLOTTED)
4" PVC COLLECTION
DRAIN (SOLID)
END CAP (TYP.)
AASHTO NO. 8 COARSE AGGREGATE
T
R
E
N
C
H
SEEPAGE
INTERCEPTION ZONE
UPSTREAM
EMBANKMENT ZONE IMPOUNDMENT
D D
FABRIFORM LINING
T
R
E
N
C
H
2" WEEP HOLES
DECANT RISER
CONCRETE
ENDWALL
26" HDPE PIPE, DR11
FLOW
COMPACTED TYPE A
CONCRETE SAND
B
B
9.40f SECTION D - D
P
R
O
T
E
C
T
IV
E
C
O
V
E
R
26" H
D
P
E
P
IP
E
, D
R
11
O
C
O
M
PAC
TED
TYPE A
C
O
N
C
R
ETE SAN
D
BOLT ON TRASH RACK
OR TEMPORARY STEEL
PLATE WITH VENT
FLOWABLE
BACKFILL
D
E
C
A
N
T
R
I
S
E
R
2
6
"
H
D
P
E
P
I
P
E
,
D
R
1
1
TERRACE EXCAVATION
FOR FLOWABLE FILL
AT RISER
PROTECTIVE COVER
PERIODIC PIPE SUPPORT
9.40e SECTION C - C
FLOWABLE
BACKFILL
FLOWABLE FILL
CRADLE
FIGURE 9.40 EXAMPLE DECANT SYSTEM AT CROSS-VALLEY IMPOUNDMENT
< PREVIOUS VIEW
9-103
Hydrology and Hydraulics
MAY 2009
Figure 9.42 illustrates a condition where the height of the natural abutment for a new refuse embank-
ment was only slightly higher than the planned embankment crest. A spillway was constructed by
making an excavation through the abutment so that storm fows would discharge into an adjacent
valley and the refuse embankment would not be endangered.
Figure 9.43 illustrates a cascade spillway system. A new spillway at the higher elevation was exca-
vated into a hillside to direct fow to a point above the beginning of the lower spillway. The plunge
pool in the lower spillway dissipates the energy of the cascading water and turns the direction of fow
downhill and away from the embankment.
The design capacity for spillway systems is usually determined by the reservoir routing procedures
presented in Section 9.8. Spillway design involves selecting the control location and method to achieve
the required fow capacity while fully utilizing the available storage capacity.
Figures 9.44 and 9.45 show typical inlet and outlet controls for spillway systems. To determine the
most appropriate control for a particular situation, the following factors should be considered:
Inlet Control Inlet control for spillway systems is desirable when it is important to minimize
the size of the transport channel, including the following situations:
The length of the channel downstream from the inlet that can be economically
constructed is limited (fow will be supercritical downstream from the inlet).
The area for construction of a downstream transport section is limited due to ter-
rain instability (hilly and steep terrain).
Substantial storage capacity is available in the impoundment and economies can
be realized by reducing the size of the transport section or channel (encountered
at some large facilities sited in long valleys).
Competent, erosion-resistant material is available at the inlet of the system but
not at the outlet (inlet control is more easily accomplished).
Outlet Control Outlet control for spillway systems is desirable when it is important to minimize
the velocity of fow or the grade of the transport section, including:
The upstream fow in the transport section must be maintained at subcritical
velocities to minimize erosion of sof channel materials. An overfow weir located
at the downstream end of the transport section or channel can be used to create
this condition.
Competent materials that provide erosion resistance are present at the outlet.
A natural overfall occurs at the outlet, where the water discharges by free fall
without causing signifcant damaging erosion.
Regardless of the point of control, the primary requirement of spillway system design is that the dis-
charged fow not adversely afect the safety of the overall coal refuse disposal facility. Section 9.7.2.3
discusses the types of outlets available for either safely discharging the fow away from the facility or
for dissipating the fow energy with a stilling basin. Generally, the spillway discharge point is located
a sufcient distance away from the embankment so that a special hydraulic structure is not required.
The examples illustrated in Figures 9.41 and 9.43 show excavated channels discharging onto steep
slopes. The major design issue for this approach concerns the frequency and severity of damage that
might result from such discharge. Discharges should be rare occurrences, and damages should be
limited to surface erosion of the steep slope without adverse impact to the disposal facility embank-
ment. Also, the discharge point and slope should be within mine property ownership and not be
< PREVIOUS VIEW
9-104
Chapter 9
MAY 2009
FIGURE 9.41 EXAMPLE SPILLWAY SYSTEM A
FINE COAL REFUSE
IMPOUNDMENT
1
6
0
0
1
6
5
0
1
7
0
0 1
7
5
0
1
2
0
0
1
3
0
0
1
4
5
0
1
5
0
0
1
4
5
0
1
4
0
0
1
5
5
0
CREST OF
EMBANKMENT
S
T
R
E
A
M
HAUL ROAD
BENCH @ EL. 1375
BENCH @ EL.1250
SPILLWAY
BENCH @ EL. 1500
SPILLWAY
SEE DETAIL 1
1
2
5
0
1
2
0
0
1
1
5
0
1
3
0
0
1
3
5
0
1
4
0
0
1
4
5
0
1
5
0
0
1
5
5
0
1
6
0
0
B
B
A
A
9.41a PLAN
FIGURE 9.41 EXAMPLE SPILLWAY SYSTEM - A
9.41c SECTION A - A
6'
1
1.5
37'
1
3
COARSE COAL
REFUSE
10'
TOP OF ROCK
2' THICK RIPRAP
EXISTING GROUND
SURFACE
TOP OF ROCK
1
1 5.5'
1
1
21' 3
1
COARSE COAL
REFUSE
9.41d SECTION B - B
1100
1150
1200
1250
1300
1350
1400
1450
S
L
O
P
E
1
0
%
BOTTOM EL. 1425
SLOPE 0.5%
BOTTOM EL. 1435
E
L
E
V
A
T
I
O
N
(
F
T
)
9.41b SPILLWAY PROFILE (NTS)
SLOPE 60%
9.41f SECTION C - C
9.41e DETAIL 1
15'
EXISTING
GROUND SURFACE
1
3
3' KEY INTO ROCK
8" CONCRETE SLAB
SET ON ROCK
BOTTOM OF
CHANNEL
20' 9'
1' REINFORCED
CONCRETE WALL
12'
1
1
'
1
2
0
'
3:1
SLOPE
21'
REINFORCED
CONCRETE WALL
9'
60% SLOPE
1
0
%
S
L
O
P
E
C
C
FIGURE 9.41 EXAMPLE SPILLWAY SYSTEM A
< PREVIOUS VIEW
9-105
Hydrology and Hydraulics
MAY 2009
FIGURE 9.42 EXAMPLE SPILLWAY SYSTEM B
2
1
2
1
50'
EMERGENCY
SPILLWAY
C
L
9.42c SECTION A - A
FIGURE 9.42 EXAMPLE SPILLWAY SYSTEM - B
9.42b SPILLWAY PROFILE
E
L
E
V
A
T
I
O
N
(
F
T
)
800
810
820
830
840
850
860
870
880
890
900
1% SLOPE
2.2% SLOPE
20' CREST AT
EL. 825
7
0
0
6
7
5
6
7
5
7
0
0
7
2
5
7
7
5
7
5
0
8
0
0
825
850
8
7
5
9
0
0
8
5
0
8
7
5
9
0
0
925
9
2
5
9
0
0
8
7
5
8
5
0
8
2
5
8
0
0
750
7
7
5
8
0
0 8
2
5
8
5
0
8
7
5
900
BENCH AT
EL. 725.0
2:1 SLOPE
EMBANKMENT
CREST
EL. 830.0
2:1 SIDE
SLOPE
EMERGENCY
SPILLWAY
A
A
CREST AT
EL. 830.0
20' CREST AT
EL. 825
2:1 SLOPE
9.42a PLAN
FIGURE 9.42 EXAMPLE SPILLWAY SYSTEM B
< PREVIOUS VIEW
9-106
Chapter 9
MAY 2009
FIGURE 9.43 EXAMPLE SPILLWAY SYSTEM C
FIGURE 9.43 EXAMPLE SPILLWAY SYSTEM - C
9.43c SECTION A - A
25'
EXISTING GROUND
SURFACE
SPILLWAY
C
L
1
2
1
2
2
1
1.5
1
TOP OF ROCK
4
0
'
C
R
E
S
T
@
E
L
.
1
6
6
0
.
0
EXISTING SLURRY
IMPOUNDMENT
A
V
E
R
A
G
E
S
L
O
P
E
2
.8
:1
1800
1675
1650
1
6
2
5
1575
1
8
0
0
1
7
7
5
1
7
5
0
1
7
2
5
1700
1675
A
SPILLWAY
SPILLWAY
9.43a PLAN
1775
1
5
0
0
1750
1725
PLUNGE POOL
1
5
2
5
1700
1550
1
6
0
0
A
9.43b SPILLWAY PROFILE
300'
EL. 1647
SPILLWAY
ENTRANCE
375'
EL. 1644
40'
EL. 1570
EL. 1600
EL. 1605
PLUNGE POOL
40'
FIGURE 9.43 EXAMPLE SPILLWAY SYSTEM C
< PREVIOUS VIEW
9-107
Hydrology and Hydraulics
MAY 2009
FIGURE 9.44 SPILLWAY INLET CONTROL
IMPOUNDMENT
CULVERT
9.44a CULVERT INLET CONTROL
FIGURE 9.44 SPILLWAY INLET CONTROL
SECTION B - B SECTION A - A
IMPOUNDMENT
WEIR
9.44d WEIR INLET CONTROL
IMPOUNDMENT
9.44c NARROW INLET CONTROL
B
B
A
A
IMPOUNDMENT
9.44b INLET TO STEEP CHANNEL CONTROL
FIGURE 9.44 SPILLWAY INLET CONTROL
< PREVIOUS VIEW
9-108
Chapter 9
MAY 2009
FIGURE 9.45 SPILLWAY OUTLET CONTROL
upgradient from structures or public transportation facilities. More difcult to evaluate are property
and/or environmental damages that might occur below the discharge point due to vegetation loss,
surface erosion and even local landslides. The following criteria are suggested as design guides for
preventing impacts to a disposal facility embankment when the discharge of open-channel spillways
onto slopes is contemplated:
If the spillway will only be activated during very large storms (i.e., on the order of
the PMF), special provisions are seldom required at the outfow point because the
resulting erosion damage typically will not be signifcantly greater than the erosion
FIGURE 9.45 SPILLWAY OUTLET CONTROL
IMPOUNDMENT
TRANSPORT SECTION
9.45a WEIR OUTLET CONTROL
SECTION A - A SECTION B - B
IMPOUNDMENT
TRANSPORT SECTION
9.45b OUTLET TO STEEP CHANNEL CONTROL
IMPOUNDMENT
B
B
9.45c NARROW OUTLET CONTROL
TRANSPORT SECTION
A
A
FIGURE 9.45 SPILLWAY OUTLET CONTROL
< PREVIOUS VIEW
9-109
Hydrology and Hydraulics
MAY 2009
damage that would have occurred anyway, and the probability of the spillway func-
tioning many times during the operational period of the facility is low.
If the spillway is expected to be activated as ofen as every 25 years or more fre-
quently, special provisions should be made to transport the fow without excessive
erosion, or the fow should be discharged through a separate system.
High velocity fow in open channels represents a signifcant source of energy that must be controlled for safe
water conveyance. Special considerations for the design of excavated channel spillway systems include:
Initial planning and design should account for the manner in which the spillway
system may need to be extended or modifed as the disposal facility increases in size.
The channel materials (natural or constructed) must be resistant to erosion
(Section 9.7.3.1).
The stability of the excavated slopes forming the channel must be evaluated
(Section 6.6.5).
Potential erosion at directional changes, particularly during supercritical fow, must
be evaluated and accounted for in the channel design (Section 9.7.3.3).
Rigid spillway linings must be designed to resist the development of hydrostatic and
fow-induced uplif pressures.
Sufcient freeboard to contain the discharge within the spillway channel must be
provided.
An important consideration in the design of rigid spillway linings is uplif pressures. For spillways
with rigid lining, uplif is typically estimated based on the hydrostatic head associated with the
normal pool level applied at the upstream end, varying linearly to the hydrostatic head associated
with the downstream fow depth (tailwater) level. Where there is a potential for open, ofset joints
in steep spillway chutes, the velocity may be converted to dynamic pressure. While the theoretical
maximum dynamic pressure should be calculated (e.g., the stagnation condition where all velocity is
converted to pressure), surface efects will limit the dynamic pressure. The USBR (2007b) presents the
results of testing for a range of fow velocities, joint widths, and ofset dimensions, which is intended
to be a refnement of the estimation of associated potential uplif forces where open joints are a
concern. However, the uplif associated with the impoundment level must be estimated separately
based upon site-specifc conditions. For example, the uplif pressure for dynamic forces is typically
50 to 75 percent of the stagnation pressure for joint ofsets ranging between and inch (e.g., at an
average fow velocity of 35 feet per second, measured pressure was approximately 10 feet of head as
compared to a 14-foot stagnation value for a -inch ofset and -inch joint). Such uplif forces should
be considered in situations where lining failure would trigger impoundment failure. Measures for
reducing or controlling the potential development of uplif pressures include:
Grouting to control foundation seepage, where applicable.
Installation of an underdrain system consisting of perforated pipe within a graded
sand and gravel flter to control seepage.
Use of rigid foam insulation between the concrete spillway and underdrain system
to prevent freezing in cold climates.
Installation of embedded waterstops in foor joints to prevent the fow of water
through the joints to the foundation.
Use of longitudinal reinforcement and transverse cutofs at joints to prevent relative
displacement.
Increasing the weight of the channel lining.
< PREVIOUS VIEW
9-110
Chapter 9
MAY 2009
For channels on rock foundations where the above measures may not be efective, structural
design of the channel lining with anchors to resist uplif pressure should be employed.
9.7.4.3 Diversion and Collection Ditches
Diversion and collection ditches are important because the collection and control of runof from a
refuse embankment surface, from the slopes of an embankment, and from hillsides draining to and
away from the embankment should minimize environmental damage to downstream waters, prevent
damage to the embankment, and reduce maintenance and repair eforts related to site erosion. This is
particularly true when a refuse disposal facility is reclaimed and the drainage facilities must function
with limited repair for a long period of time. Basic procedures for designing diversion and collection
ditches are the same as those for other types of channels, as discussed in Sections 9.7.1 and 9.7.3.
The following are the most important considerations in the design of drainage channels:
Diversion ditches should be designed to reduce the amount of water reaching the
disposal facility during moderate storms when it can be shown that they will appro-
priately reduce operating or environmental concerns. It is seldom practical to design
diversion ditches that will not fail during very large storms such as the PMF.
The constantly changing size and confguration of refuse disposal facilities ofen
makes it necessary to provide short-term runof diversion and collection ditches at
intermediate development stages. Erosion protection criteria for these structures
must be established on a case-by-case basis and may difer from the erosion protec-
tion procedures for more permanent ditches and channels.
When a diversion or collection ditch will later serve as a permanent channel, the
capacity and type of erosion protection material should meet the requirements for
the permanent installation.
When a collection ditch will also serve to route the design storm discharge from an
impoundment outlet (by receiving the discharge from a decant or open-channel spill-
way), its capacity and durability should be based on the impoundment design storm.
In this instance, the concern is that the collection ditch must be able to function with-
out damage to the integrity of the embankment for storms up to the impoundment
design storm.
9.7.4.4 Culverts
Culverts are typically used beneath access roads that permit runof from minor storms to pass with-
out loss of road use. The appropriate design storm for such culverts depends on the importance of
the roadway to the overall operation and the efort and cost of repairing the road if it is overtopped
or washed out. An exception might be a roadway to an impounding embankment that must remain
open during or immediately following a large storm. However, the design criteria established for
impoundments in Section 9.5 are based on the assumption that it is not practical to maintain access to
the impoundment during that time period.
The basic design requirements for culverts were previously discussed in Sections 9.7.1 and 9.7.2.
9.7.4.5 Natural Streams
A natural stream fowing adjacent to a coal refuse embankment can cause signifcant damage if
the water at food elevation can reach the refuse embankment toe and cause erosion. Flow in
natural streams can be determined based on the principles of open-channel fow (Section 9.7.2.2),
although the analysis to determine maximum fow depth may be more involved due to the irregu-
< PREVIOUS VIEW
9-111
Hydrology and Hydraulics
MAY 2009
lar cross sections of natural streams. Computer sofware for estimating the food levels along natu-
ral streams is readily available.
As with constructed channels, the fow in a stream can be controlled by upstream or downstream
conditions, depending on the stream slope, confguration and discharge. Normally, the fow may be
assumed to be uniform and Mannings equation (Equation 9-17 or 9-18) can be used to approximate
the fow depth and velocity, except immediately upstream of bridges, road embankment crossings,
natural channel constrictions, etc. that cause backwater efects requiring the use of open-channel
profle analysis for determination of fow depth and velocity. The roughness coefcient n for natural
streams may be in the higher ranges shown in Table 9.13 due to vegetation including trees, varia-
tions in alignment, and the irregularity of cross sections. The hydraulic radius can be calculated in a
manner similar to that for a constructed channel, except that the areas and weted perimeter must be
determined from topographic maps or cross-section drawings.
Based on the peak runof to a natural stream, the approximate water depth can be calculated. If the
computed fow depth indicates that the water in the stream could encroach upon the refuse embank-
ment toe, provisions to prevent embankment erosion or disruption of the facility hydraulic system
should be employed.
9.8 RESERVOIR ROUTING
As part of the design of the hydraulic system for an impounding refuse disposal facility, reservoir
routing analyses are typically performed to determine the outlet spillway discharge and impound-
ment storage requirements. This is critical for open-channel spillway systems, but is also important
for sites that rely on storage of the design storm and discharge of the runof through the decant,
because these facilities have 10 days to discharge the storm infow in accordance with the criteria pre-
sented in Section 9.5. The methodology and key parameters for routing analyses are described herein,
and references to the computer sofware typically employed are provided. For specifc applications,
frequent reference is made to texts and publications on hydraulic design and engineering such as
Chow (1959), USBR (1987a), and the USDA National Engineering Handbook (1956).
9.8.1 Basic Routing Methodology
Reservoir routing is performed by analyzing the infow hydrograph (Section 9.6), the storage capac-
ity of the impoundment (Section 9.7), and the discharge-head relationship for the spillway outlet
(Section 9.7) to determine the reservoir level and spillway outfow hydrograph. The spillway outlet
may consist of a conduit system (decant or principal spillway), emergency open-channel spillway, or
a combination of conduit and open-channel spillway. Flood routing analyses should be based upon
an initial impoundment water level no lower than the lowest functional decant inlet. For impound-
ments that rely solely on a conduit system, the majority of the runof from the design storm must be
stored within the impoundment because of the limited discharge capacity of the conduit. However,
the conduit must be capable of discharging 90 percent of the runof within a 10-day period following
the design storm.
Impoundments with an emergency open-channel spillway generally have signifcant discharge capac-
ity, and thus less of the design storm runof must be stored within the reservoir. The open-channel
spillway inlet will be above the normal pool level, such that some initial storage or accumulation of
runof from the infow hydrograph occurs before outfow through the open-channel spillway is initi-
ated. Subsequently, the spillway discharges at a rate dependent on the reservoir level, which in turn
is a function of the infow hydrograph and storage capacity. Afer the peak of the infow hydrograph
passes, the reservoir level will continue to rise until the infow rate and spillway outfow capacity are
equal. Thereafer, the reservoir level will decline as spillway discharge becomes predominant. This
relationship is shown in the hydrographs presented in Figure 9.3. The development of the reservoir-
< PREVIOUS VIEW
9-112
Chapter 9
MAY 2009
storage relationship is discussed in Section 9.6.1.4, and the spillway-discharge relationship is dis-
cussed in Section 9.7.2.
USBR (1987a) and the USDA National Engineering Handbook (1956) present mathematical procedures
for computation of food routing, using an iterative process to arrive at the outfow hydrograph.
Computer programs such as HEC-1 and HEC-HMS are frequently employed to perform the routing
analysis USACE (1998b, 2000).
9.8.2 Basic Routing Parameters
An important factor that diferentiates coal refuse disposal facility impoundments from other types
of impoundments is that the embankment and impoundment confgurations continually change
with both the disposal of coarse refuse on the embankment and the disposal of fne refuse slurry into
the impoundment. These efects must be accounted for in the hydraulic design and reservoir routing
for a refuse disposal facility.
Figure 9.46 presents a sectional view of an impoundment facility illustrating features that impact the
routing of foods and the development of design parameters, including normal pool elevation, mini-
mum pool level, surcharge storage, normal freeboard and design-storm freeboard.
The design-storm freeboard for an impounding embankment is a function of the wave height and
the wave run-up conditions at the upstream face of the embankment. Guidance for the evaluation
of design storm freeboard for reservoirs is presented in USBR (1987a). Coal refuse impounding
embankments are typically required to have a minimum design storm freeboard of 3 feet above
the maximum reservoir pool level associated with the design storm, consistent with a fetch of less
than 1 mile.
Since the crest elevation of a slurry impoundment can change frequently, the facility plans and
specifcations should include a graph or table that shows the maximum allowable normal pool
level and allowable spillway and decant levels for each stage of operation. To ensure that adequate
freeboard is available to handle the design storm, the normal pool level and spillway inlets must
not be raised until the appropriate crest elevation has been reached. The disposal of fne coal refuse
within an impoundment afects the reservoir storage capacity. While most of the accumulation
occurs below the impoundment operational water level, the slurry discharge results in the build-
up of deposits forming deltas or beaches above the pool level. While this impact is frequently
insignifcant for many valley-fll type impoundments, it can have an impact on routing and free-
board particularly at diked-type facilities. This loss of storage capacity is generally estimated based
DECANT
INLET
DESIGN-STORM FREEBOARD
NORMAL POOL
ELEVATION
NORMAL
FREEBOARD
SURCHARGE
SETTLED FINE
REFUSE
EARTH OR
COARSE REFUSE
MINIMUM POOL LEVEL
CONTROLLED BY PUMPING
REFUSE
EMBANKMENT
FIGURE 9.46 SECTIONAL VIEW THROUGH IMPOUNDING FACILITY
FIGURE 9.46 SECTIONAL VIEW THROUGH IMPOUNDING FACILITY
FIGURE 9.46 SECTIONAL VIEW THROUGH IMPOUNDING FACILITY
< PREVIOUS VIEW
9-113
Hydrology and Hydraulics
MAY 2009
upon the position and elevation of the planned slurry discharge and an assumed slope of the
deposit (typically between 1 and 3 percent).
9.9 DAM BREACH ANALYSIS AND INUNDATION MAPPING
9.9.1 Background
Impoundments are assigned a hazard-potential classifcation based on the consequences to down-
stream workers, residents and property in the event the dams were to fail. FEMA (2004a) classifes
dams as high hazard potential, signifcant hazard potential, or low hazard potential, and the
states typically have comparable classifcation systems. A high-hazard-potential dam or embank-
ment is one whose catastrophic failure would likely result in loss of human life. A signifcant-haz-
ard-potential dam or embankment is one whose failure would not be expected to result in the loss
of human life, but could cause substantial property damage. Minimal property damage would be
expected from the failure of a low-hazard-potential dam or embankment. As discussed in Section 3.1,
MSHA requires evaluation of the hazard potential for other impoundment breach pathways, such as
breakthrough into underground mines.
In practice the hazard-potential classifcation for a dam may be apparent from site conditions; for
example, a large impoundment located upstream from a populated area will likely be classifed as
having a high hazard potential. To aid in determining the hazard-potential classifcation and to assist
in preparation of an Emergency Action Plan, a dam-breach analysis is performed to determine the
downstream consequences of a dam failure.
An EAP should be prepared for high- and signifcant-hazard-potential dams and embankments, so that
procedures are in place for responding to an emergency at the dam and to conditions in the potentially-
inundated area downstream (Chapter 14.0). This is a requirement in several states, and MSHA encour-
ages EAP preparation for high- and signifcant-hazard-potential dams in order to protect the public that
would potentially be afected by a dam failure (MSHA, 2007). An important step in the preparation of
an EAP is to perform a dam-breach analysis so that the potentially-inundated area downstream from
the dam or refuse embankment can be defned. Scenarios for dam failure, methods of analysis, sofware
used for analysis, data requirements, and other aspects of the dam-failure fow release and determina-
tion of the resulting inundated area are discussed in the following paragraphs.
9.9.2 Failure Scenarios
Dam failures can occur in a number of ways, but most result from: (1) overtopping of the dam due
to inadequate spillway capacity, (2) failure of the dam structure as the result of an earthquake, or (3)
the fow of water through the embankment leading to development of a breach in the embankment
(i.e., piping).
The frst scenario is the most common, as there are many existing dams in the U.S. that do not have
a spillway capacity adequate to handle the most extreme rainfall events. Consequently, states are
requiring upgrades at these facilities to meet current requirements. These upgrades typically include
such remedial measures as raising the dam crest and increasing spillway capacity. In some cases,
measures such as armoring the dam with roller-compacted concrete are employed to allow overtop-
ping to occur while the structure remains intact.
Piping failures result from pathways through a dam embankment where seepage gradually increases
transport of fne materials until a point is reached where pore pressures are high over a relatively
large area and a breach initiates. Spillway/decant pipes extending through a dam embankment are
vulnerable locations for this phenomenon to occur, and care must be taken to minimize the potential
for seepage fow along these structures.
< PREVIOUS VIEW
9-114
Chapter 9
MAY 2009
A major earthquake can result in an increase in soil pore pressure and sliding failure, usually in the
upstream portion of the embankment. The crest of the dam drops during the embankment failure,
resulting in a breach. Upgrades to prevent this type of failure typically require major and very costly
repairs.
Catastrophic dam failures can also result from causes such as landslides, foundation failure, sabotage
or damage to operational equipment.
While some technical studies of breach formation have been carried out (Wahl, 1998), the composi-
tion of earth embankment dams is highly variable and there is litle actual data available for calibrat-
ing model results. Thus, breach development has not been accurately related to a specifc dam failure
scenario.
At the present time, modeling of potential dam failures usually involves two basic scenarios. The
frst is overtopping of the dam during high fow (Infow Design Flood) conditions and the second is
catastrophic failure of the dam on a day with normal (sunny day or fair weather) fow conditions.
These two scenarios provide a reasonable representation of the range of conditions resulting from the
possible failure modes.
Coal refuse disposal facility embankments are developed in stages over several years, ultimately
resulting in a massive structure sometimes 1,000 feet or more in width and several hundred feet high.
Dam breach analyses should consider intermediate stages (which may have a narrower embankment
cross section) as well as the fnal facility confguration, because the failure of an intermediate stage
could occur more rapidly and could result in a greater breach fow and thus more signifcant down-
stream inundation than breach of the facility in its fnal confguration.
An important consideration for coal refuse slurry impoundments is the volume of fne coal refuse
that could be released during a hypothetical dam failure. Based upon data from a wide range of tail-
ings dam failures, Vick (2000) estimates that, while on average about 25 percent of the impounded
contents are released, the release of impounded contents can approach 100 percent. Rico et al.
(2008) evaluated historical records of tailings dam failures and releases to identify factors afect-
ing the runout distance and peak discharge, fnding that, on average, one-third of the tailings and
water (to the post-failure level) was released. Overtopping from foods appears to mobilize and
release more tailings.
While considerable uncertainty exists, the above estimate appears to be consistent with observations
reported by Owens (2008) for incidents at coal refuse impoundments. The setled fne coal refuse
frequently remains in a sof or loose condition until sufciently consolidated and thus may be in a
fowable state. Michael et al. (2005) performed a literature review to evaluate the ability of fne coal
refuse to fow. While no specifc test data are available for fne coal refuse, evaluation of other tailings
materials led them to conclude that saturated refuse may be susceptible to high pore pressure and
static liquefaction when containment is breached. In order to conservatively estimate the amount of
fne coal refuse that could be released from a dam breach, some states prescribe consideration of all
water, slurry and setled fne coal refuse contained within the impoundment from the breach invert
to the crest, when computing downstream fows. Adoption of a reduced volume of setled fne coal
refuse will generally require site-specifc information concerning the consistency and resistance to
fow of the material.
In addition to failure of the dam, breakthrough of the impoundment into an underground mine may
represent another type of release pathway. This pathway could lead to signifcant discharges in streams
in other watersheds, depending on the alignment and extent of extraction of the underground mine.
< PREVIOUS VIEW
9-115
Hydrology and Hydraulics
MAY 2009
9.9.3 Analytical Methods
The failure of a dam results in a condition referred to as rapidly varied unsteady fow. This is a very
complex fow condition that can be modeled with computer sofware. However, programs that pro-
vide the most sophisticated modeling of rapidly varied unsteady fow can be difcult to use. Thus,
the popular programs used for dam breach analysis represent simplifcations to some degree of rap-
idly varied unsteady fow analysis. A listing of sofware frequently used for dam breach analysis and
the analytical approach employed is provided in Table 9.20.
HEC-1 is frequently used for dam breach analysis for mine impoundments, and the other sofware
listed in the table are less commonly employed. In terms of sophistication, HEC-RAS, DAMBRK and
FLDWAV are the most technically advanced and should provide more accurate results than HEC-1.
HEC-1 and HEC-RAS were developed by the USACE Hydrologic Engineering Center, while the later
three programs in the preceding table are National Weather Service programs.
HEC-RAS (River Analysis System) is a second generation program from the USACE Hydrologic
Engineering Center and is the successor program to HEC-2. It was frst released in 1995 and gained
unsteady fow analysis capabilities in 2000. The unsteady fow portion of the program was adapted
from UNET. The program has the capability of modeling mixed fow regimes and can account for
channel constrictions and of-channel storage.
DAMBRK was developed by Fread (1988) for modeling unsteady fow associated with dam breaches.
FLDWAV, which was introduced in 1998, is the successor to DAMBRK and DWOPER and provides
advanced capabilities over both programs. SMPDBK is based upon an approximate methodology, and
under some circumstances can provide results within 10 percent of the results provided by DAMBRK.
In terms of modeling accuracy, both HEC-1 and SMPDBK have clear limitations. The accuracy of
these programs diminishes in situations involving channel constrictions and resulting backwater.
The fnal version of HEC-1 was released in 1998, and the program has been replaced by HEC-HMS,
which ofers one-dimensional kinematic wave routing for dam breach analyses. This methodology
does not account for inertial and pressure forces, and the energy slope is assumed to be equal to the
channel slope. Thus, HEC-HMS is best suited to relatively steep channels and urban areas where
natural channels have been modifed to regular shapes and constant slopes.
FLDWAV is the most sophisticated program currently (2009) available. It utilizes fnite-diference
approximations to solve the Saint-Venant equations for one-dimensional unsteady fow and can
account for natural features such as of-channel storage and channel constrictions. The program is
capable of handling a wide range of channel confgurations and data input. However, FLDWAV
requires some calibration for optimum accuracy. Other programs such as HEC-RAS and predeces-
TABLE 9.20 INUNDATION ANALYSIS SOFTWARE
Program Method of Analysis
HEC-1
Muskingum-Cunge
Modifed Puls
HEC-RAS One-Dimensional Unsteady State Flow
SMPDBK Approximate Method
DAMBRK One-Dimensional Unsteady State Flow
FLDWAV One-Dimensional Unsteady State Flow
< PREVIOUS VIEW
9-116
Chapter 9
MAY 2009
sors to both HEC-RAS and FLDWAV are capable of providing adequate results depending upon the
nature of the breach, outfow hydrograph and downstream channel confguration.
GIS-based sofware is gaining in popularity in hydrology and hydraulics applications and has been
used in combination with unsteady fow analysis sofware. WMS (a GIS-based hydrologic model) can
be used in conjunction with sofware such as HEC-RAS and SMPDBK. BREACH and FLDWAV have
reportedly been used in combination with GIS-based sofware to assess dam breaches and inunda-
tion mapping. Also ESRI, the developer of ArcGIS, and the USACE Hydrologic Engineering Center
have worked together to create HEC-GeoRAS, which allows the results of food routing analyses to
be displayed in a GIS environment. Eventually, GIS-based models, either packaged with or used in
combination with sophisticated unsteady fow analysis sofware, will be the accepted approach for
dam breach modeling and presentation of results.
Another issue related to sofware selection is the available topographic data for the analyses, the level of
accuracy required, and the users familiarity with the sofware. It is common to obtain topographic and
cross-section information from USGS quadrangle maps with some feld observation and verifcation.
Generally, a high degree of accuracy is not required for defning inundation limits and for identifying
potential evacuation requirements, particularly in remote areas. Thus selection of less sophisticated
sofware for EAP development is quite ofen adequate. However, use of a breach analysis to support
a hazard-potential classifcation other than high hazard potential may require careful evaluation of
the assumptions incorporated into the sofware. As discussed previously, communication should be
maintained with dam safety regulators (both state and federal) relative to sofware usage.
9.9.4 Input Data
9.9.4.1 Breach Parameters
A breach in an earth embankment dam is generally assumed to be trapezoidal in shape and can be
defned by depth, width, side slopes and development time. A comprehensive study of breach devel-
opment was carried out by Wahl (1998) for earth dams. While an accurate depiction of breach develop-
ment is desirable, the process is highly variable and difcult to predict, and conservative assumptions
based on published guidelines are normally accepted. Discussions with agency personnel should
prove valuable in this regard. In situations where the downstream channel is relatively broad in com-
parison to the volume of the release, the use of conservative breach development geometry and time
may not mater, as discussed in Section 9.9.6. Some state agencies require that the release include the
entire volume of tailings. When a site-specifc evaluation is performed, it should be compared with
the following guidance for estimating the minimum volume of release for fne refuse:
An equal volume of fne refuse (as stored) and food water
One-third of the fne refuse stored
The volume of fne refuse released can then be used with other site-specifc factors to estimate the
breach depth and confguration and duration of outfow. Some guidelines for determination of
breach parameters (FERC, 1993) are presented in Figure 9.47. Physical limitations such as the width
and depth of the valley should be considered when applying the guidelines.
Programs used for routing dam breach fows (HEC-1 or HEC-RAS, DAMBRK, FLDWAV, etc.) typi-
cally have input parameters for defning the breach geometry and development time. These pro-
grams expand the size of the breach from zero to the full dimensions in the specifed development
time using an internal algorithm.
One sofware program for breach development is the National Weather Service (NWS) program
BREACH, which was developed by Fread (1988). This program is a physically-based breach simula-
< PREVIOUS VIEW
9-117
Hydrology and Hydraulics
MAY 2009
tion model, but concerns regarding the model have been raised (Wahl, 1998). The hydrograph calcu-
lated by BREACH can be used as the dam breach hydrograph by unsteady fow modeling sofware.
Coal refuse impoundments can include massive embankment stages, such that an intermediate
embankment stage confguration may represent a more critical breach geometry and time of failure
than the fnal development confguration. Modeling a partial failure of a coal refuse dam using pro-
grams such as SMPDBK may result in a peak outfow discharge that is higher than that from breaching
the full height of the dam all the way down to the natural valley botom. Since slurry impoundments,
especially upstream-construction dams, ofen contain a considerable amount of consolidated slurry
and have a relatively small water storage volume as compared to conventional water supply, food
control, or multipurpose dams, this may be more representative of the actual consequences of a
slurry dam failure. However, some state regulatory agencies require that all of the saturated fne coal
DAM CREST INITIAL BREACH
APPROXIMATE
FINAL BREACH
CONFIGURATION
BR
1
Z Z
1
FIGURE 9.47 BREACH PARAMETERS
HD
(ADAPTED FROM FERC, 1993)
FOR NON-ENGINEERED DAMS OF SLAG, REFUSE MATERIALS:
BR 0.8 x CREST LENGTH
>
>
1 Z 2
>
0.1 TFH 0.3 HOUR
>>
FOR ENGINEERED EARTHEN DAMS:
>
> 0.25 Z 1
>
0.1 TFH 1.0 HOUR
>>
HD BR 5HD >
WHERE:
BR = AVERAGE BREACH WIDTH
TFH = TIME TO FULLY FORM BREACH
Z = HORIZONTAL COMPONENT OF BREACH SIDE SLOPE
HD = HEIGHT OF DAM
FIGURE 9.47 BREACH PARAMETERS
FIGURE 9.47 BREACH PARAMETERS
< PREVIOUS VIEW
9-118
Chapter 9
MAY 2009
refuse be treated as fowable and that a breach analysis be based upon a breach extending the full
height of the embankment (from the fnal crest to the foundation).
9.9.4.2 Initial Conditions
Generally, the fow into an impoundment and in the channel downstream from the dam is assumed
to be constant and equal to the design storm condition immediately prior to dam breach. Frequently
the design storm is the PMF, which represents an extreme upper-bound infow to the impoundment
that will only rarely be approached (Section 9.5). Sometimes the design storm will be less than the
PMF, but as long as it exceeds the spillway capacity, the dam could be overtopped and catastrophic
failure could occur. The impoundment water surface elevation used in the dam breach model should
be the minimum required for breach initiation.
For a sunny day breach, normal steady-state stream fows into the impoundment and in the down-
stream channel should be assumed. These can usually be obtained from published stream gage
data or on-site records. Estimates of fow can be developed based upon channel dimensions and
slope, estimated roughness and calculated normal fow velocity in the absence of recorded fow
data. The water surface in the impoundment is typically assumed at normal pool elevation for a
sunny day breach.
9.9.4.3 Flow Channel Geometry and Roughness
All dam-breach fow models require a description of the downstream channel and foodplain
geometry (i.e., cross sections) and roughness. Roughness is usually defned in terms of Mannings
n, and values for Mannings n are available in the literature for a wide range of conditions (Chow,
1959). Typically, out-of-channel fow encounters substantial resistance from brush, trees, debris
and even dwellings, so that the Mannings n for the foodplain is much higher than for normal
channel (USGS, 1989).
A key factor that can cause elevated food levels is the presence of constrictions such as bridge and
railroad embankment crossings or severe natural channel narrowings in the reach downstream from
the dam. These can cause temporary backwater elevation and localized increased fooding. Addi-
tional cross section data (i.e., closer cross section spacing) may be needed in these constricted areas.
The presence of an existing downstream impoundment or impoundments may result in the need to
extend the fow model farther downstream.
Typically for a dam breach analysis under design storm conditions, the area of interest along the
channel terminates when the fow reaches a certain increment above the fow elevation without the
dam breach (typically 1 to 2 feet). For a sunny day breach, the area of interest along the channel
normally terminates when the fow returns from the foodplain to the natural channel or to a level
associated with a specifed recurrence interval food. Thus, downstream channel data should extend
past the points where these control points are anticipated to occur.
In addition to natural channels, another pathway for an impoundment breach is via breakthrough to
an underground mine, with discharge through the mine and out of associated mine openings. This
can lead to potential inundation in watersheds other than that in which the impoundment is located,
depending on the size and extent of the mine.
9.9.5 Results of Analysis/Inundation Mapping
The output from fow model sofware will be water surface elevations and fow velocity at each
channel cross section in the model. Additionally, the model sofware can provide the time of
< PREVIOUS VIEW
9-119
Hydrology and Hydraulics
MAY 2009
arrival of the food wave from the breach, which is useful for the EAP development. The maxi-
mum water surface elevation at each cross section following the dam breach will defne the extent
of inundation. It is important to note that the maximum inundation elevation for an unsteady
fow analysis does not occur at the same time at each cross section. Therefore care must be taken
to record the highest level at each cross section and to use these values in determining the limits
of inundation.
The fow velocity is also computed for each channel cross section, which may be helpful in assessing
the potential threat to occupied structures or roadway travelers, as discussed in Section 9.5.1.3.2. It
is important to note that while the average velocity in the foodplain may be provided as part of the
model output, it is advisable to perform an independent computation of the food velocity for the
structure location, considering the maximum water surface elevation and energy grade line.
As discussed in the previous section, the inundation map should extend downstream from the dam
to the farther of the termination points associated with a breach occurring under design storm and
sunny day conditions.
Ploting the extent of inundation can be tedious if done manually because the water surface ele-
vation is falling relative to a fxed datum and thus the extent of inundation will not match or be
parallel to any ground surface elevation contours. GIS-based sofware can provide plots of the
inundation limits as part of the normal output and thus eliminate the need for manual ploting.
9.9.6 Sensitivity Analyses
It may be useful to perform multiple dam breach analyses with variations in selected input data to
evaluate the efect of the variation on the analysis results. For dam-breach analyses, the parameters
associated with the breach development (i.e., dimensions and development time) are likely to be the
most controversial. Since a breach analysis can directly afect the safety of downstream residents, it
is prudent that conservative dam-breach analyses be performed. If, for example, substantially reduc-
ing the breach development time does not signifcantly alter the results in terms of inundation levels
and extent of inundation, then conservative breach assumptions can be used in the analysis and may
expedite regulatory review of the EAP.
9.9.7 Hazard Classifcation
Most commonly, a dam-breach analysis is performed as part of preparing an EAP for a dam that has
already been classifed as having high or signifcant hazard potential. Another purpose for a dam-
breach analysis could be to establish the hazard classifcation in the frst place. In this event, the FEMA
(2004a) or applicable state criteria should be followed. The FEMA criteria are listed in Table 9.21.
TABLE 9.21 HAZARD POTENTIAL CLASSIFICATIONS
Hazard-Potential
Classifcation
Loss of Human Life
Economic, Environmental,
Lifeline Losses
Low None expected
Low and generally
limited to owner
Signifcant None expected Yes
High Probable One or more Yes
(FEMA, 2004a)
< PREVIOUS VIEW
9-120
Chapter 9
MAY 2009
If populated areas are impacted, particularly areas located close to the dam, a high-hazard-potential
classifcation should normally be assigned. As discussed in Section 9.5.1.3.2, other criteria, such as
depth or velocity of fow, may also be considered by designers and accepted by regulatory agencies
for determining the potential signifcance of the inundation level and to assign the associated hazard
classifcation. Other classifcations should be consistent with Table 9.21 and the limits of inundation.
< PREVIOUS VIEW
Chapter 10
ENVIRONMENTAL
CONSIDERATIONS
10-1 MAY 2009
Environmental considerations for coal refuse disposal facilities generally involve potential impacts to
streams and wetlands, air quality, and water quality. Impacts to streams and wetlands originate with
facility siting. Air quality issues arise from dust and burning associated with coal refuse embank-
ments. Water quality issues are typically related to the generation of acid leachates by coal refuse or
to erosion and sedimentation at refuse surfaces or disturbed areas under development. Liner sys-
tems have been used to provide protection of groundwater, and reclamation of coal refuse disposal
embankments can mitigate air and water impacts.
Federal and state air and water quality regulatory programs govern site discharges and must be
considered in coal refuse disposal facility design. Thus, review of applicable regulatory programs
and permit requirements should precede the design of coal refuse disposal facilities. Similarly, liner
systems are generally regulated by states.
In light of the above, this chapter provides a general discussion of environmental issues associated
with coal refuse disposal facility design, construction, and reclamation.
10.1 STREAMS AND WETLANDS
Coal refuse disposal facilities ofen impact streams and wetlands regulated by the Clean Water Act
(CWA). This legislation was originally enacted in 1972 and was subsequently amended in 1977. When
a planned coal refuse disposal facility will impact streams and wetlands, several types of permits
and certifcations may be required by CWA regulations. Although the U.S. Environmental Protection
Agency (USEPA or EPA) has regulatory authority over the CWA, the permits and certifcations may
be administered and enforced by other federal, as well as state or local agencies. These agencies may
include the USEPA, the U.S. Army Corps of Engineers (USACE), the U.S. Fish and Wildlife Service
and state Departments of Environmental Protection (state DEPs).
The CWA was enacted with the intent of restoring and maintaining the chemical, physical and bio-
logical integrity of the waters of the United States. The term waters of the United States includes
the following:
1. All waters which are currently used, or were used in the past, or may be susceptible
to use in interstate or foreign commerce, including all waters which are subject to the
ebb and fow of the tide;
< PREVIOUS VIEW
10-2
Chapter 10
MAY 2009
2. All interstate waters including interstate wetlands;
3. All other waters such as intrastate lakes, rivers, streams (including intermitent
streams), mudfats, sandfats, wetlands, sloughs, prairie potholes, wet meadows,
playa lakes, or natural ponds, the use, degradation or destruction of which could
afect interstate or foreign commerce including any such waters:
i. Which are or could be used by interstate or foreign travelers for recre-
ational or other purposes; or
ii. From which fsh or shellfsh are or could be taken and sold in interstate
or foreign commerce; or
iii. Which are used or could be used for industrial purpose by industries in
interstate commerce;
4. All impoundments of waters otherwise defned as waters of the United States under
the defnition;
5. Tributaries of waters identifed in paragraphs 1-4 of this section;
6. The territorial seas;
7. Wetlands adjacent to waters (other than waters that are themselves wetlands) identi-
fed in paragraphs (a) (1)-(6) of this section.
Waste treatment systems, including treatment ponds or lagoons designed to meet
the requirements of CWA (other than cooling ponds as defned in 40 CFR 123.11(m)
which also meet the criteria of this defnition) are not waters of the United States.
8. Waters of the United States do not include prior converted cropland. Notwithstand-
ing the determination of an areas status as prior converted cropland by any other
federal agency, for the purposes of the Clean Water Act, the fnal authority regarding
Clean Water Act jurisdiction remains with the EPA.
Coal refuse disposal facilities that impact waters of the United States must be permited and certifed
under the federal regulations outlined in the CWA. The various sections of the CWA regulate the
activities described below:
Section 401 Water Quality Certifcation
This section of the CWA requires that any applicant for a federal permit to construct
and operate a coal refuse disposal facility that may result in the discharge of any pol-
lutant must obtain certifcations for those activities from the state in which the dis-
charge originates. This certifcation is referred to as the Water Quality Certifcation
for the project.
Section 402 NPDES Regulations
The 1972 amendments to the CWA established the National Pollutant Discharge
Elimination System (NPDES) permit program to control discharges of pollutants
from point sources. The NPDES permit may be administered and enforced by a local
USEPA branch or state DEPs. Some states have additional requirements for storm-
water discharges that may impact planned coal refuse disposal facilities and are not
covered by the CWA.
Section 404 Dredge/Fill Permiting
This section of the CWA established a permit program to regulate the discharge
of dredged or fll material into waters of the United States. This permit program
is administered by the USACE under a memorandum of agreement between the
< PREVIOUS VIEW
10-3
Environmental Considerations
MAY 2009
Department of the Army and the USEPA. Under Section 404 of the CWA an individ-
ual or general permit may be needed based on the proposed activities.
In addition to the Clean Water Act, other statutes and regulations such as the Surface Mining and
Reclamation Act (SMCRA) of 1977 and the Safe Water Drinking Act (1974) may be applicable to coal
refuse disposal facilities with respect to streams and wetlands. These regulations may result in addi-
tional permiting not covered by the CWA. The Ofce of Surface Mining (OSM), U.S. Department
of the Interior, is responsible for the national program to regulate the surface efects of coal mining
activities, although each state may take on primary responsibility if the states regulatory program is
approved by the OSM.
Consideration should be given early in the design process to the permits and certifcations required
for coal refuse disposal facilities as they relate to streams and wetlands. The time involved in the
permiting process is typically lengthy and must be accounted for in coal refusal disposal facility
design. Agencies such as the USEPA, USACE, U.S. Fish and Wildlife Service, state DEP, and local
municipalities should be contacted prior to permit preparation to determine what permits and cer-
tifcations will be required and which agencies will administer and enforce them. Once the required
permits are determined, it may be benefcial to hold a pre-submital meeting with the appropriate
agencies. Afer the meeting, the permit applications should be submited in a timely manner, allow-
ing for responses to permit application comments. Some states have moved to a combined applica-
tion process, although generally permit applications are submited separately and at various times
during the design process.
10.2 AIR QUALITY
Coal refuse disposal can create two types of air quality problems: (1) fugitive dust and particulate
mater and (2) noxious gases originating from burning refuse embankments. Fugitive dust becomes
airborne due to wind and coal refuse handling and placement. Sources may include: emissions from
haul roads; wind erosion from exposed surfaces, storage piles and spoil piles; reclamation opera-
tions; and other material or earth disturbance activities. Fugitive dust can be ingested by humans
and animals and can also be harmful to vegetation. High concentrations of sulfur dioxide associ-
ated with the combustion of coal refuse are toxic to nearby vegetation. Also, sulfur dioxide, organics
(polynuclear aromatic hydrocarbons such as benzo(a)pyrene), and metals (mercury and arsenic) are
harmful if inhaled in signifcant volumes by humans.
Dust is regulated as an air emission by state DEPs or, if no approved state program exists, by the
USEPA. If amendments are being considered or co-disposal with combustion waste is planned, dust
control requirements can take on greater signifcance than with normal construction. If accidental
combustion occurs at coal refuse disposal facilities, air emissions can become a signifcant health and
safety concern, and methods to address burning may need to be developed and implemented as part
of a remedial action.
The following sections discuss measures for controlling dust and for reducing the potential for com-
bustion or controlling burning should it occur.
10.2.1 Dust Control
The transportation and placement of coal refuse can create a considerable amount of fne particu-
late mater that is susceptible to wind erosion. Coal refuse is compacted and crushed by machinery
during placement and further deteriorates through physical weathering and chemical decomposi-
tion. When refuse-related dust problems occur, they can be mitigated by stabilizing the surface layer
of the refuse. This can be accomplished by applying water or a dust suppressant solution over dis-
turbed areas, establishing windbreaks of trees or hedgerows that alter both the direction and the
< PREVIOUS VIEW
10-4
Chapter 10
MAY 2009
velocity of wind over the refuse material, or performing reclamation by covering and vegetating the
disturbed surface (Coalgate et al., 1973).
In situations where a relatively quick dust control procedure is needed or where vegetation is for
some reason impractical, stabilization has been achieved using various commercially available chem-
ical agents. Chemical seals have been accomplished through application of: (1) a lime chip-sodium or
potassium silicate topdressing over the refuse material, (2) a resinous or bituminous-base adhesive,
(3) calcium, ammonium and sodium lignin sulfonates and bark extracts, (4) resin and wax emulsions
or neoprene, and (5) elastomeric organic polymers (Coalgate et al., 1973; Dean and Havens, 1972;
and Eigenbrod, 1971). When applying such products to areas such as haul roads that will experience
truck or heavy-equipment trafc, the efect on traction should be considered.
Erosion control mats that have plant seeds incorporated within the binding material have been suc-
cessfully used to vegetate disturbed construction areas and to control dust.
10.2.2 Combustion Control
Current practices in the mining industry have virtually eliminated coal refuse fres. The reason is
two-fold. First, the amount of coal in coal refuse has been greatly reduced because of more efcient
removal of coal during mining and processing. Secondly, current embankment construction practices
involve thorough compaction of refuse material, thus restricting the fow of air and moisture that can
create a favorable environment for heat generation. Thus, the discussion provided herein is mainly
applicable to older existing embankments.
Components of air emissions from burning coal refuse may include carbon, nitrogen, sulfur com-
pounds and metals such as arsenic and mercury. These emissions can impact human health and the
environment. Air emissions along with elevated temperatures can degrade existing vegetation and
make establishment of new vegetation impossible.
Coal refuse embankment fres have been caused by spontaneous combustion and in some instances
from careless burning of trash or other debris. Coal refuse fres have also been intentionally started
to obtain red dog material for use as a road construction base or have been accidentally ignited by
natural causes such as lightning or forest fres. Historically, the most common cause of coal refuse
fres has been spontaneous combustion resulting from the self-heating tendencies of coal. The poten-
tial for spontaneous combustion is greatly increased if oxidizing materials such as pyrites are present
and if these oxidizing materials are wet (Coalgate et al., 1973; Mihok and Chamberlain, 1968; Nicho-
las and Hutnik, 1971).
Self-heating of coal refuse generally occurs due to exposure of organic and carbonaceous materials to
moisture and oxygen, creating reactions that generate heat. When the generation rate of heat exceeds
the rate of heat loss, temperatures within a refuse pile can reach the ignition temperature of the
remaining coal and carbonaceous materials. The generation rate of heat is a function of the concentra-
tion of reactants (thermophillic bacteria, carbon and oxygen), surface area of the pile, particle sizes of
the coal refuse and ambient air temperature (Kim and Chaiken, 1990). When coal refuse is exposed
to water and oxygen, heat can be generated from the respiration of bacteria up to a temperature of
about 120 to 170 degrees Fahrenheit ( F), when the bacteria die. Beyond this temperature range, oxi-
dation of carbon and carbonaceous materials has to occur if the ignition temperature of coal (in the
approximate range of 620 to 788 F for bituminous coal and 842 to 950 F for anthracite coal) is to be
reached (Maneval, 1969).
In addition to creating air quality problems, burning refuse embankments can also create potentially
dangerous working situations. The most common of these is the creation of burned-out voids or pock-
ets within the interior of the refuse embankment that can lead to surface cave-ins and/or hazardous
slides. Atempts to extinguish smoldering refuse facilities with water can cause violent explosions if
< PREVIOUS VIEW
10-5
Environmental Considerations
MAY 2009
these burned-out voids become flled with pressurized steam. Explosions can also occur in the vicinity
of burning material as a result of airborne coal dust produced during the handling of coal refuse.
Under current disposal conditions, the likelihood of coal refuse igniting is extremely low because of
low pyrite and/or coal content. When coal refuse is spread and compacted in lifs in stable embank-
ments, fres rarely occur. Other standard construction practices that should be followed for mitigat-
ing combustion potential include:
Prior to placement of any coal refuse material at a new site, all vegetation and other
combustible materials should be removed from the area where refuse will be placed.
All refuse materials with high pyritic and coal content should be compacted as the facil-
ity is constructed, and all large rocks should be crushed or removed to a separate loca-
tion to prevent the creation of air pockets in the embankment (Coalgate et al., 1973).
If present, waste materials with high pyritic and coal content should be allowed to
weather separately prior to their placement at a refuse facility in order to lessen the
chance of a thermal buildup due to oxidation.
If oxidation is a potential problem, coal refuse facilities should be designed and con-
structed in a manner that minimizes the amount of exposed surface area in order to
decrease the air infltration (Coalgate et al., 1973).
Typically, detection of burning is based upon on-site visual observation (i.e., noting the presence or
absence of smoke and/or sulfur dioxide fumes). However, there is no inexpensive means of detecting
overheated refuse materials below the embankment surface prior to their combustion. Methods that
have been used to detect combustion of conditions leading to combustion include:
Gas Emission Monitoring Carbon monoxide (CO) is a by-product of coal refuse
oxidation and can be detected very early in the oxidation process. Surface monitor-
ing of CO emissions can thus indicate the potential for spontaneous combustion.
Hydrogen sulfde (H
2
S) is also a by-product of coal oxidation. Concentrations of this
noxious gas will be present prior to combustion and can also be detected through
monitoring (Chamberlain and Hall, 1973; Chamberlain et al., 1970; Guney, 1968).
Direct Thermal Monitoring The internal temperatures of refuse embankments can
be monitored by inserting temperature probes into driven pipes or drilled holes. The
temperature buildup associated with oxidizing refuse material can thus be profled.
Remote Sensing Thermal and optical images from an airborne platform can be
used to identify the location, depth, size and propagation of hot spots and fres
(Zhang et al., 2004). Landsat TM imagery and airborne thermal scanner data have
been employed in remote sensing studies for measuring ground surface tempera-
tures. The surface temperature data can then be used for estimating the extent and
depth of coal fres using thermodynamic models.
Electrical Resistivity Geophysical Survey Some researchers have employed surface
DC electrical resistivity for distinguishing burnt sedimentary rock with relatively
high resistivity from non-impacted sedimentary rock. The burnt rock has a higher
porosity, more cracks and lower water content, which allows it to be distinguished
from the non-impacted rock.
10.2.3 Refuse Fire Extinguishment
Extinguishing coal refuse fres is normally not a problem confronting engineers and designers of new
coal refuse facilities. However, when an existing facility is being modifed or added to, fre abatement
< PREVIOUS VIEW
10-6
Chapter 10
MAY 2009
can be an important part of the engineering and design process. Fire extinguishment can also be a
critical consideration when a refuse embankment is being prepared for abandonment.
Studies have determined that refuse embankment fres generally burn in a temperature range
between 600 and 2000 F. It has also been found that once refuse materials have reached a tempera-
ture of approximately 200 F, either through spontaneous heat buildup or through heat transfer from
adjacent areas, they will eventually self-ignite given favorable conditions such as an abundant supply
of air and moisture (Magnuson and Baker, 1974).
Since the reactions that create heat are inherently variable, no single safe temperature has been
identifed below which heat buildup and refuse ignition will not occur. Ignition temperatures vary
with each embankment and with location within the embankment and are largely a function of
available air and the site-specifc characteristics of the coal refuse. It is therefore not enough to
extinguish the burning portion of a refuse embankment. Steps must also be taken to: (1) lower the
temperature of the refuse below the point of re-ignition and (2) eliminate embankment conditions
that could lead to temperature buildup and future re-ignition.
Temperatures in coal refuse embankments that are sufcient for combustion have been measured at
depths of 100 feet or more. However, at that depth the amount of available oxygen is minimal and
ignition will not occur. If, however, hot spots are exposed through the excavation of overburden
or through some other embankment modifcation, the additional available oxygen may cause these
areas to ignite. Critical extinguishment depths are therefore related to site-specifc conditions and
may be afected by future actions that may alter these conditions.
As indicated previously, the most critical concerns facing those atempting to extinguish a coal refuse
fre are the unique dangers involved in using water and in excavating materials in ways that may
cause airborne dust. Explosions that can result from such practices can hurl hot debris over nearby
areas and can lead to failure of the refuse embankment. Similarly dangerous are smoldering internal
voids created when a refuse embank ment burns. These areas of potential cave-in can be extremely
dangerous to workers and fre fghters alike. Carbon monoxide poisoning is also a danger.
Despite these potential dangers, a number of fre-fghting techniques have proven successful in cer-
tain situations. For purposes of discussion these techniques can be grouped into three general catego-
ries: (1) physical removal of the burning refuse, (2) quenching and/or sealing by surface treatment,
and (3) quenching and/or sealing by injection into the burning refuse. These methods are briefy
discussed in the following paragraphs and are also summarized in Table 10.1.
10.2.3.1 Excavation and Removal
Excavation and removal has historically been the predominant method for extinguishing refuse
embankment fres (Kim and Chaiken, 1993). This approach has several variations, each generally
involving the removal of burning materials from the refuse embankment. The removed materials
may be extinguished by quenching, cooling, and sufocation, or they may simply be allowed to burn
out. This method can be efectively used when the burning areas are relatively small and accessible
and when removal activities do not adversely afect embankment stability. Extreme care must be
taken to minimize airborne coal dust when handling burning refuse materials. This dust can ignite
and cause violent explosions. Also, any time that equipment is working over burned-out areas, there
is a danger that large voids created by the fre will collapse under the weight of the equipment. Varia-
tions of the excavation and removal approach include:
Excavation Small and readily-accessible burning areas can be extinguished by
removing the burning refuse material from the embankment using construction
equipment. The removed material can then be extinguished through quenching, or
< PREVIOUS VIEW
10-7
Environmental Considerations
MAY 2009
TABLE 10.1 FIRE EXTINGUISHMENT TECHNIQUES
Method Brief Description Limitations
P
h
y
s
i
c
a
l
R
e
m
o
v
a
l
Excavation
Burning refuse excavated from embankment;
extinguished or allowed to burn itself out; facility
regraded and sealed
Dust and noxious fumes
Access to burning material
Possible cave-ins
Weakens refuse facility
Water cannons
Water cannons used to dislodge and quench
burning refuse; quenched material replaced and
recompacted on refuse facility
Source of quenching water
Weakens refuse facility
Potential for dust explosion
Isolation
Burning zone isolated by excavating trenches;
burning zone quenched or buried with inert sealing
material; trenches reflled with inert material
Access to burning material
Controlled
burnout
Burning refuse is allowed to burn under monitored
and controlled conditions
Access to burning material
Duration is uncertain
Weakens refuse facility
S
u
r
f
a
c
e
T
r
e
a
t
m
e
n
t
Blanketing or
sealing
Entire burning embankment covered with mantle of
clay or soil; compacted; burning is smothered
Limited to small facilities
Maintaining seals integrity
Possible cave-ins
Source of clay or soil
Foam covering
Entire refuse facility is sealed with a commercial
foam blanket; oxygen denied the refuse; burning is
extinguished
Facility size
Maintaining a seal
Cant use where burning is near
surface
Rice paddy
technique
Suited for fat refuse areas; dikes constructed
around perimeter and area fooded; water
percolates into burning zone; fre quenched.
Supply of water
Possible cave-ins
Slow
Stability
Water sprinklers
Burning refuse facilities are wet-down or
saturated by a system of sprinklers until burning is
extinguished
Water source
Saturation weakens structure
Reignition possible
I
n
t
e
r
n
a
l
T
r
e
a
t
m
e
n
t
Multiple well-
point system
Horizontal insertion of perforated metal piping
near base of embankment; water injected; pipes
removed and reinserted in higher strata; process
repeated for total structure
Source of quenching water
Slow
Weakens refuse facility
Slurry injection
Vertical or angle holes drilled into burning
embankment at various depths; liquid slurry
injected into burning voids; steam vent pipes
inserted; heat reduction monitored
Slow
Stability
it can be allowed to burn at a safe distance from the refuse embankment. Once the
burning material has been removed from the embankment, the excavated portion
should be backflled, regraded, compacted and covered with a sealing material that
will limit air fow (Coalgate et al., 1973; Jolley and Russell, 1959). A major drawback
to this approach is that machinery operators may be exposed to large doses of nox-
ious and toxic gases that are dangerous if exposure is prolonged. Health and safety
monitoring, air monitoring and use of personal protective equipment are required
for this activity.
Water cannons Water cannons similar to those used by fre departments have been
used to dislodge and quench burning refuse materials when they are near embank-
< PREVIOUS VIEW
10-8
Chapter 10
MAY 2009
ment surfaces. Removal of the quenched material can be accomplished by: (1)
hydraulic sluicing using a water cannon, (2) excavation by dragline, and (3) loading
on trucks for dumping elsewhere. For all three alternatives the extinguished material
should be re-spread and compacted in accordance with facility plans and specifca-
tions. The use of this technique is contingent upon the availability of water and the
stability of the embankment during hydraulic excavation (McNay, 1971).
Isolation Burning materials can be isolated from the remainder of the refuse facility
by cuting trenches around them. To eliminate heat transfer, such excavations should
be at least 6 feet wide and should extend into the embankment foundation. Once the
burning material is isolated, it can be extinguished with water, by applying a sealant,
or by burying under a blanket of non-combustible material. The exposed trench faces
should be sealed with clay or fne-grained soil to restrict air fow, or the trenches
should be backflled with non-combustible material such as soil. To prevent heat
transfer from the burning portion of the embankment to non-burning areas, sand or
other heat-conducting material should not be used as backfll (Coalgate et al., 1973;
Jolley and Russell, 1959).
To mitigate the potential for explosions, excavations into refuse materials that are known or suspected
to be burning must be performed with extreme care if hot or burning materials will be exposed to
airborne coal dust and/or moisture. Through monitoring, areas of high material temperature can be
mapped (if boreholes are used, they should be sealed to prevent airfow). Excavation should be per-
formed in stages and monitored with the intent of avoiding opening up burning areas to moisture and
coal dust in confned spaces. Work should proceed downwind (from upwind areas) using equipment
that can operate from above and away from burning areas. Upon completion of the excavation, backfll
materials should be placed in lifs and compacted, which will minimize the potential for rekindling.
10.2.3.2 Surface Treatment
The methods described in this section require that the embankment be relatively small and have
accessible slope faces. Basically, surface treatment involves sealing of the entire surface of an embank-
ment to restrict air fow to the fre. The primary problem with surface seals is maintaining them
until sufcient cooling has occurred to prevent re-ignition. This maintenance period can exceed 20
years (Kim and Kociban, 1994), which is greater than the efective life of many types of surface seals.
Common surface treatment methods are described in the following:
Blanketing or sealing In some instances, it may be practical to extinguish burning
refuse by blanketing the entire embankment with about 2 feet of non-combustible
material such as fy ash, clay or other soil. This cover should be compacted as it is
applied, thereby smothering the burning refuse. Breaks in the seal can occur through
water erosion, heat cracks, cave-in of burned-out voids, or even wind erosion (Coal-
gate et al., 1973; Jolley and Russell, 1959; McNay, 1971; Myers et al., 1966). In extreme
cases, where the need to extinguish an embankment fre exceeds normal economic
constraints, commercial foam sprays (e.g., polyurethane) have been applied (Magnu-
son and Baker, 1974).
Rice-paddy technique This procedure is only suited for large, stable, fat-topped
refuse facilities. Since minimal fumes and dust are created, it is ideal for sites located
near residential areas. Dikes are constructed around the top perimeter of the burning
refuse facility and at appropriate intermediate locations. Each diked area or pond is
then fooded, and the impounded water percolates into the embankment. Draglines
can be used periodically to stir the botoms of the ponds to increase the rate of per-
colation. The use of this fre-abatement procedure is dependent upon an abundant
< PREVIOUS VIEW
10-9
Environmental Considerations
MAY 2009
supply of water and is further dependent upon the ability of the burning embank-
ment to support earth-moving equipment during dike construction (Coalgate et
al., 1973; McNay, 1971). The impact of dike construction and water irrigation on the
stability of the coal refuse embankment must be evaluated prior to implementation
of this method.
Water sprinklers In some instances, water sprinklers have been used to wet down
burning embankments and to provide a continuous supply of water over and
through the refuse material. The success of this procedure is largely dependent upon
the hydraulic conductivity of the embankment, and vertical drilling may be required
to increase percolation into the embankment interior. The saturation of an impound-
ing embankment can be dangerous, as its stability may be greatly reduced (Coalgate
et al., 1973; Myers et al., 1966).
Surface treatment methods should be implemented sequentially with monitoring of explosion and
emission hazards, particularly if concurrent or subsequent excavation activities are planned, as pre-
viously discussed.
10.2.3.3 Water and Slurry Injection
This approach involves injection of water or slurry into the burning zones under pressure. The
injected material quenches and smothers the burning material. The use of an injection method can
ofer one or more of the following advantages:
While usually more expensive on a unit volume basis, injection is well suited to spot
treatment of smaller burning areas within a larger embankment in contrast to exca-
vation and removal or surface treatment, which require remedial work over a much
larger area.
Inaccessible areas on steep slopes can be treated. Pipes can be driven with air ham-
mers while other equipment (mixers, pumps, etc.) can be placed at a nearby level
location.
Men and equipment do not have to work directly over burning areas.
There are basically two types of injection methods:
Multiple well-point system This procedure entails driving perforated pipes in a
single horizontal plane near the toe of the embankment and pumping water into
the pipes. The pipes are placed relatively close to each other (approximately 2 feet
on center) so that the injected water thoroughly saturates the entire zone. Once the
burning is extinguished in that zone, the pipes are withdrawn and then re-inserted
a short distance above their previous location. Water is again introduced to extin-
guish the fre in this new area. This procedure is repeated until all the burning areas
within the embankment are extinguished. It should be emphasized that in order to
minimize the potential for re-ignition of the refuse material, this procedure should
progress from the botom of an embankment upward. Because the burning portion
of the embankment becomes saturated, the use of this method is not recommended if
stability is an issue.
Slurry injection When slurry is injected into an embankment, voids and air
channels are blocked and air access is restricted (McNay, 1971). The slurries most
commonly used are suspensions of fy ash, limestone dust, vermiculite, sodium
bicarbonate or mine drainage sludge in water. Pipes are typically driven vertically
into the burning zone on 10- to 15-foot centers. Slurry is injected under low pres-
< PREVIOUS VIEW
10-10
Chapter 10
MAY 2009
sure (usually 10 to 15 psi) to depths of 40 feet or more. When the slurry is no longer
accepted, the pipes are raised and injection is resumed. The interior or deepest por-
tion of the burning zone is treated frst in order to prevent further penetration of the
fre. Injection then progresses toward the surface of the embankment. Because of the
danger of explosions, open pipes should be inserted next to the injection holes to
vent steam. Use of cryogenic slurry consisting of liquid nitrogen and granular carbon
dioxide to enable quick cooling of the burning material has been proposed. Some
initial testing demonstrating the ability of this approach to lower temperatures over
an extended period was conducted (Kim and Kociban, 1994).
10.3 WATER QUALITY
As indicated previously, coal refuse facilities can substantially degrade the quality of water in nearby
drainage courses if they are improperly constructed. In addition to adversely afecting surface-water,
drainage from refuse facilities can also afect the groundwater. Although a variety of water quality
problems can be created by coal refuse drainage, the most common efects are: (1) increased turbidity
and suspended solids and (2) water quality degradation due to acidic leachates (Martin, 1974).
Water pollution problems created by coal refuse can be substantial. Coal refuse leachates can be
acidic, can contain elevated concentrations of metals such as iron, aluminum and manganese, and
can also be corrosive. When leachates enter a stream, aquatic environments may be greatly altered
and desirable organisms may be reduced or eliminated entirely. When refuse leachates percolate into
the groundwater, aquifers can be signifcantly impacted. The following sections provide a discussion
of mine refuse water quality issues and various procedures and techniques for controlling and/or
mitigating their adverse efects.
10.3.1 Erosion and Sedimentation
Erosion and sedimentation control plans must be submited to state and local regulatory authorities
as part of refuse disposal facility designs. These plans typically include a variety of measures for
diverting drainage from disturbed areas, for controlling erosion, and for removing sediment from
runof before release of surface water from the refuse disposal site. As part of these plans, efuent
monitoring programs are typically established to verify that erosion and sedimentation control mea-
sures are efective.
10.3.1.1 Prevention
When coal refuse and earthen materials are exposed to weathering, erosion and sedimentation can
occur. The following practices can be implemented to minimize erosion and sedimentation:
Stripping of vegetation from a disposal site should be limited to only the area that
is needed for construction. Future fll areas should be stripped immediately prior to
construction.
Topsoil that is removed from a construction area and stockpiled for future use
should be stored in a manner that minimizes erosion and should be revegetated as
soon as possible.
During the construction process, care should be taken to preserve vegetation on
areas surrounding the disturbed construction area.
Collection ditches and sedimentation ponds should be constructed at the down-
stream end of the construction site.
All fll material exposed during construction should be graded in a manner that min-
imizes the potential for runof over the downstream face of the embankment. This is
particularly important for the crest and downstream face of the refuse embankment.
< PREVIOUS VIEW
10-11
Environmental Considerations
MAY 2009
Completed embankment surfaces should be reclaimed and vegetated as soon as
practical, while accommodating seepage control measures such as extension of
underdrains or installation of collection and discharge systems at the embank-
ment toe.
10.3.1.2 CONTROL
Control procedures for reducing the amount of suspended material entering streams are presented
in the following subsections.
10.3.1.2.1 SEDIMENTATION PONDS
Sedimentation ponds are structures designed to intercept and retain water-borne sediment and
debris. They are primarily intended for use during construction prior to the establishment of efec-
tive vegetation on the disturbed area. Sedimentation ponds should be sized and constructed in accor-
dance with criteria prescribed by state mining regulation agencies. These structures normally do not
retain water for long periods and are usually maintained with low water surface levels except follow-
ing rainfall. Engineering design criteria and standards for sedimentation ponds have evolved from
requirements for surface mining operations. In most instances, these standards are also applicable to
coal refuse (Davis, 1973).
OSM rules for sedimentation ponds under 30 CFR 816.46 to 49 generally include the following:
Sedimentation ponds can be used individually or in series.
They should be located as near as possible to the disturbed area and not in perennial
streams.
They should provide adequate detention time to meet efuent standards and should
contain or treat the runof from the 10-year, 24-hour precipitation event.
They should provide sediment storage capacity with periodic sediment removal suf-
fcient to maintain adequate volume.
Ponds with embankments that meet or exceed the impoundment size criteria
or other conditions indicated in 30 CFR 216 (20 acre-feet capacity or 20 feet in
height) should have principal and emergency spillways designed to safely pass the
runof from a 100-year precipitation event or larger, depending upon the hazard
potential classifcation. For ponds that do not meet or exceed the impoundment
size criteria, the principal and emergency spillways should be designed to safely
pass runof from the 25-year precipitation event or greater, as specifed by the state
regulatory authority.
State agencies generally provide additional guidance regarding determination of the sediment stor-
age capacity and may require specifc design storm parameters or values for sizing the principal and
emergency spillways.
In situations where very fne particulate material is suspended in the refuse drainage, the amount
of time required for natural setlement or clarifcation in a setling basin can be long. If the drainage
is carrying a signifcant volume of suspended solids, clarifcation can be accelerated through use
of chemical focculants. This practice may also be considered when the capacity of a sedimentation
pond is relatively small.
Sediment/sludge removal is required in order to sustain sedimentation pond capacity. In the event
that such removal is not practical, sedimentation ponds should be designed with a capacity large
enough to accommodate sedimentation over the appropriate operating period.
< PREVIOUS VIEW
10-12
Chapter 10
MAY 2009
10.3.1.2.2 Sediment Traps and Check Dams
Sediment traps and check dams may be useful as intermediate structures between erosion sources and
sedimentation ponds or can be employed where sedimentation ponds are prohibited or unfeasible.
They should be located within site drainage structures and should not cause channel overfow under
design fow conditions. Design and installation should be in accordance with state regulations.
10.3.1.2.3 Silt Fences
Silt fences are temporary structures for detaining sediment-laden overland (sheet) fow long enough
that the larger-sized particles are deposited and silt-sized particles are fltered out. State regulatory
publications provide design and construction guidance for silt fences, and manufacturers provide
similar information for their products. The following are general guidelines for silt fences:
The drainage area should not exceed 0.25 acres per 100 feet of silt fence length.
For slopes between 50:1 and 5:1, the maximum allowable upstream fow path length
to the silt fence should be 100 feet.
The flter material should be able to retain at least 75 percent of the sediment.
The botom edge of the silt fence should be tied or anchored into the ground to pre-
vent underfow.
There should be no ponding behind silt fences.
Silt fences should be regularly maintained.
Appropriate state guidelines should be reviewed prior to installation of silt fences.
10.3.1.2.4 Erosion Control Blankets and Reinforcement Mats
Erosion control blankets can be used to stabilize freshly seeded slopes and drainage or ditches until
such time that a cover of vegetation is established. Typically, they are most efective on slopes up to 3:1
and in drainage ditches with slopes up to 20:1. Erosion control blankets typically degrade within 6 to 24
months of installation, depending on their composition (straw, fber, and plastic systems). Design and
installation guidance are available in state regulatory publications and manufacturers literature.
Reinforcement mats are similar to erosion control blankets, but provide greater protection because
of the use of synthetic fbers that reinforce vegetation and result in more erosion-resistant construc-
tion. Reinforcement mats are used for steep slopes (greater than 3:1) and channels with slopes in the
range of 15:1 to 10:1. Design and installation guidance are available in state regulatory publications
and manufacturers literature.
10.3.1.2.5 Vegetation
Erosion and stream turbidity are best minimized by establishing a protective layer of vegetation
on embankment slopes and along exposed ditch surfaces. The establishment of grasses in drainage
ditches reduces fow velocity and, consequently, erosion.
Vegetation covers on embankment slopes are not practical until construction has proceeded far enough
that relatively stable slope conditions are achieved. Vegetation is further discussed in Section 10.5.5.
10.3.2 Acid Generation and Control
The potential for acid generation from coal refuse materials can be estimated, and measures can be
implemented to control acid formation or migration. State regulatory programs vary in terms of pre-
diction methodology and the measures required to control or contain acid mine drainage.
< PREVIOUS VIEW
10-13
Environmental Considerations
MAY 2009
10.3.2.1 Background
Acid generation is principally the result of pyrite oxidation. Pyrites are commonly associated
with coal formations and surrounding strata. Several types of pyrites may be present, and the
reactivity of diferent forms varies signifcantly (Kleinmann, 2000). Acidity is produced by the
oxidation of pyrites (sulfde components and iron components), which leads to the dissolution
of metals (ferric iron, manganese, and aluminum, and occasionally other metals such as copper,
zinc, and nickel). Rock strata may contain carbonate materials that neutralize acidity; however,
coal refuse is material segregated from coal and generally includes minimal overburden materi-
als that will neutralize acidity.
Acid mine drainage is a major problem in the northern Appalachian Basin (particularly within the
Allegheny Group stratigraphic section) and less signifcantly in the Midwest (Kleinmann, 2000; Appa-
lachian Regional Commission, 1969; Wetzel and Hofman, 1989). Kleinmann (2000) provides a discus-
sion of geology, hydrology and prediction of acid generation, including acid-base accounting (static or
whole rock analysis) and simulated weathering tests (kinetic testing such as leaching tests in various
columns and chamber arrangements). Testing procedures associated with acid-base accounting can be
applied to individual samples of overburden and spoil materials for predicting acid generation or rec-
lamation performance. Table 10.2 presents a summary of suggested criteria for interpreting the results
of acid-base accounting analysis. While simulated weathering tests are not routinely used for coal
mine drainage prediction, they can provide data for estimating the relative concentrations of net acid-
ity, metals and sulfate, and they can be useful for evaluating the efectiveness of various amendments
for mitigating problem water quality conditions. Kleinmann (2000) provides a detailed discussion of
criteria for determining whether to conduct kinetic testing as well as testing methods.
Mitigation of acid generation can also be accomplished by hydrologic controls that minimize water
contact with air and refuse. This typically involves: (1) compaction of the refuse surface, (2) sealing
of the refuse surface and diversion of runof from active disposal areas, and (3) capping and covering
of completed refuse disposal areas. The USEPA (2000) developed a best management practices guid-
ance manual for remining of refuse disposal sites providing specifc guidance related to erosion and
sedimentation controls and mitigation of acid generation.
10.3.2.2 Grading, Compaction and Sealing
Grading, compaction and sealing of coal refuse embankment surface areas will minimize the poten-
tial for infltrating water contacting pyrites and thus reduce the potential quantity of acid genera-
tion and groundwater migration. Grading facilitates control of surface water fows, and compaction
reduces the hydraulic conductivity of the refuse material. Regular sealing of the refuse embankment
surface using smooth-drum rolling equipment facilitates runof and thus reduces infltration and
the generation of acid leachates. Before subsequent placement of additional lifs, the sealed surface
should be scarifed to enhance bonding between lifs and to minimize potential stratifcation.
10.3.2.3 Amendments
A number of amendments for neutralizing acidity have been used with coal refuse, including coal
combustion waste (lime-containing materials), kiln dust, phosphate rock, lime and other prod-
ucts. The amount of amendment material required for neutralizing acidity is a function of several
factors, as described by Kleinmann (2000), USEPA (2000), and Brady et al. (1998). Stewart et al.
(1997, 2001) evaluated neutralization and leaching from various blends of combustion waste and
acid-producing refuse based upon a series of multi-year unsaturated column experiments. With
sufcient combustion ash (20 percent and greater for the cited ash and coal refuse), no evidence
of acid conditions was detected and low levels of most metals were observed, although high
concentrations of boron and sulfate were reported. In column tests where the combustion ash
< PREVIOUS VIEW
10-14
Chapter 10
MAY 2009
TABLE 10.2 SUMMARY OF SUGGESTED CRITERIA FOR INTERPRETING
ACID-BASE ACCOUNTING
Criteria Application References
Rocks with NNP less than -5 parts/1000
considered potentially toxic.
Coal overburden rocks in northern
Appalachian basin for root zone
media in reclamation.
Smith et al., 1974, 1976; West
Virginia Surface Mine Drainage Task
Force, 1979; Skousen et al., 1987
Rocks with paste pH less than 4.0
considered acid toxic.
Coal overburden rocks in northern
Appalachian basin for root zone
media.
Smith et al., 1974, 1976; Surface
Mine Drainage Task Force, 1979
Rocks with greater than 0.5% sulfur
may generate signifcant acidity.
Coal overburden rocks in northern
Appalachian basin, mine drainage
quality.
Brady and Hornberger, 1990
Rocks with NP greater than 30
parts/1000 and fzz are signifcant
sources of alkalinity.
Coal overburden rocks in northern
Appalachian basin, mine drainage
quality.
Brady and Hornberger, 1990
Rocks with NNP greater than 20
parts/1000 produce alkaline drainage.
Coal overburden rocks in northern
Appalachian basin. Base and
precious metal mine waste rock and
tailings in Canada.
Skousen et al., 1987;
British Columbia Acid Mine Drainage
Task Force, 1989;
Ferguson and Morin, 1991
Rocks with NNP less than -20
parts/1000 produce AMD.
Base and precious metal mine
waste rock and tailings in Canada.
British Columbia Acid Mine Drainage
Task Force, 1989; Ferguson and
Morin, 1991
Rocks with NNP greater than 0 do not
produce acid. Tailings with NNP less
than 0 produce AMD.
Base and precious metal mine
waste rock and tailings in Canada.
Patterson and Ferguson, 1994;
Ferguson and Morin, 1991
NP/MPA ratio less than 1 likely results
in AMD.
Base and precious metal mine
waste rock and tailings in Canada.
Patterson and Ferguson, 1994;
Ferguson and Morin, 1991
NP/MPA ratio classifed as less than
1 (likely AMD), between 1 and 2
(possible AMD), and greater than 2 (low
probability of AMD).
Base and precious metal mine
waste rock and tailings in Canada.
Ferguson and Robertson, 1994
Price et al., 1997
Theoretical NP/MPA ratio of 2 needed
for complete acid neutralization.
Coal overburden rocks in northern
Appalachian basin, mine drainage
quality.
Cravotta et al., 1990
NP/MPA ratio used with NP threshold
to determine confdence levels for acid
producing samples. 80% confdence of
no acid production if NP/MPA ratio of
6.5 and NP threshold of 3.3%.
Coal overburden samples from 4
states: PA, WV, TN, and KY.
Bradham and Caruccio, 1995
Use actual NP and MPA values as
well as ratios to account for buffering
capacity of the system.
Base metal mine waste rock, United
States.
Filipek et al., 1991
Note: NP = Neutralization potential
NNP = Net neutralization potential
MPA = Maximum potential acidity
(ADAPTED FROM KLEINMANN, 2000)
< PREVIOUS VIEW
10-15
Environmental Considerations
MAY 2009
was insufciently alkaline or where insufcient ash was combined with the refuse, acid genera-
tion ultimately exceeded the neutralizing alkalinity of the ash, resulting in a decline in pH and
increased concentrations of metals. Stewart et al. (2001) recommend that careful atention be paid
to balancing the acid-generating potential of refuse with the alkalinity of combustion ash. Some
practitioners recommend increasing the alkalinity by some factor in order to prevent acidic condi-
tions (Daniels et al., 1996).
Daniels et al. (2002) evaluated various combustion ash and coal refuse mixing strategies (including
layering and partial blending) to determine their efectiveness in reducing acidity; they demonstrated
the value of blending in alkaline materials as close as possible to the area where acid generation is
occurring. Rich and Hutchison (1990) discuss the use of kiln dust for neutralizing combined coal
refuse. Use of limestone, oxides, phosphate rock and other materials for neutralization is addressed
by Skousen et al. (1998).
If amendments are used, provisions should be included in the design plans to verify that proper
placement and/or mixing are achieved. The efect of amendments on the geotechnical characteristics
of the refuse materials, particularly the strength and hydraulic conductivity of materials placed in
structural embankment zones, should be assessed, as discussed in Sections 5.1.5 and 6.2.3.5.
10.3.2.4 Reclamation and Vegetative Cover
Reclamation and vegetative cover following completion of disposal operations provides drainage
control and limits contact of the coal refuse with infltrating water. The USEPA (2000) provides a
qualitative discussion of improvements in the control of acid generation associated with reclama-
tion and vegetation and cites supporting quantitative studies. Gentile et al. (1997) describe a cover
system for an Illinois refuse disposal facility consisting of a compacted clay liner and protective soil
cover designed to reduce infltration by 84 percent. Meek (1994) describes the use of a PVC liner that
reduced acid loads from a spoil pile by 70 percent.
While placement of barriers to infltration as part of reclamation can address acid generation, provi-
sions such as drainage systems should also be incorporated, so that internal seepage can discharge
from the toe of a refuse embankment without raising the phreatic surface.
10.3.3 Water Quality Control
10.3.3.1 Diversion of Runof
Drainage from undisturbed portions of a watershed should be conveyed around coal refuse disposal
facilities to the extent practical using diversions. Thus, the amount of drainage contacting coal refuse
and potentially subject to water quality impacts will be minimized. State regulatory guidelines pro-
vide criteria for the design and construction of diversion systems for control of runof from undis-
turbed areas. While use of diversion ditches for impoundments can assist with controlling runof,
their capacity can only be considered in the impoundment food routing if they are designed and
constructed to handle the associated impoundment design storm (e.g., the Probable Maximum Flood
for a high-hazard potential impoundment).
10.3.3.2 Treatment
Treatment of acid mine drainage typically involves neutralization of acidity and precipitation of
metal ions to meet applicable efuent standards (USEPA, 1983). To meet the required standards, a
variety of treatment methods including active and passive treatment technologies can be employed.
Selection of an active treatment system involves evaluation of the fow rate, pH, total suspended
solids, acidity/alkalinity, iron and manganese concentrations, the receiving streams fow rate and
< PREVIOUS VIEW
10-16
Chapter 10
MAY 2009
use, availability of electric power, the distance from the point of chemical addition to the point where
the water enters a setling pond, and the volume and confguration of the setling pond. Most active
chemical treatment systems consist of an infow pipe or channel (sometimes a raw water storage
pond and aerator for large fows), a storage tank or bin for treatment chemicals, a chemical metering
system, a setling pond for precipitated metal oxyhydroxides, and a discharge point for treated water.
Table 10.3 presents a summary of chemical compounds used for acid mine drainage (AMD) treatment
and an equation for estimating the quantity of chemicals required based on the stream fow and the
acidity of the AMD. Aeration enhances oxidation of metals such that chemical treatment is more ef-
cient. Oxidants and pH adjusters are also sometimes used in the oxidation process to enhance metal
oxyhydroxide precipitation and reduce metal foc volume. Mechanical surface aerators are generally
used for large fows where aeration is required; simpler aeration systems using gravity to cascade
water over rocks or splash blocks may be useful in smaller applications. Chemicals for neutralizing
acidity are generally selected based on technical and cost factors. Skousen et al. (1998) discuss active
treatment system design and costs and provide case studies.
Passive treatment technologies that take advantage of naturally occurring chemical and biologi-
cal processes to cleanse impacted water and do not require continuous chemical inputs have been
developed. The primary passive technologies include constructed wetlands, anoxic limestone drains
(ALD), vertical fow systems such as successive alkalinity producing systems (SAPS), limestone
ponds, and open limestone channels (OLC). Table 10.4 presents design factors and references for pas-
sive treatment systems. Skousen et al. (1998) discuss passive treatment system design and costs and
provide case studies.
10.3.4 Water Quality Impacts on Construction Materials
The corrosive nature of coal refuse and leachates from coal refuse makes construction material selec-
tion important if facility appurtenant structures are to function as intended for long periods of time
TABLE 10.3 CHEMICAL COMPOUNDS USED IN AMD TREATMENT
Common Name Chemical Name Formula
Conversion
Factor
(1)
Neutralization
Effciency
(2)
Limestone Calcium Carbonate CaCO
3
1.00 50%
Hydrated Lime Calcium Hydroxide Ca(OH)
2
0.74 95%
Pebble Quicklime Calcium Oxide CaO 0.56 90%
Soda Ash Sodium Carbonate Na
2
CO
3
1.06 60%
Solid Caustic Soda Sodium Hydroxide NaOH 0.80 100%
20% Liquid Caustic Soda Sodium Hydroxide NaOH 784 100%
50% Liquid Caustic Soda Sodium Hydroxide NaOH 256 100%
Ammonia Anhydrous Ammonia NH
3
0.34 100%
Note: 1. The conversion factor may be multiplied by the estimated tons of acid per year to get tons of chemical
needed for neutralization per year. For liquid caustic, the conversion factor gives gallons needed for
neutralization.
2. Neutralization effciency is an estimate of the relative effectiveness of a chemical in neutralizing AMD
acidity. For example, if 100 tons of acid per year is the amount of acid to be neutralized, then 78 tons of
hydrated lime would be needed to neutralize the acidity in the water (100 x 0.74 / 0.95).
(ADAPTED FROM SKOUSEN ET AL., 1998)
< PREVIOUS VIEW
10-17
Environmental Considerations
MAY 2009
including abandonment. Table 11.6 lists common construction materials used for facility appurte-
nant structures and corrosion or deterioration mechanisms. The potential for chemical reaction and
for clogging of drainage materials are critical considerations in the design of drainage systems. For
drainage structures that are in contact with coal refuse or leachate, measures such as sulfate-resistant
cement and coatings applied to metal surfaces should be used, as appropriate.
10.3.5 Hydrogeology
Groundwater recharge, unsaturated groundwater fow and saturated groundwater fow are hydro-
geologic mechanisms that can afect migration of coal refuse constituents from a refuse disposal
site. Groundwater fow is the primary migration mechanism, as erosion and sedimentation control
measures are generally capable of controlling overland fow processes. Table 10.5 presents an over-
view of the hydrogeologic process and signifcance of the saturated and unsaturated groundwater
regimes. Some hydrogeologic features and their efect on the design of coal refuse disposal facili-
ties include the following:
Springs To minimize the potential for contact of water with coal refuse, natural
hillside spring fows should be collected and controlled. Spring collection drains
provide a means to collect and convey spring water from the source to down-
stream locations.
Mine discharges and underground mine workings In some instances, discharges
to and from mines may be important hydrogeologic features, because mines collect
and convey groundwater. Similar to springs, discharges from mine openings can be
controlled by collection drains that convey the mine water from the source to down-
TABLE 10.4 DESIGN CONSIDERATIONS FOR AMD PASSIVE TREATMENT SYSTEMS
Treatment
System
Raw Water
Conditions
Construction
Design Factors to Size
Treatment System
References
Aerobic Wetland Net alkaline water
Overland fow,
cattails planted in
substrate
10 to 20 g Fe/m
2
/d
0.5 to1.0 g Mn/m
2
/d
Hedin et al. (1993)
Horizontal-
Flow Anaerobic
Wetland
Net acidic water,
generally low fow
rate
Horizontal fow
above organic
substrate
3.5 g acidity/m
2
/d
Hydraulic conductivity of
substrate generally 10
3
to 10
4
cm/sec
Rate of sulfate reduction
(~300 mmoles/m
3
/day)
Hydraulic loading
Hedin et al. (1993),
Eger (1994),
Wildeman et al.
(1993)
Anoxic Limestone
Drain (ALD)
Net acidic water
DO, Fe
3+
, Al < l.0
mg/l
Horizontal fow
through buried
limestone
15 hours contact time
6- to 15-cm-diameter
limestone
Lifetime limestone
consumption
Hedin et al. (1994)
Successive
Alkalinity
Producing
Systems (SAPS)
Net acidic water
Vertical fow
through an
organic layer
overlying a
limestone bed
15- to 30-cm organic matter
with adequate permeability
15 hours contact time in
limestone
Lifetime limestone
consumption
6- to15-cm-diameter
limestone
Kepler and
McCleary (1994)
(SKOUSEN ET AL., 1998)
< PREVIOUS VIEW
10-18
Chapter 10
MAY 2009
stream locations. Impoundments may require construction of a barrier to control
fow of slurry into the mine workings. Additionally, the underground workings may
act as a sink for groundwater migration, including seepage from the impoundment.
Groundwater Groundwater fow beneath a disposal site may be afected by seep-
age from refuse materials. If adverse water quality impacts are anticipated, liner
systems and amendments can be used to mitigate these concerns.
Surface water Surface-water bodies may be a recharge source or receiving body.
Disposal sites located near surface water bodies or impounding facilities may require
measures such as liners, cutofs, and other barriers to protect the hydrogeologic regime.
In addition to provisions for protecting the hydrogeologic regime, state regulatory agencies will
require monitoring systems for detecting potential impacts to groundwater quality. This require-
TABLE 10.5 OVERVIEW OF THE HYDROGEOLOGIC PROCESS
Source and Flow Process Description References
Recharge
Recharge into the disposal facility may occur from:
Infltration of precipitation and runoff
Seepage from impoundment waters
Kleinmann, 2000
Unsaturated fow in
embankment materials
Unsaturated fow in embankment materials is
infuenced by the recharge rate and unsaturated
hydraulic conductivity and generally migrates
vertically toward saturated embankment zones.
Hutchison and Ellison,
1992
Saturated fow in embankment
materials
Saturated fow in the embankment materials is
infuenced by underlying barriers such as foundation
materials and internal drainage structures designed
to control phreatic levels. Saturated fow generally
migrates horizontally along foundation surfaces,
although a component of fow can be into foundation
soils. A liner system may be employed to restrict this
component of fow.
Hutchison and Ellison,
1992
Groundwater fow in embank-
ment foundation soils
Saturated groundwater fow in foundation soils is
infuenced by underlying aquicludes or bedrock
barrier and generally migrates horizontally along such
surface, although a component of fow can be into
deeper horizons or bedrock. Monitoring well systems
are typically employed to monitor groundwater quality
conditions beyond the limits of disposal sites.
Hutchison and Ellison,
1992
Groundwater fow in bedrock
fracture system
Saturated groundwater fow in the bedrock is
infuenced by the fracture system (and, in some
cases bedrock primary porosity) and generally
migrates horizontally toward groundwater discharge
zones. Monitoring well systems may be employed to
monitor groundwater quality conditions beyond the
limits of disposal sites.
Kleinmann, 2000
Groundwater interaction with
underground mines
Saturated groundwater fow may interact with
underground mines, which may act as a discharge
zone. Flow may follow discharge gradients in
response to coal seam dip or pressure head within
the mine. Monitoring well systems may be employed
to monitor groundwater quality conditions.
< PREVIOUS VIEW
10-19
Environmental Considerations
MAY 2009
ment is usually satisfed through installation of monitoring wells located upgradient and down-
gradient from the disposal facility. Guidance for monitoring programs is typically available from
state regulatory agencies. General guidance for groundwater monitoring systems is provided by
Hutchison and Ellison (1992).
10.4 LINER SYSTEMS
Site-specifc factors that should be considered in liner system design are summarized in Table 10.6.
Liner systems are generally used for containment in situations where acid generation from coal
refuse may impact the groundwater. Liner systems are an option in addition to amendments that can
be considered for neutralizing acid generation.
Liner systems for protection of groundwater are cited in some state regulatory guidance for coal
refuse disposal. Generally the reference is to a single-component, low-hydraulic-conductivity layer.
Liner systems employed for other waste containment systems such as combustion waste (DiGioia et
al., 1995) generally comprise multiple layers. The layers from the botom up typically include: (1) sub-
TABLE 10.6 SITE-SPECIFIC FACTORS TO CONSIDER IN LINER SYSTEM DESIGN
Potential Waste Material Toxicity
Chemical properties of refuse and coal preparation additives
Net acid generation potential
Soluble constituents for anticipated environmental conditions
Special treatment or neutralization procedures utilized
Total mass of soluble constituents
General Water Resource Values at Site
Adequate quality for benefcial use
Adequate quantity for benefcial use
Existing or identifed benefcial uses
Probable locations of future benefcial uses
Leachate Availability to the Environment
Waste material characteristics
Thickness of waste
Site climatic conditions
Provisions at closure to restrict infltration
Site Factors
Topography
Geology, including predictability of uniformity and/or potential for discontinuities
Unsaturated zone thickness, continuity, hydraulic conductivity and natural water content
Potential migration time for seepage to groundwater
Effects of climatic conditions on long-term unsaturated zone mitigation characteristics
Constituent attenuation potential
Waste Disposal Facility Management Practices
Facility type
Waste placement method
Protection of liner system from environmental damage
Controls on the hydraulic head
Risk reduction practices such as placement of underdrains, sub-aerial deposition, limited time of operations
Non-liner barriers such as cutoff walls
Installation of special early warning monitoring systems
(HUTCHISON AND ELLISON, 1992)
< PREVIOUS VIEW
10-20
Chapter 10
MAY 2009
grade or cushion materials, (2) leak detection zone, (3) liners (primary, secondary and/or composite),
and (4) leachate collection layer.
10.4.1 Design Requirements
Design and performance requirements for liner systems are generally determined by the following
(Hutchison and Ellison, 1992):
Waste material characteristics including chemical composition, grain-size distribu-
tion, hydraulic conductivity, and the presence of free liquids.
Waste disposal facility characteristics including liner hydraulic conductivity,
slope of the liner, depth and slope of waste placed on the liner, waste placement
method, hydraulic head controls, and the duration of operation for all or portions
of the facility.
Site characteristics including location and depth of the water resource to be pro-
tected, unsaturated zone conditions, and climatic conditions.
The potential for release of leachate is a function of the magnitude of the hydraulic head above the
liner, the thickness and efective hydraulic conductivity of the liner material (considering the fre-
quency of discontinuities in the liner such as cracks or holes), and the length of time the hydraulic
head is applied to the liner. Leachate from coal refuse is generally not reactive with liner materials,
but, if organic chemicals or strong bases are used in the coal preparation process and remain present
in the waste, the issue of liner material compatibility may need to be addressed.
A liner system generally consists of a single low-hydraulic-conductivity layer (clay soil or geosyn-
thetic material). Clay soil liners may include an overlying protection layer to protect the liner from
erosion and desiccation. Where a geosynthetic material is used, an underlying cushion layer and an
overlying protection layer are usually employed to minimize the potential for penetrations. Addi-
tionally, single liner systems may include overlying hydraulic head controls such as a pervious layer
above the liner. Such systems reduce the head on the liner and thus further limit potential migration
of leachate from the disposal facility. Composite double liners and leachate collection and removal
systems are used when redundant systems are needed, although such liner systems are not generally
used at coal refuse disposal facilities. Table 10.7 summarizes materials and handling and construction
procedures associated with individual components of liner systems.
Major considerations in choosing materials for soil liners are availability and composition. Soils
must contain a sufcient portion of clay material such that the constructed liner has low hydraulic
conductivity, high plasticity, and chemical stability. Suitable soils are usually classifed CL, CH, or
SC in the Unifed Soil Classifcation System (USCS) with a liquid limit between 35 and 60 and a
plasticity index of 10 or greater. Material for soil liners can consist of on-site or local borrow mate-
rials, imported bentonite, or mixtures thereof. To achieve the low hydraulic conductivity required
for a containment layer, soils must have consistent properties and may need thorough mixing,
preprocessing (e.g., removal of rocks, breakdown of soil clods, addition of bentonite), condition-
ing (e.g., adjustment of water content), placement in controlled lifs, and compaction. Imperfec-
tions such as gravel zones, organics and roots should be removed during construction. Protection
against cracking from drying or shrinking may also be required.
The required thickness and hydraulic conductivity of the barrier layer are a function of the hydraulic
heads, refuse material characteristics, and state policies or regulations. Typically, requirements for a
soil liner or barrier layer are a minimum thickness of 2 feet and a hydraulic conductivity less than 5 x
10
-5
cm/sec ( 50 f/yr) and in some applications less than 1 x 10
-7
cm/sec ( 0.1 f/yr). Variations from
these criteria are generally dependent upon in-situ conditions.
< PREVIOUS VIEW
10-21
Environmental Considerations
MAY 2009
TABLE 10.7 AVAILABLE MATERIALS OR PROCEDURES FOR LINER SYSTEM COMPONENTS
A. Low-Hydraulic-Conductivity Liners
A.1 Low-Hydraulic-Conductivity Natural Soil or Rock Natural soils or rock may be used as a low-hydraulic-
conductivity liner so long as it is possible to demonstrate by feld investigations that the material is continuous
and of suffcient thickness and properties over the entire area requiring the liner. This demonstration may be
particularly diffcult for rock because of jointing and fracture conditions.
A.2 Constructed Low-Hydraulic-Conductivity Liners Low-hydraulic-conductivity liners that are constructed
beneath a mine waste disposal facility may consist of any of the following materials:
Compacted, low-hydraulic-conductivity soils (e.g., clayey-silt to clay depending upon the required
hydraulic conductivity)
Soil and bentonite or cement mixtures
Pre-formed fexible geotextile impregnated with bentonite or pre-formed, granulated bentonite laminated
to a geomembrane, referred to as geosynthetic clay liner (GCL)
Pre-formed fexible membrane liners made from a variety of available polymeric material, generally
referred to as geomembranes; varying in thickness from about 20 to 100 mils
Field-applied liners, varying from about 80 mils of spray-on asphaltic materials to 6 inches of
conventionally-placed asphaltic materials
Composite liners, consisting of combinations of soil and geomembrane low-hydraulic-conductivity layers
A.3 Waste Material Settled or mechanically placed tailings often have a low hydraulic conductivity and can be
used as part of the long-term liner system, provided the tailings serve one of the low-hydraulic-conductivity liner
functions.
B. Cushion or Liner Protection Materials
B.1 Geotextiles Synthetic geotextile materials varying in weight from 4 to 20 ounces per square yard may be
used above or below geomembranes to protect against penetrations from rock particles due to loads from
construction activities or the weight of the waste material. The suitability of a geotextile to act as a cushioning
layer varies and is defned by the method employed by its fabrication (needle-punch non-woven versus woven).
B.2 Fine-grained Soil for Geomembrane Protection Soils varying from clay to sand can also be used to protect
most geomembranes from equipment traffc or static loading of the waste material. Small gravel-size material
has also been used to protect thick geomembranes. The protective soil must be relatively free of large rock
particles that could cause stress concentrations on the liner.
B.3 Cover Material for Clay Liner Protection Cover protection may also be required for a compacted soil liner if
the liner could be subjected to extreme loads, such as construction equipment traffc, or exposed to drainage or
desiccation.
C. Hydraulic Head Control Components
C.1 Free-Draining Gravel Layer Several inches of free-draining gravel (including coarse sand) are usually
adequate to rapidly remove small volumes of leakage. However, thicker layers (8 to 18 inches) are usually
placed to facilitate construction and protect the liner layer from being damaged. The waste material itself may
serve this purpose if the material is granular, durable and relatively free draining.
C.2 Perforated Pipes Closely-spaced perforated pipes can be used to control hydraulic head above the liner.
The required spacing is calculated based on the maximum desired head and the fow rate and hydraulic
conductivity of the waste material between the pipes.
C.3 Geocomposite Systems Composite systems consisting of synthetic drainage associated with geotextile flters
have been developed for a wide range of drainage control functions. Performance of these systems under load
must be confrmed.
D. Leachate Collection and Removal Systems
D.1 Synthetic Geonet Materials Geonets are net-like polymer products designed to allow high rates of transverse
fow. Typical thicknesses of these materials vary from 0.16 to 0.30 inches.
D.2 Free-draining Gravel Layer (See Item C.1)
(HUTCHISON AND ELLISON, 1992)
< PREVIOUS VIEW
10-22
Chapter 10
MAY 2009
Geomembranes made from high-density polyethylene (HDPE), polyvinyl chloride (PVC) and very
low density polyethylene (VLDPE) have been used as liners. Important considerations in the selec-
tion of geomembranes are thickness, strength, durability, chemical resistivity, cost, cover mate-
rial needed for cushioning above and below the barrier, method of construction, and the method
for seaming the liner (Hutchison and Ellison, 1992). Most geomembranes are manufactured with
ultraviolet inhibitors (e.g., carbon black) and can be expected to last more than 50 years even when
exposed to sunlight. Geosynthetic clay liners (GCLs) consisting of bentonite sandwiched between
two geotextiles (woven or non-woven synthetic fabrics) that are glued or sewn together or ben-
tonite laminated to an HDPE geomembrane have also been used. GCLs are resistant to damage
due to handling during installation, but they lose shear strength as the bentonite is hydrated, thus
decreasing stability.
Geotextiles and soil materials above and below the geomembrane layer may be needed for protection
against penetrations by underlying rocks or sharp objects during construction. The protective soil
layers should be relatively free of large rocks and roots that could cause concentrated stresses in the
liner. State agencies can provide guidance on the use of geomembranes and may specify a minimum
thicknes and requirements for compatibility with the refuse materials.
An efective QA/QC program is essential for installation of soil liners, geomembranes and GCLs.
Past failures have been atributed to poor material placement, seaming, and protection (Daniel and
Koerner, 1995). Composite systems that consist of a combination of soil and a geomembrane have less
potential for quality control problems, but may only be economically feasible when suitable soils are
available on site.
Figure 10.1 shows three examples of soil and geomembrane liner designs used at coal refuse disposal
facilities. Compacted subgrade, as shown in Figure 10.1a, is acceptable in many situations for con-
tainment of coal refuse. Soil liners and synthetic liner systems, as shown in Figures 10.1b and 10.1c,
respectively, may be atractive in some situations. Some soil and synthetic liner systems may require
a prepared subgrade and protective cover materials.
Other layers can be added to a liner system if warranted, including a leachate collection layer
and a leachate detection layer. The leachate collection layer should be positioned above the liner
to collect and convey seepage from the refuse and to limit the buildup of hydraulic head on the
liner. The thickness and hydraulic conductivity of the leachate collection layer should be designed
based upon the potential leachate fow, liner confguration (slopes and other geometry that afect
seepage), and any restrictions on hydraulic head associated with the liner. The leachate collection
layer typically consists of sand and/or gravel designed to be more hydraulically conductive than
the waste itself. Geotextiles may be used between this layer and the liner for cushioning and to
improve stability.
A network of perforated pipes is sometimes provided within the leachate collection layer to increase
capacity, and these pipes must be properly designed to withstand crushing under the embankment
weight. The leachate collection layer typically drains to one or more central collector or header pipes.
Solid-wall pipes convey the leachate from the disposal area to holding areas for eventual treatment (if
required) and discharge. Manholes may be installed at bends and at regular intervals for pipe inspec-
tion and cleaning; cleanout ftings may also be used.
Geonets and geocomposite drainage products have been used in some applications for leachate col-
lection (DiGioia et al., 1995) if chemical compatibility and fow capacity under the applied load is
acceptable. These products are also sometimes used for leak detection zones beneath the primary
liner when conditions warrant.
< PREVIOUS VIEW
10-23
Environmental Considerations
MAY 2009
FIGURE 10.1 LINER SYSTEMS USED AT COAL REFUSE
DISPOSAL FACILITIES
10.1a COMPACTED SUBGRADE
10.1c SYNTHETIC LINER
GEOMEMBRANE OR
GEOSYNTHETIC CLAY LINER
COAL REFUSE
PROTECTIVE COVER MATERIAL
COMPACTED SUBGRADE
NATURAL SOILS OR ROCK
COMPACTED SUBGRADE
COAL REFUSE
NATURAL SOILS OR ROCK
10.1b SOIL LINER
PROTECTIVE COVER MATERIAL
COMPACTED SUBGRADE
NATURAL SOILS OR ROCK
COAL REFUSE
2-FOOT-THICK SOIL LAYER
COMPACTED TO 95 TO 100
PERCENT STANDARD PROCTOR
FIGURE 10.1 LINER SYSTEMS USED AT COAL REFUSE DISPOSAL FACILITIES
10.4.2 Stability
Structural stability of a liner system is a critical element in design. For embankments, the follow-
ing types of waste stability issues could cause liner damage: (1) sloughing of loose uncompacted
material from surfcial zones, (2) block failure of the waste material moving laterally with shearing
occurring predominantly within the liner system (particularly for sloping liner system), and (3)
dynamic slope instability and permanent displacement related to earthquake or other dynamic
loading. Liner damage in the vicinity of the impoundment can arise from erosion due to slurry
discharge or from natural runof, as well as instability of adjacent hillsides. A more detailed discus-
FIGURE 10.1 LINER SYSTEMS USED AT COAL REFUSE DISPOSAL FACILITIES
< PREVIOUS VIEW
10-24
Chapter 10
MAY 2009
TABLE 10.8 STABILITY CONSIDERATIONS IN LINER DESIGN
Problem Liner Stress Free Body Diagram
Required Properties
Typical
Factor of
Safety
Geomembrane Landfll
1. Liner self-weight Tensile
T
W
FL
T
W
FL
G, t,
allow
,
L
, H 10 to 100
2. Weight of flling Tensile
W
FL
FU T
t,
allow
,
U
,
L
, h, , H 0.5 to 10
3. Impact during
construction
Impact
I
I d, W 0.1 to 5
4. Weight of landfll Compression
n
allow
, H 10 to 50
5. Puncture Puncture
P
I
p
, H, P, A
p
0.5 to 10
6. Anchorage Tensile
T
FU
FL
t,
allow
,
U
,
L
, , 0.7 to 5
7. Settlement of
landfll
Shear
,
U
, , H 10 to 100
8. Subsidence
under landfll
Tensile
T z
n
FU
FL
t,
allow
,
U
,
L
, z , , H 0.3 to 10
Legend:
G = specifc gravity = slope angle
T = tensile force H = landfll height
t = thickness = unit weight
allow
= allowable strength h = lift height
= shear strength = subsidence angle
I = impact resistance = friction angle
p
= puncture strength d = drop height
U
= friction coeffcient with material above W = weight
L
= friction coeffcient with material below P = puncture force
F
U
= friction force upper A
p
= puncture area
F
L
= friction force lower z = mobilization distance
(ADAPTED FROM KOERNER, 2006)
sion of slope stability for situations involving liners is provided in Section 6.6.4. Koerner (2006) also
presents procedures for analysis of liner stability. Table 10.8 summarizes stability issues associated
with liner system design.
10.4.3 Performance Considerations
Liner system performance is measured in terms of the extent of control of leachate seepage from
the refuse disposal facility. Liner system performance is related to the types of waste material pres-
ent, the hydraulic head, and subsurface conditions, and these factors can mitigate or exacerbate the
potential hydrologic impacts. Table 10.9 provides a summary of guidance related to performance
< PREVIOUS VIEW
10-25
Environmental Considerations
MAY 2009
T
A
B
L
E
1
0
.
9
L
I
N
E
R
S
Y
S
T
E
M
P
E
R
F
O
R
M
A
N
C
E
F
A
C
T
O
R
S
A
N
D
C
O
N
S
I
D
E
R
A
T
I
O
N
S
L
i
n
e
r
T
y
p
e
T
h
i
c
k
n
e
s
s
(
f
t
)
H
y
d
r
a
u
l
i
c
C
o
n
d
u
c
t
i
v
i
t
y
(
c
m
/
s
e
c
)
E
x
a
m
p
l
e
H
y
d
r
a
u
l
i
c
H
e
a
d
E
s
t
i
m
a
t
e
d
M
i
n
.
S
t
e
a
d
y
-
S
t
a
t
e
L
e
a
k
a
g
e
R
a
t
e
Q
(
g
p
a
d
)
T
r
a
v
e
l
T
i
m
e
t
h
r
o
u
g
h
L
i
n
e
r
(
y
e
a
r
s
)
E
q
u
a
t
i
o
n
s
C
l
a
y
2
1
x
1
0
-
6
1
0
3
,
5
0
0
0
.
1
2
1
x
1
0
-
7
1
0
5
1
0
1
G
e
o
m
e
m
b
r
a
n
e
M
e
m
b
r
a
n
e
w
i
t
h
s
i
n
g
l
e
p
e
r
f
o
r
a
t
i
o
n
s
o
f
0
.
0
3
c
m
2
/
a
c
r
e
(
f
o
r
l
i
n
e
r
e
v
a
l
u
a
t
i
o
n
s
)
1
0
7
.
9
(
1
)
N
A
M
e
m
b
r
a
n
e
w
i
t
h
s
i
n
g
l
e
p
e
r
f
o
r
a
t
i
o
n
s
o
f
1
c
m
2
/
a
c
r
e
(
f
o
r
l
e
a
k
a
g
e
d
e
t
e
c
t
i
o
n
c
a
p
a
c
i
t
y
c
a
l
c
u
l
a
t
i
o
n
s
)
1
0
1
3
.
7
(
1
)
N
A
N
o
t
e
:
1
.
S
u
b
s
o
i
l
h
y
d
r
a
u
l
i
c
c
o
n
d
u
c
t
i
v
i
t
y
=
1
x
1
0
-
5
c
m
/
s
e
c
.
L
e
g
e
n
d
:
Q
=
s
e
e
p
a
g
e
f
o
w
i
n
g
a
l
l
o
n
s
p
e
r
a
c
r
e
p
e
r
d
a
y
(
g
p
a
d
)
D
t
=
t
h
i
c
k
n
e
s
s
o
f
s
a
t
u
r
a
t
e
d
w
a
s
t
e
(
l
e
n
g
t
h
)
q
=
s
e
e
p
a
g
e
p
e
r
u
n
i
t
a
r
e
a
(
l
e
n
g
t
h
/
t
i
m
e
)
h
f
=
s
o
i
l
p
o
r
e
-
w
a
t
e
r
p
r
e
s
s
u
r
e
b
e
l
o
w
l
i
n
e
r
(
l
e
n
g
t
h
)
k
l
=
h
y
d
r
a
u
l
i
c
c
o
n
d
u
c
t
i
v
i
t
y
o
f
l
i
n
e
r
(
l
e
n
g
t
h
/
t
i
m
e
)
n
=
e
f
f
e
c
t
i
v
e
p
o
r
o
s
i
t
y
(
d
i
m
e
n
s
i
o
n
l
e
s
s
)
D
l
=
t
h
i
c
k
n
e
s
s
o
f
l
i
n
e
r
(
l
e
n
g
t
h
)
i
=
i
n
i
t
i
a
l
i
n
-
s
i
t
u
v
o
l
u
m
e
t
r
i
c
m
o
i
s
t
u
r
e
c
o
n
t
e
n
t
(
d
i
m
e
n
s
i
o
n
l
e
s
s
)
k
t
=
h
y
d
r
a
u
l
i
c
c
o
n
d
u
c
t
i
v
i
t
y
o
f
w
a
s
t
e
(
l
e
n
g
t
h
/
t
i
m
e
)
h
c
=
s
o
i
l
p
o
r
e
-
w
a
t
e
r
p
r
e
s
s
u
r
e
(
l
e
n
g
t
h
)
d
=
d
i
a
m
e
t
e
r
o
f
h
o
l
e
i
n
g
e
o
m
e
m
b
r
a
n
e
(
l
e
n
g
t
h
)
k
s
=
h
y
d
r
a
u
l
i
c
c
o
n
d
u
c
t
i
v
i
t
y
o
f
s
u
b
s
o
i
l
(
l
e
n
g
t
h
/
t
i
m
e
)
q
r
=
r
e
d
u
c
e
d
s
p
e
c
i
f
c
l
e
a
k
a
g
e
(
l
e
n
g
t
h
)
H
=
d
e
p
t
h
o
f
p
o
n
d
e
d
l
i
q
u
i
d
(
h
y
d
r
a
u
l
i
c
h
e
a
d
)
a
b
o
v
e
t
o
p
o
f
l
i
n
e
r
(
l
e
n
g
t
h
)
T
=
t
r
a
v
e
l
t
i
m
e
(
A
D
A
P
T
E
D
F
R
O
M
H
U
T
C
H
I
S
O
N
A
N
D
E
L
L
I
S
O
N
,
1
9
9
2
)
q
=
k
s
q
r
2
4
d
=
4
q
r
e
x
p
[
(
H
h
f
)
]
2
q
r
T
=
(
n
i
)
[
D
l
(
H
-
h
c
)
l
n
(
2
D
l
+
H
h
f
)
]
k
l
D
l
+
H
h
f
q
=
k
l
[
D
t
+
D
l
h
f
]
D
l
+
D
t
k
l
k
t
< PREVIOUS VIEW
10-26
Chapter 10
MAY 2009
evaluation and provides some specifc examples. Measuring the performance of a liner system
typically involves monitoring of drain discharges and down gradient groundwater conditions.
10.5 RECLAMATION
Reclamation requirements vary according to land use, climate, and state regulations. The purpose of
this section is to provide general guidelines for reclamation grading, impoundment elimination, soil
and topsoil covering, and revegetation of coal refuse disposal facilities.
10.5.1 Design Considerations
The content of a reclamation plan is related to the planned post-mining land use, site terrain and
disposal facility confguration, climate, and pre-mining and adjacent area conditions. Generally, post-
mining land use is open space and wildlife habitat and may be oriented to specifc wildlife species
and vegetation biodiversity (e.g., forest and grass land mix). Other land use possibilities, although
rarely considered, include agriculture, recreation, and site development. All of these land uses typi-
cally require the establishment of persistent, low-maintenance vegetation for controlling erosion.
Site access and topography signifcantly afect future land use, and the engineering properties of the
embankment and the method of construction are important if structural foundations are planned.
Table 10.10 presents a summary of potential fnal land uses and related key requirements and con-
siderations.
In evaluating potential land uses, the availability of resources, and specifcally soils for revegetation,
must be determined. Other resources include water, access roads and existing site infrastructure.
Preparation of an inventory of resources is an integral step in the development of a reclamation plan.
Because soils are used for a variety of applications besides reclamation (e.g., starter dams, liners, etc.),
an understanding of the quantity and quality of soils available at or near the site is essential. During
the planning and design phases, geotechnical exploration should include feld characterization of soil
TABLE 10.10 POTENTIAL FINAL LAND USES
Land Use Examples Key Requirements and Considerations
Wildlife Habitat and Open Space
Adequate cover of appropriate vegetation for desired wildlife
species.
Agriculture
Pasture and Hay
Fiber Crops
Tree Nursery
Agricultural land uses should include assessment for trace
elements.
Recreation
Active Recreation (sports felds,
golf courses, ski/biking facilities)
Access to site, topography, erosion and drainage control.
Passive Recreation (hunting,
hiking, nature study)
Access to site, topography.
Site Development
Commercial and Industrial
(buildings, storage areas)
Access to site, topography, structural support, erosion and drainage
control. Residential (housing, parks)
Infrastructure (highways, airports)
(DIGIOIA ET AL., 1995)
< PREVIOUS VIEW
10-27
Environmental Considerations
MAY 2009
properties such as thickness, texture, and color. Evaluation of soil pH and potential lime requirements
during geotechnical laboratory testing of soils will enable improved planning of complex reclamation
sites. Analyses related to soil and other material handling should be performed with the capabilities
and limitations of the available excavation equipment in mind, so that costs associated with recovery
and segregation of soils during excavation, stockpiling, and redistribution are realistic.
Soils in the eastern U.S. and the Midwest tend to be neutral to acidic and, with addition of appropri-
ate amounts of lime and fertilizers, can support plant growth without irrigation if appropriate spe-
cies are selected. For practical purposes, lime will neutralize soils only to the depth of incorporation
(plow depth). If lime cannot be incorporated to a sufcient soil depth, plant species with tolerance to
low pH should be selected. Soils from arid or semi-arid regions in the west tend to be high in soluble
salts and/or sodium, and revegetation in this material can be challenging. All soils should be tested
for nutrient availability before lime and fertilizer application is specifed (Page et al., 1982). Table
10.11 presents design considerations for reclamation soils.
10.5.2 Grading
Final grading plans for reclamation should include development of surfaces and slopes in order to
achieve efective site drainage and to facilitate access for placement of soil and topsoil, vegetation, and
maintenance. While plans for a refuse disposal facility provide an anticipated fnal confguration and
slopes, the facility may not reach its planned capacity prior to reclamation. In such circumstances, a
reclamation grading plan providing site drainage (eliminating impounding conditions as necessary)
and minimizing erosion potential should be developed. The confguration of embankments, slopes,
benches and drainage channels at abandonment is subject to state regulatory criteria, which gener-
ally include requirements for overall embankment slopes, benches, and top surface grades.
TABLE 10.11 DESIGN CONSIDERATIONS FOR RECLAMATION SOILS
Support for Plants
Non-toxic for plants, capable of root penetration and storing suffcient
amounts of plant-available water
Soil Type and Thickness
Typical regulatory requirements are up to 4 feet in total thickness, with
equivalent topsoil placement to pre-mining condition. Alternate cover
and growth medium may be considered.
Root Anchorage
For large shrubs and trees, soil thickness of greater than 4 feet may be
required for anchorage against wind and gravity.
Water Storage
Field capacity and wilting point can be measured or estimated from
texture or grain-size distribution.
Establishment of Vegetation
Ability to add nutrients, pH adjustments, soil conditioners for
acceptable growth medium
pH
Most plant species grow best at a pH between 6.0 and 7.5. Soils
between pH of 3.5 and 6.0 can be limed; the cause of excessive pH
should be determined before adjustments are made.
Salt Stress
Sodium and salt soils can be evaluated using electrical conductivity
and sodium adsorption ratio.
Nutrient/Trace Element Availability
Nutrient and toxic element testing (Baker, 1988) can be used to identify
fertilizer and amendment requirements.
Species Selection
In addition to soil conditions, species should be selected based on
short- and long-term availability of irrigation water, short-term erosion
control requirements, and maintenance intensity and methods.
Yield of Vegetation for Land Use Adequate balance of nutrients and trace metals for sustained yield
Engineering Properties
Acceptable erosion resistance, hydraulic conductivity, load bearing
capacity, resistance to traffc, etc.
(ADAPTED FROM DIGIOIA ET AL., 1995)
< PREVIOUS VIEW
10-28
Chapter 10
MAY 2009
10.5.3 Impoundment Elimination
Elimination of slurry impoundments requires special measures for grading and reclamation. The
impoundment surface may be wet, dry, dessicated, or vegetated, but the underlying materials typi-
cally remain sof and can exhibit sudden shearing under equipment operation. The impoundment
elimination plan should address factors associated with: (1) fne refuse properties, (2) impoundment
size and depth, (3) the presence of water, and (4) the availability of and access to borrow sources for
regrading and covering materials. Preparation of an impoundment elimination plan may require
characterization of the fne refuse materials (including drainage and consolidation properties), speci-
fcation of the borrow material for covering the fne refuse, and specifcation of the equipment for
implementing the work. Section 11.5.2 provides guidance for upstream construction that should be
considered in developing and implementing plans for covering of an impoundment. The following
are typical guidelines for impoundment elimination:
Drainage toward the impoundment should be collected and routed away, and
ponded water should be removed.
Access into the impoundment area for delivery of borrow materials to cover the fne
coal refuse should be developed. This may involve construction of access roads and
designation of temporary stockpile areas in preparation for initial pushout of borrow
materials over the fne refuse.
An initial lif of borrow materials should be pushed out over the fne refuse using
a bulldozer. This initial lif should typically be between 4 and 6 feet thick, with the
lower end of this range more desirable for minimizing displacement of the fne
refuse. The pushout should not be performed into standing water, and dewater-
ing measures in the fne refuse (prolonged drying, drainage sumps, wick drains,
etc.) may be needed to facilitate placement of the initial lif. The initial lif should
be advanced from frm areas along the perimeter of the impoundment, creating a
wide area of operation rather than a narrow one. The initial lif should generally be
advanced a distance of at least 50 feet before additional lifs are placed or trucks or
haulage equipment are allowed onto covered areas. This distance should be main-
tained until the impoundment surface has been covered, and trucks should generally
not be allowed into this zone for delivery of borrow materials. Monitoring should be
conducted throughout the initial pushout period.
Subsequent lifs should be placed in accordance with geotechnical requirements for
the disposal embankment and should not exceed a thickness of 2 feet. Generally,
material for impoundment elimination is considered placed fll unless structural fll
is required by fnal land use. Depending on the geotechnical design requirements,
restrictions on the rate of fll placement may be warranted in order to limit loading
and to allow consolidation of the fne coal refuse.
Should displacement of fne refuse occur during pushout, the following measures
should be considered: (1) slowing the advance of the pushout to allow dissipation
of pore pressure in the fne refuse, (2) use of low-ground-pressure equipment, (3)
improving drainage within the fne refuse (e.g., sumps, wick drains), and (4) stabi-
lization of the fne refuse or reinforcement of the pushout lif using geotextiles or
geogrids. If displacement is unavoidable, the impound ment elimination plan should
include provisions for containment of the displaced material.
The material used to cover the fnes and eliminate the impounding capability
should be cambered such that when setlement occurs due to consolidation of the
underlying fnes, the surface will always provide positive drainage of the site. The
amount of long-term setlement should be estimated based on the consolidation
characteristics of the fnes.
< PREVIOUS VIEW
10-29
Environmental Considerations
MAY 2009
In the development of an impoundment elimination plan, the safety of equipment operators covering slurry
deposits should be addressed as well as monitoring of the work. The following general guidance applies:
Initial and periodic review sessions covering the procedures and anticipated perfor-
mance of the initial pushout, delivery of borrow material, and subsequent lif placement
should be held by engineering personnel for equipment operators and supervisors.
The impoundment should be maintained in a dewatered condition to the extent practical.
Initial pushouts should be restricted to daylight hours or times when the work area
is sufciently illuminated to provide good visibility.
Radio communication for pushout equipment operators and supervisors should be
provided, and it is recommended that equipment operations be within sight of mine
personnel during the initial pushout and that operators be provided with foatation
devices (e.g., life jackets).
The work area should be examined frequently for signs of instability such as crack-
ing or sinking, and work should be suspended in areas exhibiting such indications.
Monitoring of the work should be performed by engineering personnel with an under-
standing of technical issues such as slope stability, displacement, and deformation.
Pore pressures within the fne refuse and deformations or displacements may be
monitored with instrumentation, if warranted.
10.5.4 Soil and Topsoil Cover
Soil and topsoil cover materials with the properties that meet regulatory requirements for reclama-
tion should be stockpiled and recovered from locations near the disposal facility. While OSM and
state regulations typically require 4 feet of soil and topsoil cover, there are situations where a vari-
ance in cover thickness may be considered. Also, isolation of the refuse materials from infltrating
water may be necessary. In these circumstances, supplemental materials and/or modifed placement
procedures may be needed:
In Appalachian regions, there may be insufcient soil and topsoil for reclamation,
and reduced cover thicknesses may be necessary. Dove et al. (1987) and Daniels (2005)
evaluated direct seeding and reduced topsoil thickness alternatives to determine the
optimal combination of soil amendments and topsoil thicknesses for successful vegeta-
tion of refuse with varying levels of potential acidic leachate generation. Daniels (2005)
indicates that for moderately acid producing refuse, acceptable vegetation can be
established with less than 12 inches of soil cover if lime is added to the refuse surface.
Alternatives such as bio-solids and combustion ash may be considered. Bio-solids
(sewage sludge) can be plowed into refuse surfaces and used to establish an alterna-
tive growth medium. Combustion ash can also be applied or mixed into the refuse
surface to establish an alternative growth medium.
If isolation of the refuse materials requires a low-hydraulic-conductivity cap, use of
clay or a geo membrane may be appropriate. The evaluation and design of caps gen-
erally follows the procedures for liner systems presented in Section 10.4.
Achieving good adhesion of soil placed on refuse surfaces may require special procedures. The refuse
surface should be scarifed by tracking up and down slopes with a bulldozer, by shallow tillage
along contour lines, or by other methods that will loosen the surface. Soil should be placed in a rela-
tively dry condition with low-ground-pressure equipment, avoiding excessive compaction (unless
required for construction of a low hydraulic conductivity cap). If soils have a low pH (below 5) and
require amendment with lime or gypsum, it may be appropriate to place and amend the soil in lifs,
with incorporation (plowing in) of amendments through the entire lif prior to placement of the next
lif. Table 10.12 presents a summary of reclamation guidelines (DiGioia et al., 1995).
< PREVIOUS VIEW
10-30
Chapter 10
MAY 2009
TABLE 10.12 SUMMARY OF RECLAMATION GUIDELINES
Task Recommended Guidelines
Application of Lime Gypsum
and Fertilizer
Lime and gypsum should be plowed into the entire lift thickness, and fertilizers and
other plant nutrient sources should be applied evenly and plowed under within 24
hours and to a depth of at least 2 inches. If seedbed preparation includes creation of
furrows, seedbed preparation may be done in concert with fertilizer incorporation. If
hydroseeding is utilized, apply no more than 40-80-40 pounds N-P
2
O
5
-K
2
O per acre
with seed and do not leave seed and inoculants in contact with fertilizer-containing
solutions for more than 1 hour.
Furrowing and Land-
Imprinting
Where management of water or reduction in wind or salt stress is desired, deep
furrowing (6 to 10 inches), land imprinting, or other methods should performed.
Furrows should be oriented parallel to site contours or on fat surfaces, perpen dicular
to prevailing winds.
Seedbed Preparation
If the furrowing or land imprinting procedures are performed more than a few days
before seeding, or a crust has formed on the soil surface, these procedures should be
repeated just prior to seeding.
Seeding and Inoculating
Seeding depths using the drill seeding method should be set for the shallowest
seeded species. To maximize the opportunity for biological nitrogen fxation, legume
seed can be inoculated with Rhizobium strain specifc to the species being sown.
Broadcast and hydroseeding work best when the seeds are promptly covered with soil
and mulch.
Selection of Planting of
Woody Species
The emphasis is on establishment of herbaceous, not woody plants. Guidance on
selection, planting, maintenance and specifcation of woody plants is presented in
Vogel (1987) and Himelick (1981).
Mulching and Tacking
Mulch should be applied within a day of seeding and before rain. Straw and/or hay
applied at a rate of 3,000 to 6,000 pounds per acre and wood cellulose fber mulch
applied at a rate between 1,200 and 2,500 pounds per acre are accept able mulch for
most purposes. Tacking can be performed by crimping mulch into soil with large disks
set along the direction of travel or by application of wood cellulose fber mulch over
the straw/hay using a hydroseeder. Crimping tech niques that leave some straw/hay
standing up in the soil crease and in rows at right angle to the prevailing wind are
desirable for dry, windy sites.
Watercourse Protection
For watercourse protection (swales, ditches) wood excelsior, coconut fber, nylon, and/
or jute blankets should be used according to manufacturers instructions.
Irrigation
Irrigation should be considered during the establishment year in arid and semi-arid
regions and other areas where the gains from improvements in establish ment rate
and long-term survival outweigh the risk of failure.
(ADAPTED FROM DIGIOIA ET AL., 1995)
10.5.5 Vegetation
Species for vegetation and reclamation should be selected based on their adaptability and tolerance
to site climate and soil (or alternative media) conditions and their suitability for the fnal land use and
compatibility with regulatory provisions. Additional considerations include erosion and sedimenta-
tion control requirements, the need for irrigation water, and maintenance requirements. To the extent
possible, local expertise should be sought for development of vegetation plans for specifc land uses.
Where available, state erosion and sed imentation publications and university agronomy studies can
provide important guidance related to seeding/planting mixtures and cultivation practices. Potential
vegetation species and their adaptability to various climates and soil conditions are summarized in
Table 10.13 (DiGioia et al., 1995).
In humid regions, winter rye and redtop mixed with more slow-to-establish perennial species such
as birdsfoot trefoil and deertongue grass are ofen used to provide cover. Too high a seeding rate of
< PREVIOUS VIEW
10-31
Environmental Considerations
MAY 2009
T
A
B
L
E
1
0
.
1
3
C
H
A
R
A
C
T
E
R
I
S
T
I
C
S
O
F
C
O
M
M
O
N
S
P
E
C
I
E
S
P
O
T
E
N
T
I
A
L
L
Y
S
U
I
T
A
B
L
E
F
O
R
R
E
C
L
A
M
A
T
I
O
N
R
e
g
i
o
n
s
o
f
U
.
S
.
(
1
)
C
o
m
m
o
n
N
a
m
e
S
c
i
e
n
t
i
f
c
N
a
m
e
a
n
d
C
u
l
t
i
v
a
r
(
2
)
G
r
o
w
t
h
H
a
b
i
t
/
F
o
r
m
s
(
3
,
4
)
G
r
o
w
t
h
S
e
a
s
o
n
(
5
)
N
a
t
i
v
e
S
p
e
c
i
e
s
(
6
)
D
r
a
i
n
-
a
g
e
P
r
e
c
i
p
.
R
a
n
g
e
(
7
)
(
i
n
)
p
H
R
a
n
g
e
(
7
)
L
a
n
d
U
s
e
s
(
8
)
S
e
e
d
i
n
g
R
a
t
e
(
9
)
(
l
b
/
a
c
)
M
i
s
c
e
l
l
a
n
e
o
u
s
N
o
t
e
s
N
C
,
S
C
,
W
T
h
i
c
k
s
p
i
k
e
W
h
e
a
t
g
r
a
s
s
A
g
r
o
p
y
r
o
n
d
a
s
y
s
t
a
c
h
y
u
m
p
e
r
e
n
.
/
g
r
a
s
s
c
o
o
l
y
e
s
p
o
o
r
-
w
e
l
l
1
0
-
1
7
5
.
0
-
8
.
5
H
,
A
,
R
,
D
4
-
6
S
o
d
-
f
o
r
m
i
n
g
b
u
n
c
h
g
r
a
s
s
.
N
C
,
W
C
r
e
s
t
e
d
W
h
e
a
t
g
r
a
s
s
A
g
r
o
p
y
r
o
n
d
e
s
e
r
t
o
r
u
m
p
e
r
e
n
.
/
g
r
a
s
s
c
o
o
l
n
o
w
e
l
l
1
0
-
1
7
5
.
5
-
8
.
5
R
,
D
4
-
6
E
s
t
a
b
l
i
s
h
e
s
s
l
o
w
l
y
.
I
n
t
o
l
e
r
a
n
t
t
o
f
o
o
d
i
n
g
.
B
u
n
c
h
g
r
a
s
s
.
S
t
r
a
t
i
f
y
s
e
e
d
.
N
C
,
S
W
,
S
C
T
a
l
l
W
h
e
a
t
g
r
a
s
s
A
g
r
o
p
y
r
o
n
e
l
o
n
g
a
t
u
m
a
n
n
u
a
l
/
g
r
a
s
s
c
o
o
l
y
e
s
p
o
o
r
-
w
e
l
l
1
2
-
2
0
+
6
.
0
-
8
.
0
H
,
A
4
-
6
U
s
e
a
d
a
p
t
e
d
c
u
l
t
i
v
a
r
s
.
E
s
t
a
b
l
i
s
h
e
s
e
a
s
i
l
y
.
S
o
d
-
f
o
r
m
i
n
g
b
u
n
c
h
g
r
a
s
s
.
B
o
r
o
n
i
n
t
o
l
e
r
a
n
t
.
N
C
,
W
S
t
r
e
a
m
b
a
n
k
W
h
e
a
t
g
r
a
s
s
A
g
r
o
p
y
r
o
n
r
i
p
a
r
i
u
m
a
n
n
u
a
l
/
g
r
a
s
s
c
o
o
l
y
e
s
p
o
o
r
-
w
e
l
l
1
1
-
1
9
6
.
5
-
8
.
5
R
,
D
2
-
4
S
p
r
e
a
d
s
r
a
p
i
d
l
y
.
S
o
d
-
f
o
r
m
i
n
g
.
N
C
,
S
C
,
M
,
W
,
N
E
W
e
s
t
e
r
n
W
h
e
a
t
g
r
a
s
s
A
g
r
o
p
y
r
o
n
s
m
i
t
h
i
i
p
e
r
e
n
.
/
g
r
a
s
s
c
o
o
l
y
e
s
p
o
o
r
-
w
e
l
l
5
-
2
0
+
4
.
5
-
9
.
0
H
,
A
,
R
,
D
4
-
6
U
s
e
a
d
a
p
t
e
r
c
u
l
t
i
v
a
r
s
.
S
o
d
-
f
o
r
m
i
n
g
.
N
C
,
S
C
,
W
S
l
e
n
d
e
r
W
h
e
a
t
g
r
a
s
s
A
g
r
o
p
y
r
o
n
t
r
a
c
h
y
c
a
u
l
u
m
p
e
r
e
n
.
/
g
r
a
s
s
c
o
o
l
y
e
s
m
e
d
.
-
p
o
o
r
1
5
-
2
0
+
6
.
5
-
8
.
5
H
,
A
2
-
4
G
r
o
w
s
f
a
s
t
,
s
h
o
r
t
-
l
i
v
e
d
.
S
C
,
M
,
N
E
,
S
R
e
d
t
o
p
A
g
r
o
s
t
i
s
a
l
b
a
p
e
r
e
n
.
/
g
r
a
s
s
c
o
o
l
n
o
m
e
d
.
-
p
o
o
r
2
0
+
4
.
0
-
7
.
5
R
,
D
2
-
4
S
o
d
-
f
o
r
m
i
n
g
s
h
o
r
t
-
l
i
v
e
d
.
N
C
,
S
C
,
M
B
i
g
B
l
u
e
s
t
e
m
A
n
d
r
o
p
o
g
o
n
g
e
r
a
r
d
i
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
y
e
s
m
e
d
.
-
p
o
o
r
1
6
-
2
0
+
5
.
5
-
7
.
5
H
,
A
,
R
,
D
4
-
8
U
s
e
a
d
a
p
t
e
d
c
u
l
t
i
v
a
r
s
.
D
e
e
p
r
o
o
t
s
.
A
l
l
C
o
m
m
o
n
O
a
t
s
A
v
e
n
a
s
a
t
i
v
a
a
n
n
u
a
l
/
g
r
a
s
s
c
o
o
l
n
o
w
e
l
l
1
0
-
5
5
.
5
-
7
.
0
H
,
A
,
R
,
D
3
0
-
5
0
U
s
e
a
d
a
p
t
e
d
c
u
l
t
i
v
a
r
s
.
C
o
v
e
r
c
r
o
p
.
C
o
l
d
t
o
l
e
r
a
n
t
.
N
e
e
d
s
N
.
S
K
i
n
g
R
a
n
c
h
B
l
u
e
s
t
e
m
B
o
t
h
r
i
o
c
h
l
o
a
i
s
c
h
a
e
m
u
m
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
n
o
w
e
l
l
2
0
+
5
.
0
-
7
.
0
A
2
-
6
M
i
x
w
i
t
h
S
e
r
c
i
c
a
L
e
s
p
e
d
e
z
a
.
N
C
,
S
C
,
M
S
i
d
e
o
a
t
s
G
r
a
m
a
B
o
u
t
e
l
o
u
a
c
u
r
t
i
p
e
n
d
u
l
a
V
a
u
g
h
n
o
r
E
l
r
e
n
o
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
y
e
s
m
e
d
.
-
w
e
l
l
1
2
-
2
0
+
6
.
0
-
7
.
5
H
,
A
2
-
4
C
a
l
c
a
r
e
o
u
s
s
o
i
l
s
.
B
u
n
c
h
g
r
a
s
s
.
N
C
,
S
C
,
W
B
l
u
e
G
r
a
m
a
B
o
u
t
e
l
o
u
a
g
r
a
c
i
l
i
s
L
o
v
i
n
g
t
o
n
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
y
e
s
w
e
l
l
8
-
2
0
6
.
0
-
8
.
5
H
,
A
,
R
,
D
1
0
-
1
2
U
s
e
a
d
a
p
t
e
d
c
u
l
t
i
v
a
r
s
.
S
o
d
-
f
o
r
m
i
n
g
b
u
n
c
h
g
r
a
s
s
.
A
n
i
m
a
l
f
o
r
a
g
e
.
< PREVIOUS VIEW
10-32
Chapter 10
MAY 2009
T
A
B
L
E
1
0
.
1
3
C
H
A
R
A
C
T
E
R
I
S
T
I
C
S
O
F
C
O
M
M
O
N
S
P
E
C
I
E
S
P
O
T
E
N
T
I
A
L
L
Y
S
U
I
T
A
B
L
E
F
O
R
R
E
C
L
A
M
A
T
I
O
N
(
C
o
n
t
i
n
u
e
d
)
R
e
g
i
o
n
s
o
f
U
.
S
.
(
1
)
C
o
m
m
o
n
N
a
m
e
S
c
i
e
n
t
i
f
c
N
a
m
e
a
n
d
C
u
l
t
i
v
a
r
(
2
)
G
r
o
w
t
h
H
a
b
i
t
/
F
o
r
m
s
(
3
,
4
)
G
r
o
w
t
h
S
e
a
s
o
n
(
5
)
N
a
t
i
v
e
S
p
e
c
i
e
s
(
6
)
D
r
a
i
n
-
a
g
e
P
r
e
c
i
p
.
R
a
n
g
e
(
7
)
(
i
n
)
p
H
R
a
n
g
e
(
7
)
L
a
n
d
U
s
e
s
(
8
)
S
e
e
d
i
n
g
R
a
t
e
(
9
)
(
l
b
/
a
c
)
M
i
s
c
e
l
l
a
n
e
o
u
s
N
o
t
e
s
A
l
l
S
m
o
o
t
h
B
r
o
m
e
B
r
o
m
u
s
i
n
e
r
m
u
s
p
e
r
e
n
.
/
g
r
a
s
s
c
o
o
l
n
o
m
e
d
.
-
w
e
l
l
1
7
-
2
0
+
5
.
5
-
8
.
0
R
,
D
1
0
-
1
5
U
s
e
a
d
a
p
t
e
d
c
u
l
t
i
v
a
r
s
.
E
s
t
a
b
l
i
s
h
e
s
e
a
s
i
l
y
f
o
r
q
u
i
c
k
c
o
v
e
r
.
S
o
d
-
f
o
r
m
i
n
g
.
N
C
,
S
C
B
u
f
f
a
l
o
G
r
a
s
s
B
u
c
h
l
o
e
d
a
c
t
y
l
o
i
d
e
s
S
h
a
r
p
s
I
m
p
r
o
v
e
d
o
r
T
e
x
o
k
a
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
y
e
s
m
e
d
.
-
p
o
o
r
1
2
-
2
0
+
6
.
5
-
8
.
0
H
,
A
,
R
,
D
5
-
1
0
S
o
d
-
f
o
r
m
i
n
g
,
s
t
o
l
o
n
-
i
f
e
r
o
u
s
.
S
t
r
a
t
i
f
y
s
e
e
d
.
M
o
n
o
e
c
i
o
u
s
.
N
C
,
S
C
,
W
P
r
a
i
r
i
e
S
a
n
d
r
e
e
d
C
a
l
a
m
o
v
i
l
f
a
l
o
n
g
i
f
o
l
i
a
G
o
s
h
e
n
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
y
e
s
w
e
l
l
1
0
-
2
0
6
.
0
-
8
.
0
H
,
A
,
R
,
D
6
-
9
S
o
d
-
f
o
r
m
i
n
g
,
r
h
i
z
o
m
a
t
o
u
s
.
S
R
h
o
d
e
s
G
r
a
s
s
C
h
l
o
r
i
s
g
a
y
a
n
a
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
n
o
m
e
d
.
2
0
+
4
.
0
-
7
.
0
H
,
A
,
R
,
D
8
-
1
2
V
e
r
y
t
o
l
e
r
a
n
t
t
o
s
a
l
i
n
i
t
y
.
S
t
o
l
o
n
i
f
e
r
o
u
s
.
N
C
,
S
,
S
C
,
N
E
B
e
r
m
u
d
a
G
r
a
s
s
C
y
n
o
d
o
n
d
a
c
t
y
l
o
n
Q
u
i
c
k
s
a
n
d
C
o
m
m
o
n
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
n
o
m
e
d
.
-
p
o
o
r
1
6
-
2
5
+
4
.
5
-
7
.
5
A
,
R
,
D
3
-
5
F
o
u
r
-
y
e
a
r
c
o
v
e
r
.
R
h
i
z
o
m
a
t
o
u
s
a
n
d
s
e
e
d
s
p
r
e
a
d
i
n
g
.
N
C
,
M
,
N
E
,
S
O
r
c
h
a
r
d
G
r
a
s
s
D
a
c
t
y
l
u
s
g
l
o
m
e
r
a
t
a
p
e
r
e
n
.
/
g
r
a
s
s
c
o
o
l
n
o
m
e
d
.
-
w
e
l
l
2
0
+
5
.
0
-
7
.
5
H
,
A
,
R
,
D
5
-
8
U
s
e
a
d
a
p
t
e
d
c
u
l
t
i
v
a
r
s
.
B
u
n
c
h
g
r
a
s
s
.
M
,
N
E
,
S
D
e
e
r
t
o
n
g
u
e
G
r
a
s
s
D
i
c
h
a
n
t
h
e
l
i
u
m
c
l
a
n
d
e
s
t
i
n
u
m
T
i
o
g
a
(
P
a
n
i
c
u
m
c
.
)
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
y
e
s
p
o
o
r
-
w
e
l
l
2
0
+
3
.
8
-
8
.
0
H
,
A
,
R
6
-
2
0
S
t
r
a
t
i
f
y
s
e
e
d
.
W
e
t
a
n
d
d
r
y
s
i
t
e
s
.
E
s
t
a
b
l
i
s
h
e
s
s
l
o
w
l
y
.
N
C
,
S
C
,
W
S
a
l
t
g
r
a
s
s
D
i
s
t
i
c
h
l
i
s
s
t
r
i
c
t
a
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
y
e
s
m
e
d
.
-
p
o
o
r
1
4
-
2
0
+
6
.
0
-
8
.
5
H
,
R
,
D
4
-
9
S
a
l
t
t
o
l
e
r
a
n
t
.
N
C
,
W
B
a
s
i
n
W
i
l
d
r
y
e
E
l
y
m
u
s
c
i
n
e
r
i
u
s
p
e
r
e
n
.
/
g
r
a
s
s
c
o
o
l
y
e
s
p
o
o
r
1
4
-
2
0
+
6
.
0
-
8
.
0
H
,
A
,
R
4
-
8
L
o
n
g
-
l
i
v
e
d
b
u
n
c
h
g
r
a
s
s
.
N
C
,
W
R
u
s
s
i
a
n
W
i
l
d
r
y
e
E
l
y
m
u
s
j
u
n
c
e
u
s
p
e
r
e
n
.
/
g
r
a
s
s
c
o
o
l
n
o
m
e
d
.
-
w
e
l
l
1
3
-
2
0
6
.
0
-
8
.
0
R
,
D
4
-
8
D
r
o
u
g
h
t
t
o
l
e
r
a
n
t
.
N
C
,
W
B
e
a
r
d
l
e
s
s
W
i
l
d
r
y
e
E
l
y
m
u
s
t
r
i
t
i
c
o
i
d
e
s
p
e
r
e
n
.
/
g
r
a
s
s
c
o
o
l
y
e
s
p
o
o
r
1
6
-
2
0
+
6
.
0
-
8
.
0
H
,
A
,
R
,
D
4
-
8
E
s
t
a
b
l
i
s
h
e
s
s
l
o
w
l
y
.
W
B
o
e
r
L
o
v
e
g
r
a
s
s
E
r
a
g
r
o
s
t
i
s
c
h
l
o
r
o
m
e
l
a
s
A
-
8
4
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
n
o
m
e
d
.
-
w
e
l
l
1
2
-
1
9
6
.
0
-
8
.
0
R
,
D
4
-
8
H
o
t
,
d
r
y
c
l
i
m
a
t
e
s
.
B
u
n
c
h
g
r
a
s
s
< PREVIOUS VIEW
10-33
Environmental Considerations
MAY 2009
T
A
B
L
E
1
0
.
1
3
C
H
A
R
A
C
T
E
R
I
S
T
I
C
S
O
F
C
O
M
M
O
N
S
P
E
C
I
E
S
P
O
T
E
N
T
I
A
L
L
Y
S
U
I
T
A
B
L
E
F
O
R
R
E
C
L
A
M
A
T
I
O
N
(
C
o
n
t
i
n
u
e
d
)
R
e
g
i
o
n
s
o
f
U
.
S
.
(
1
)
C
o
m
m
o
n
N
a
m
e
S
c
i
e
n
t
i
f
c
N
a
m
e
a
n
d
C
u
l
t
i
v
a
r
(
2
)
G
r
o
w
t
h
H
a
b
i
t
/
F
o
r
m
s
(
3
,
4
)
G
r
o
w
t
h
S
e
a
s
o
n
(
5
)
N
a
t
i
v
e
S
p
e
c
i
e
s
(
6
)
D
r
a
i
n
-
a
g
e
P
r
e
c
i
p
.
R
a
n
g
e
(
7
)
(
i
n
)
p
H
R
a
n
g
e
(
7
)
L
a
n
d
U
s
e
s
(
8
)
S
e
e
d
i
n
g
R
a
t
e
(
9
)
(
l
b
/
a
c
)
M
i
s
c
e
l
l
a
n
e
o
u
s
N
o
t
e
s
S
C
,
M
,
N
E
,
S
W
e
e
p
i
n
g
L
o
v
e
g
r
a
s
s
E
r
a
g
r
o
s
t
i
s
c
u
r
v
u
l
a
E
r
m
e
l
o
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
n
o
w
e
l
l
2
0
+
4
.
5
-
8
.
0
R
,
D
1
-
3
B
u
n
c
h
g
r
a
s
s
,
s
h
o
r
t
-
l
i
v
e
d
i
n
t
h
e
N
E
.
A
l
l
T
a
l
l
F
e
s
c
u
e
F
e
s
t
u
c
a
a
r
u
n
d
i
n
a
c
e
a
K
e
n
t
u
c
k
y
-
3
1
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
y
e
s
m
e
d
.
-
p
o
o
r
1
8
-
2
0
+
5
.
0
-
8
.
0
R
,
D
1
0
-
1
2
U
s
e
a
d
a
p
t
e
d
c
u
l
t
i
v
a
r
s
B
u
n
c
h
g
r
a
s
s
U
s
e
o
n
l
y
e
n
d
o
p
h
y
t
e
-
f
r
e
e
v
a
r
i
e
t
i
e
s
f
o
r
a
g
r
i
c
u
l
t
u
r
a
l
l
a
n
d
u
s
e
s
.
S
C
,
M
,
N
E
,
S
R
e
d
F
e
s
c
u
e
F
e
s
t
u
c
a
r
u
b
r
a
p
e
r
e
n
.
/
g
r
a
s
s
c
o
o
l
n
o
p
o
o
r
-
w
e
l
l
2
0
+
5
.
0
-
7
.
5
A
,
R
,
D
4
-
6
U
s
e
a
d
a
p
t
e
d
c
u
l
t
i
v
a
r
s
.
E
s
t
a
b
l
i
s
h
e
s
s
l
o
w
l
y
.
S
o
d
-
f
o
r
m
i
n
g
.
W
G
a
l
l
e
t
a
H
i
l
a
r
i
a
j
a
m
e
s
i
i
V
i
v
a
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
y
e
s
m
e
d
.
-
w
e
l
l
6
-
1
2
6
.
0
-
8
.
0
H
,
A
,
R
,
D
6
-
1
0
V
e
r
y
d
r
o
u
g
h
t
t
o
l
e
r
a
n
t
,
r
h
i
z
o
m
a
t
o
u
s
.
W
B
i
g
G
a
l
l
e
t
a
H
i
l
a
r
i
a
r
i
g
i
d
a
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
y
e
s
m
e
d
.
9
-
1
4
6
.
0
-
8
.
0
H
,
A
,
R
,
D
6
-
1
0
V
e
r
y
d
r
o
u
g
h
t
t
o
l
e
r
a
n
t
,
r
h
i
z
o
m
a
t
o
u
s
.
N
C
,
S
C
M
,
N
E
,
S
B
a
r
l
e
y
H
o
r
d
e
u
m
v
u
l
g
a
r
e
a
n
n
u
a
l
/
g
r
a
s
s
c
o
o
l
n
o
m
e
d
.
-
w
e
l
l
2
0
+
6
.
0
-
7
.
8
H
,
A
,
R
,
D
1
0
-
2
5
W
i
n
t
e
r
c
o
v
e
r
c
r
o
p
.
N
C
,
S
C
,
M
A
n
n
u
a
l
(
I
t
a
l
i
a
n
)
R
y
e
g
r
a
s
s
L
o
l
i
u
m
m
u
l
t
i
f
o
r
u
m
a
n
n
u
a
l
/
g
r
a
s
s
c
o
o
l
n
o
m
e
d
.
-
w
e
l
l
2
0
+
5
.
5
-
7
.
5
A
,
R
,
D
4
-
6
Q
u
i
c
k
c
o
v
e
r
c
r
o
p
.
S
C
K
l
e
i
n
G
r
a
s
s
P
a
n
i
c
u
m
c
o
l
o
r
a
t
u
m
S
e
l
e
c
t
i
o
n
7
5
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
y
e
s
m
e
d
.
1
8
-
2
0
+
6
.
0
-
8
.
0
A
,
R
,
D
6
-
1
0
T
e
x
a
s
.
N
C
,
S
C
,
M
,
N
E
,
S
S
w
i
t
c
h
g
r
a
s
s
P
a
n
i
c
u
m
v
i
r
g
a
t
u
m
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
y
e
s
m
e
d
.
-
p
o
o
r
1
8
-
2
0
+
5
.
0
-
7
.
5
H
,
A
,
R
2
-
5
U
s
e
a
d
a
p
t
e
d
c
u
l
t
i
v
a
r
s
a
n
d
e
c
o
t
y
p
e
s
.
B
l
a
c
k
w
e
l
l
v
a
r
i
e
t
y
i
s
c
o
m
p
e
t
i
t
i
v
e
.
R
h
i
z
o
m
a
t
o
u
s
.
G
o
o
d
w
i
n
t
e
r
c
o
v
e
r
.
S
B
a
h
i
a
G
r
a
s
s
P
a
s
p
a
l
u
m
n
o
t
a
t
u
m
P
e
n
s
a
c
o
l
a
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
n
o
m
e
d
.
-
w
e
l
l
2
0
+
4
.
5
-
7
.
5
H
,
A
2
5
-
3
5
S
t
r
a
t
i
f
y
s
e
e
d
s
.
S
o
d
-
f
o
r
m
i
n
g
s
e
m
i
-
b
u
n
c
h
g
r
a
s
s
.
N
C
,
N
E
,
M
R
e
e
d
C
a
n
a
r
y
G
r
a
s
s
P
h
a
l
a
r
i
s
a
r
u
n
d
i
n
a
c
e
a
p
e
r
e
n
.
/
g
r
a
s
s
c
o
o
l
n
o
p
o
o
r
1
8
-
2
0
+
5
.
0
-
7
.
5
H
,
A
5
-
8
W
e
t
s
i
t
e
s
.
< PREVIOUS VIEW
10-34
Chapter 10
MAY 2009
T
A
B
L
E
1
0
.
1
3
C
H
A
R
A
C
T
E
R
I
S
T
I
C
S
O
F
C
O
M
M
O
N
S
P
E
C
I
E
S
P
O
T
E
N
T
I
A
L
L
Y
S
U
I
T
A
B
L
E
F
O
R
R
E
C
L
A
M
A
T
I
O
N
(
C
o
n
t
i
n
u
e
d
)
R
e
g
i
o
n
s
o
f
U
.
S
.
(
1
)
C
o
m
m
o
n
N
a
m
e
S
c
i
e
n
t
i
f
c
N
a
m
e
a
n
d
C
u
l
t
i
v
a
r
(
2
)
G
r
o
w
t
h
H
a
b
i
t
/
F
o
r
m
s
(
3
,
4
)
G
r
o
w
t
h
S
e
a
s
o
n
(
5
)
N
a
t
i
v
e
S
p
e
c
i
e
s
(
6
)
D
r
a
i
n
-
a
g
e
P
r
e
c
i
p
.
R
a
n
g
e
(
7
)
(
i
n
)
p
H
R
a
n
g
e
(
7
)
L
a
n
d
U
s
e
s
(
8
)
S
e
e
d
i
n
g
R
a
t
e
(
9
)
(
l
b
/
a
c
)
M
i
s
c
e
l
l
a
n
e
o
u
s
N
o
t
e
s
M
,
N
E
,
S
T
i
m
o
t
h
y
G
r
a
s
s
P
h
l
e
u
m
p
r
e
t
e
n
s
e
p
e
r
e
n
.
/
g
r
a
s
s
c
o
o
l
n
o
p
o
o
r
-
w
e
l
l
2
0
+
4
.
5
-
8
.
0
H
,
A
,
R
4
-
7
U
s
e
a
d
a
p
t
e
d
c
u
l
t
i
v
a
r
s
R
h
i
z
o
m
a
t
o
u
s
.
S
h
o
r
t
-
l
i
v
e
d
.
N
C
,
S
C
,
M
B
i
g
B
l
u
e
g
r
a
s
s
P
o
a
a
m
p
l
a
p
e
r
e
n
.
/
g
r
a
s
s
c
o
o
l
y
e
s
w
e
l
l
1
2
-
2
0
6
.
0
-
8
.
0
H
,
A
6
-
1
0
B
u
n
c
h
g
r
a
s
s
.
P
h
e
a
s
a
n
t
n
e
s
t
i
n
g
h
a
b
i
t
a
t
.
A
L
L
L
i
t
t
l
e
B
l
u
e
s
t
e
m
S
c
h
i
z
a
c
h
y
r
i
u
m
s
c
o
p
a
r
i
u
m
(
A
n
d
r
o
p
o
g
o
n
s
c
o
p
a
r
i
u
s
)
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
y
e
s
w
e
l
l
1
2
-
2
0
+
6
.
0
-
8
.
0
H
,
A
,
R
,
D
4
-
8
U
s
e
a
d
a
p
t
e
d
c
u
l
t
i
v
a
r
s
.
N
C
,
S
C
,
M
,
S
,
N
E
R
y
e
S
e
c
a
l
e
c
e
r
e
a
l
e
B
a
l
b
o
a
n
n
u
a
l
/
g
r
a
s
s
c
o
o
l
n
o
w
e
l
l
2
0
+
5
.
5
-
7
.
5
H
,
A
,
R
,
D
3
0
-
6
0
C
o
v
e
r
c
r
o
p
.
N
C
,
S
C
,
M
,
S
,
N
E
I
n
d
i
a
n
G
r
a
s
s
S
o
r
g
h
a
s
t
r
u
m
n
u
t
a
n
s
p
e
r
e
n
.
/
g
r
a
s
s
w
a
r
m
y
e
s
m
e
d
.
-
p
o
o
r
1
5
-
3
5
4
.
0
-
7
.
5
R
,
D
2
-
4
S
o
d
-
f
o
r
m
i
n
g
,
s
h
o
r
t
-
l
i
v
e
d
.
S
C
,
S
J
o
h
n
s
o
n
G
r
a
s
s
S
o
r
g
h
u
m
S
u
d
a
n
e
s
e
a
n
n
u
a
l
/
g
r
a
s
s
w
a
r
m
n
o
m
e
d
.
-
w
e
l
l
2
0
+
5
.
5
-
7
.
5
A
,
R
,
D
1
5
-
2
0
U
s
e
a
d
a
p
t
e
d
c
u
l
t
i
v
a
r
s
.
C
o
v
e
r
c
r
o
p
.
N
C
,
S
C
,
W
A
l
k
a
l
i
S
a
c
a
t
o
n
S
p
o
r
o
b
o
l
u
s
a
i
r
o
i
d
e
s
p
e
r
e
n
.
/
g
r
a
s
s
c
o
o
l
y
e
s
p
o
o
r
-
w
e
l
l
6
-
1
8
7
.
0
-
8
.
5
H
,
A
,
R
,
D
6
-
1
0
U
s
e
l
o
c
a
l
e
c
o
t
y
p
e
s
.
B
u
n
c
h
g
r
a
s
s
.
N
e
e
d
s
i
n
i
t
i
a
l
I
r
r
i
g
a
t
i
o
n
.
S
C
,
W
,
M
D
r
o
p
s
e
e
d
S
p
o
r
o
b
o
l
u
s
s
p
p
.
p
e
r
e
n
.
/
g
r
a
s
s
c
o
o
l
y
e
s
m
e
d
.
-
w
e
l
l
1
0
-
2
0
+
6
.
0
-
8
.
5
H
,
A
,
R
,
D
4
-
8
B
u
n
c
h
g
r
a
s
s
.
S
e
e
d
s
p
r
o
l
i
f
c
.
N
C
,
S
C
C
i
c
e
r
M
i
l
k
v
e
t
c
h
A
s
t
r
a
g
a
l
u
s
c
i
c
e
r
L
u
t
a
n
a
b
i
e
n
.
/
l
e
g
u
m
e
c
o
o
l
n
o
p
o
o
r
-
w
e
l
l
1
8
-
3
5
5
.
5
-
7
.
5
R
,
D
2
-
5
D
r
o
u
g
h
t
t
o
l
e
r
a
n
t
,
s
o
d
-
f
o
r
m
i
n
g
.
E
s
t
a
b
l
i
s
h
e
s
s
l
o
w
l
y
.
S
c
a
r
i
f
y
S
e
e
d
.
N
C
,
M
,
N
E
,
S
C
r
o
w
n
V
e
t
c
h
C
o
r
o
n
i
l
l
a
v
a
r
i
a
p
e
r
e
n
.
/
l
e
g
u
m
e
c
o
o
l
n
o
w
e
l
l
2
3
+
5
.
0
-
7
.
5
R
,
D
5
-
8
E
s
t
a
b
l
i
s
h
e
s
s
l
o
w
l
y
.
E
x
c
l
u
d
e
s
t
r
e
e
s
e
e
d
i
n
g
s
.
E
m
e
r
a
l
d
v
a
r
i
e
t
y
f
o
r
d
r
y
s
i
t
e
s
.
B
o
r
o
n
t
o
l
e
r
a
n
t
.
M
,
S
C
,
S
I
l
l
i
n
o
i
s
B
u
n
d
l
e
f
o
w
e
r
D
e
s
m
a
n
t
h
u
s
i
l
l
i
n
o
e
n
s
i
s
p
e
r
e
n
.
/
l
e
g
u
m
e
w
a
r
m
y
e
s
m
e
d
.
-
w
e
l
l
1
5
-
2
0
+
5
.
0
-
7
.
5
H
,
A
,
D
2
-
5
D
r
o
u
g
h
t
r
e
s
i
s
t
a
n
t
.
N
C
,
S
C
,
M
,
N
E
,
S
S
o
y
b
e
a
n
G
l
y
c
i
n
e
m
a
x
a
n
n
u
a
l
/
l
e
g
u
m
e
w
a
r
m
n
o
p
o
o
r
-
w
e
l
l
2
5
+
5
.
0
-
7
.
0
H
,
A
3
0
-
5
0
C
a
s
h
c
r
o
p
.
F
l
o
o
d
i
n
g
i
n
t
o
l
e
r
a
n
t
.
< PREVIOUS VIEW
10-35
Environmental Considerations
MAY 2009
T
A
B
L
E
1
0
.
1
3
C
H
A
R
A
C
T
E
R
I
S
T
I
C
S
O
F
C
O
M
M
O
N
S
P
E
C
I
E
S
P
O
T
E
N
T
I
A
L
L
Y
S
U
I
T
A
B
L
E
F
O
R
R
E
C
L
A
M
A
T
I
O
N
(
C
o
n
t
i
n
u
e
d
)
R
e
g
i
o
n
s
o
f
U
.
S
.
(
1
)
C
o
m
m
o
n
N
a
m
e
S
c
i
e
n
t
i
f
c
N
a
m
e
a
n
d
C
u
l
t
i
v
a
r
(
2
)
G
r
o
w
t
h
H
a
b
i
t
/
F
o
r
m
s
(
3
,
4
)
G
r
o
w
t
h
S
e
a
s
o
n
(
5
)
N
a
t
i
v
e
S
p
e
c
i
e
s
(
6
)
D
r
a
i
n
-
a
g
e
P
r
e
c
i
p
.
R
a
n
g
e
(
7
)
(
i
n
)
p
H
R
a
n
g
e
(
7
)
L
a
n
d
U
s
e
s
(
8
)
S
e
e
d
i
n
g
R
a
t
e
(
9
)
(
l
b
/
a
c
)
M
i
s
c
e
l
l
a
n
e
o
u
s
N
o
t
e
s
N
C
,
S
C
,
M
,
N
E
,
S
L
a
t
h
c
o
F
l
a
t
p
e
a
L
a
t
h
r
u
s
s
y
l
v
e
s
t
r
i
s
L
a
t
h
c
o
p
e
r
e
n
.
/
l
e
g
u
m
e
c
o
o
l
n
o
m
e
d
.
-
w
e
l
l
2
3
+
4
.
5
-
7
.
0
R
,
D
2
0
S
e
e
d
s
,
t
o
x
i
c
,
e
x
c
l
u
d
e
s
t
r
e
e
s
e
e
d
l
i
n
g
s
.
E
s
t
a
b
l
i
s
h
e
s
s
l
o
w
l
y
.
S
C
,
M
,
N
E
,
S
S
e
r
i
c
e
a
L
e
s
p
e
d
e
z
a
L
e
s
p
e
d
e
z
a
c
u
n
e
a
t
a
p
e
r
e
n
.
/
l
e
g
u
m
e
w
a
r
m
n
o
w
e
l
l
2
0
+
4
.
5
-
7
.
0
A
,
R
,
D
1
0
-
2
0
U
s
e
a
d
a
p
t
e
d
c
u
l
t
i
v
a
r
s
.
S
c
a
r
i
f
y
s
e
e
d
s
.
W
o
o
d
y
s
t
e
m
s
.
S
C
,
M
,
N
E
,
S
P
r
o
s
t
r
a
t
e
L
e
s
p
e
d
e
z
a
L
e
s
p
e
d
e
z
a
d
a
u
r
i
c
a
v
a
r
.
s
c
h
i
m
a
d
a
i
p
e
r
e
n
.
/
l
e
g
u
m
e
w
a
r
m
n
o
w
e
l
l
2
0
+
4
.
5
-
7
.
0
A
,
R
,
D
1
5
-
2
0
P
r
o
s
t
r
a
t
e
f
o
r
m
u
s
e
d
w
i
t
h
t
r
e
e
s
e
e
d
l
i
n
g
s
.
M
,
N
E
,
S
K
o
r
e
a
n
L
e
s
p
e
d
e
z
a
L
e
s
p
e
d
e
z
a
s
t
i
p
u
l
a
c
e
a
a
n
n
u
a
l
/
l
e
g
u
m
e
w
a
r
m
n
o
m
e
d
.
-
w
e
l
l
2
0
+
4
.
5
-
7
.
0
H
,
R
6
-
1
2
P
e
r
s
i
s
t
e
n
t
c
o
v
e
r
I
o
w
a
-
6
v
a
r
i
e
t
y
f
o
r
c
o
l
d
s
i
t
e
s
.
A
l
l
B
i
r
d
s
f
o
o
t
T
r
e
f
o
i
l
L
o
t
u
s
c
o
r
n
i
c
u
l
a
t
u
s
E
m
p
i
r
e
p
e
r
e
n
.
/
l
e
g
u
m
e
c
o
o
l
n
o
p
o
o
r
-
w
e
l
l
1
8
-
2
0
+
5
.
0
-
7
.
5
H
,
A
,
R
,
D
6
-
8
U
s
e
a
d
a
p
t
e
d
c
u
l
t
i
v
a
r
s
.
B
o
r
o
n
t
o
l
e
r
a
n
t
.
D
o
e
s
n
o
t
c
a
u
s
e
l
i
v
e
s
t
o
c
k
b
l
o
a
t
.
N
C
,
S
C
,
W
A
l
f
a
l
f
a
M
e
d
i
c
a
g
o
s
a
t
i
v
a
p
e
r
e
n
.
/
l
e
g
u
m
e
c
o
o
l
n
o
m
e
d
.
-
w
e
l
l
1
7
-
2
0
+
6
.
5
-
7
.
5
H
,
A
,
R
4
-
1
2
U
s
e
a
d
a
p
t
e
d
c
u
l
t
i
v
a
r
s
.
N
e
e
d
s
C
a
a
n
d
P
.
D
i
s
e
a
s
e
s
u
s
c
e
p
t
i
b
l
e
,
f
o
o
d
i
n
g
i
n
t
o
l
e
r
a
n
t
.
A
l
l
W
h
i
t
e
S
w
e
e
t
C
l
o
v
e
r
M
e
l
i
l
o
t
u
s
a
l
b
a
b
i
e
n
.
/
l
e
g
u
m
e
c
o
o
l
n
o
w
e
l
l
1
6
-
2
0
+
6
.
0
-
8
.
0
A
,
R
4
-
7
I
n
v
a
s
i
v
e
a
n
d
c
o
m
p
e
t
i
t
i
v
e
.
S
c
a
r
i
f
y
s
e
e
d
.
M
a
y
c
a
u
s
e
l
i
v
e
s
t
o
c
k
b
l
o
a
t
.
M
,
N
E
,
S
Y
e
l
l
o
w
S
w
e
e
t
C
l
o
v
e
r
M
e
l
i
l
o
t
u
s
o
f
f
c
i
n
a
l
i
s
b
i
e
n
.
/
l
e
g
u
m
e
c
o
o
l
n
o
w
e
l
l
1
4
-
2
0
+
6
.
0
-
8
.
0
A
,
R
4
-
7
I
n
v
a
s
i
v
e
.
M
o
r
e
d
r
o
u
g
h
t
t
o
l
e
r
a
n
t
t
h
a
n
M
.
a
l
b
a
.
S
c
a
r
i
f
y
s
e
e
d
.
M
a
y
c
a
u
s
e
l
i
v
e
s
t
o
c
k
b
l
o
a
t
.
W
,
S
C
S
a
n
f
o
i
n
O
n
o
b
r
y
c
h
i
s
v
i
c
a
e
f
o
l
i
a
p
e
r
e
n
.
/
l
e
g
u
m
e
c
o
o
l
n
o
w
e
l
l
1
0
-
1
6
6
.
0
-
7
.
5
H
,
A
,
R
,
D
1
0
-
2
0
U
s
e
a
d
a
p
t
e
d
c
u
l
t
i
v
a
r
s
.
M
,
S
C
,
N
E
,
S
P
u
r
p
l
e
P
r
a
i
r
i
e
C
l
o
v
e
r
P
e
t
a
l
o
s
t
e
m
u
m
p
u
r
p
u
r
e
u
m
K
a
n
e
b
p
e
r
e
n
.
/
l
e
g
u
m
e
w
a
r
m
y
e
s
m
e
d
.
1
2
-
1
8
6
.
0
-
8
.
0
H
,
A
,
R
,
D
6
-
8
S
h
o
w
y
d
i
s
p
l
a
y
o
f
f
o
w
e
r
s
.
N
C
,
W
S
t
r
a
w
b
e
r
r
y
C
l
o
v
e
r
T
r
i
f
o
l
i
u
m
f
r
a
g
i
f
e
r
u
m
p
e
r
e
n
.
/
l
e
g
u
m
e
c
o
o
l
n
o
p
o
o
r
1
8
-
2
0
+
6
.
0
-
7
.
0
A
,
R
,
D
2
-
3
S
c
a
r
i
f
y
s
e
e
d
.
P
r
o
v
i
d
e
s
p
a
s
t
u
r
e
.
< PREVIOUS VIEW
10-36
Chapter 10
MAY 2009
T
A
B
L
E
1
0
.
1
3
C
H
A
R
A
C
T
E
R
I
S
T
I
C
S
O
F
C
O
M
M
O
N
S
P
E
C
I
E
S
P
O
T
E
N
T
I
A
L
L
Y
S
U
I
T
A
B
L
E
F
O
R
R
E
C
L
A
M
A
T
I
O
N
(
C
o
n
t
i
n
u
e
d
)
R
e
g
i
o
n
s
o
f
U
.
S
.
(
1
)
C
o
m
m
o
n
N
a
m
e
S
c
i
e
n
t
i
f
c
N
a
m
e
a
n
d
C
u
l
t
i
v
a
r
(
2
)
G
r
o
w
t
h
H
a
b
i
t
/
F
o
r
m
s
(
3
,
4
)
G
r
o
w
t
h
S
e
a
s
o
n
(
5
)
N
a
t
i
v
e
S
p
e
c
i
e
s
(
6
)
D
r
a
i
n
-
a
g
e
P
r
e
c
i
p
.
R
a
n
g
e
(
7
)
(
i
n
)
p
H
R
a
n
g
e
(
7
)
L
a
n
d
U
s
e
s
(
8
)
S
e
e
d
i
n
g
R
a
t
e
(
9
)
(
l
b
/
a
c
)
M
i
s
c
e
l
l
a
n
e
o
u
s
N
o
t
e
s
N
C
,
W
A
l
s
i
k
e
C
l
o
v
e
r
T
r
i
f
o
l
i
u
m
h
y
b
r
i
d
u
m
p
e
r
e
n
.
/
l
e
g
u
m
e
c
o
o
l
n
o
p
o
o
r
-
w
e
l
l
1
8
-
2
0
+
5
.
0
-
7
.
5
H
,
A
,
R
3
-
5
A
t
t
r
a
c
t
s
b
e
e
s
.
S
h
o
r
t
-
l
i
v
e
d
.
S
C
,
S
C
r
i
m
s
o
n
C
l
o
v
e
r
T
r
i
f
o
l
i
u
m
i
n
c
a
r
n
a
t
u
m
b
i
e
n
.
/
l
e
g
u
m
e
c
o
o
l
n
o
m
e
d
.
2
0
+
5
.
5
-
7
.
0
A
,
R
,
D
1
5
-
2
5
C
o
v
e
r
c
r
o
p
,
b
u
t
p
e
r
s
i
s
t
e
n
t
w
i
t
h
a
d
a
p
t
e
d
c
u
l
t
i
v
a
r
s
.
N
C
,
S
C
,
W
W
h
i
t
e
C
l
o
v
e
r
T
r
i
f
o
l
i
u
m
r
e
p
e
n
s
p
e
r
e
n
.
/
l
e
g
u
m
e
c
o
o
l
n
o
m
e
d
.
1
8
-
2
-
+
6
.
0
-
7
.
0
H
,
A
,
R
,
D
2
-
4
S
o
d
f
o
r
m
i
n
g
.
A
t
t
r
a
c
t
s
b
e
e
s
.
S
,
S
C
A
r
r
o
w
l
e
a
f
C
l
o
v
e
r
T
r
i
f
o
l
i
u
m
v
e
s
i
c
u
l
o
s
u
m
a
n
n
u
a
l
/
l
e
g
u
m
e
c
o
o
l
n
o
m
e
d
.
2
0
+
5
.
0
-
7
.
0
H
,
A
1
0
-
1
5
U
s
e
a
d
a
p
t
e
d
c
u
l
t
i
v
a
r
s
.
P
e
r
s
i
s
t
e
n
t
c
o
v
e
r
.
N
C
,
S
C
,
W
H
a
i
r
y
v
e
t
c
h
V
i
c
i
a
v
i
l
l
o
s
a
a
n
n
u
a
l
/
l
e
g
u
m
e
c
o
o
l
n
o
w
e
l
l
1
8
-
2
0
+
5
.
0
-
7
.
5
R
,
D
2
0
-
3
0
S
o
d
-
f
o
r
m
i
n
g
c
o
v
e
r
c
o
r
p
.
A
l
l
C
o
m
m
o
n
Y
a
r
r
o
w
A
c
h
i
l
l
e
a
m
i
l
l
e
f
o
l
i
u
m
p
e
r
e
n
.
/
f
o
r
b
.
w
a
r
m
y
e
s
m
e
d
.
-
w
e
l
l
1
0
-
2
0
+
5
.
0
-
6
.
0
H
,
R
,
D
1
-
2
F
o
r
m
s
g
r
o
u
n
d
c
o
v
e
r
.
D
r
o
u
g
h
t
r
e
s
i
s
t
a
n
t
.
W
W
e
s
t
e
r
n
Y
a
r
r
o
w
A
c
h
i
l
l
e
a
m
i
l
l
e
f
o
l
i
u
m
l
a
n
d
u
l
o
s
a
p
e
r
e
n
.
/
f
o
r
b
.
w
a
r
m
y
e
s
m
e
d
.
-
w
e
l
l
1
2
-
1
8
6
.
0
-
8
.
0
H
,
A
,
R
,
D
1
-
2
D
r
o
u
g
h
t
r
e
s
i
s
t
a
n
t
.
N
E
,
S
D
w
a
r
f
-
e
a
r
e
d
C
o
r
e
o
p
s
i
s
C
o
r
e
o
p
s
i
s
a
u
r
i
c
u
l
a
t
a
p
e
r
e
n
.
/
f
o
r
b
.
c
o
o
l
y
e
s
w
e
l
l
2
0
+
5
.
0
-
7
.
0
R
,
D
2
-
3
S
h
o
w
y
d
i
s
p
l
a
y
o
f
f
o
w
e
r
s
.
S
t
o
l
o
n
i
f
e
r
o
u
s
.
S
C
,
W
R
a
t
t
l
e
s
n
a
k
e
W
e
e
d
E
u
p
h
o
r
b
i
a
a
l
b
o
m
a
r
g
i
n
a
t
a
p
e
r
e
n
.
/
f
o
r
b
.
w
a
r
m
y
e
s
w
e
l
l
1
4
-
2
0
+
6
.
0
-
8
.
0
H
,
R
,
D
1
-
2
C
o
l
d
t
o
l
e
r
a
n
t
.
A
l
l
A
n
n
u
a
l
S
u
n
f
o
w
e
r
H
e
l
i
a
n
t
h
u
s
a
n
n
a
u
a
s
a
n
n
u
a
l
/
f
o
r
b
.
w
a
r
m
y
e
s
w
e
l
l
1
5
+
5
.
0
-
7
.
0
H
,
A
,
D
6
-
8
S
h
o
w
y
d
i
s
p
l
a
y
.
N
C
,
W
A
n
n
u
a
l
S
u
n
f
o
w
e
r
H
e
l
i
a
n
t
h
u
s
a
n
n
u
a
s
J
a
e
g
e
r
i
a
n
n
u
a
l
/
f
o
r
b
.
w
a
r
m
y
e
s
p
o
o
r
-
w
e
l
l
6
-
1
4
+
6
.
0
-
8
.
0
H
,
A
,
D
6
-
1
0
S
h
o
w
y
d
i
s
p
l
a
y
.
S
C
,
N
C
,
M
,
N
E
,
S
G
a
y
f
e
a
t
h
e
r
L
i
a
t
r
i
s
p
y
c
n
o
s
t
a
c
h
y
a
E
u
r
e
k
a
p
e
r
e
n
.
/
f
o
r
b
.
w
a
r
m
y
e
s
p
o
o
r
-
w
e
l
l
1
0
-
2
0
+
6
.
0
-
7
.
5
H
,
A
,
R
,
D
2
-
4
S
h
o
w
y
d
i
s
p
l
a
y
.
D
r
o
u
g
h
t
r
e
s
i
s
t
a
n
t
.
M
,
S
C
,
N
E
,
S
S
m
a
r
t
f
e
e
d
P
o
l
y
g
o
n
u
m
p
e
n
n
s
y
l
-
v
a
n
i
c
u
m
a
n
n
u
a
l
/
f
o
r
b
.
w
a
r
m
y
e
s
p
o
o
r
1
5
+
5
.
0
-
7
.
0
H
,
A
1
0
-
1
2
W
e
t
s
i
t
e
s
.
R
o
o
t
s
a
t
n
o
d
e
s
.
S
c
a
r
i
f
y
s
e
e
d
.
N
C
,
S
C
,
M
,
N
E
,
S
C
o
n
e
f
o
w
e
r
R
a
t
i
b
i
d
a
c
o
l
u
m
n
i
f
e
r
a
S
u
n
g
l
o
w
p
e
r
e
n
.
/
f
o
r
b
.
w
a
r
m
y
e
s
p
o
o
r
-
w
e
l
l
1
6
+
6
.
0
-
7
.
5
H
,
A
,
R
,
D
2
-
4
S
h
o
w
y
d
i
s
p
l
a
y
o
f
f
o
w
e
r
s
.
< PREVIOUS VIEW
10-37
Environmental Considerations
MAY 2009
T
A
B
L
E
1
0
.
1
3
C
H
A
R
A
C
T
E
R
I
S
T
I
C
S
O
F
C
O
M
M
O
N
S
P
E
C
I
E
S
P
O
T
E
N
T
I
A
L
L
Y
S
U
I
T
A
B
L
E
F
O
R
R
E
C
L
A
M
A
T
I
O
N
(
C
o
n
t
i
n
u
e
d
)
R
e
g
i
o
n
s
o
f
U
.
S
.
(
1
)
C
o
m
m
o
n
N
a
m
e
S
c
i
e
n
t
i
f
c
N
a
m
e
a
n
d
C
u
l
t
i
v
a
r
(
2
)
G
r
o
w
t
h
H
a
b
i
t
/
F
o
r
m
s
(
3
,
4
)
G
r
o
w
t
h
S
e
a
s
o
n
(
5
)
N
a
t
i
v
e
S
p
e
c
i
e
s
(
6
)
D
r
a
i
n
-
a
g
e
P
r
e
c
i
p
.
R
a
n
g
e
(
7
)
(
i
n
)
p
H
R
a
n
g
e
(
7
)
L
a
n
d
U
s
e
s
(
8
)
S
e
e
d
i
n
g
R
a
t
e
(
9
)
(
l
b
/
a
c
)
M
i
s
c
e
l
l
a
n
e
o
u
s
N
o
t
e
s
N
C
,
S
C
,
W
F
o
u
r
-
w
i
n
g
S
a
l
t
b
u
s
h
A
t
r
i
p
l
e
x
c
a
n
e
s
c
e
n
s
p
e
r
e
n
.
/
s
h
r
u
b
c
o
o
l
y
e
s
p
o
o
l
-
w
e
l
l
5
-
1
5
7
.
0
-
8
.
0
H
,
A
,
R
,
D
8
-
1
2
U
s
e
a
d
a
p
t
e
d
e
c
o
t
y
p
e
s
.
E
v
e
r
g
r
e
e
n
.
S
e
e
d
s
p
r
o
l
i
f
c
.
D
r
o
u
g
h
t
t
o
l
e
r
a
n
t
.
V
e
r
y
s
a
l
t
t
o
l
e
r
a
n
t
.
S
C
,
W
S
a
l
t
b
u
s
h
s
p
e
c
i
e
s
A
t
r
i
p
l
e
x
s
p
p
.
p
e
r
e
n
.
/
s
h
r
u
b
c
o
o
l
y
e
s
p
o
o
r
-
w
e
l
l
6
-
1
2
7
.
0
-
8
.
5
H
,
A
,
R
,
D
8
-
1
2
S
o
m
e
s
p
e
c
i
e
s
d
e
c
i
d
u
o
u
s
.
N
C
,
S
C
,
W
S
i
b
e
r
i
a
n
P
e
a
S
h
r
u
b
C
a
r
a
g
a
n
a
a
r
b
o
r
e
s
c
e
n
s
p
e
r
e
n
.
/
s
h
r
u
b
w
a
r
m
n
o
w
e
l
l
1
2
-
2
0
+
4
.
0
-
8
.
5
H
,
A
,
R
,
D
1
-
2
A
b
o
r
e
a
l
,
d
e
c
i
d
u
o
u
s
.
W
A
l
k
a
l
i
R
a
b
b
i
t
b
r
u
s
h
C
h
r
y
s
o
t
h
a
m
n
u
s
n
a
u
s
e
o
s
u
s
c
o
n
s
i
m
i
l
u
s
p
e
r
e
n
.
/
s
h
r
u
b
c
o
o
l
y
e
s
w
e
l
l
8
-
1
8
+
6
.
0
-
8
.
0
H
,
R
,
D
1
-
2
D
e
c
i
d
u
o
u
s
.
D
e
e
p
r
o
o
t
s
.
N
C
,
S
C
,
W
R
u
s
s
i
a
n
O
l
i
v
e
E
l
e
a
g
n
u
s
a
n
g
u
s
t
i
f
o
l
i
a
p
e
r
e
n
.
/
t
r
e
e
w
a
r
m
n
o
p
o
o
r
-
w
e
l
l
1
0
-
2
0
+
6
.
0
-
8
.
0
A
,
R
,
D
1
-
2
I
n
v
a
s
i
v
e
,
d
e
c
i
d
u
o
u
s
,
s
h
r
u
b
b
y
.
N
-
f
x
e
r
.
N
C
,
S
C
,
W
W
i
n
t
e
r
f
a
t
E
u
r
o
t
i
a
l
a
n
a
t
a
p
e
r
e
n
.
/
s
h
r
u
b
c
o
o
l
y
e
s
w
e
l
l
6
-
1
5
+
6
.
0
-
8
.
0
H
,
A
1
-
2
E
v
e
r
g
r
e
e
n
.
N
o
t
f
o
r
M
o
n
t
a
n
a
.
W
G
r
a
y
M
o
l
l
K
o
c
h
i
a
A
m
e
r
i
c
a
n
a
V
e
s
t
i
t
a
p
e
r
e
n
.
/
s
h
r
u
b
w
a
r
m
y
e
s
p
o
o
r
-
w
e
l
l
4
-
2
0
6
.
0
-
9
.
0
H
,
A
,
R
,
D
1
-
2
A
l
k
a
l
i
n
e
s
i
t
e
s
.
N
C
,
S
C
,
W
S
u
m
m
e
r
C
y
p
r
e
s
s
K
o
c
h
i
a
p
r
o
s
t
r
a
t
a
p
e
r
e
n
.
/
s
h
r
u
b
w
a
r
m
n
o
w
e
l
l
1
2
-
1
8
6
.
0
-
8
.
0
R
,
D
.
1
-
2
L
o
n
g
-
l
i
v
e
d
v
e
r
y
d
r
o
u
g
h
t
t
o
l
e
r
a
n
t
.
P
r
o
s
t
r
a
t
e
f
o
r
m
.
N
C
,
S
C
,
W
W
e
e
p
i
n
g
M
u
l
b
e
r
r
y
M
o
r
u
s
a
l
b
a
K
i
n
g
a
n
p
e
r
e
n
.
/
t
r
e
e
w
a
r
m
n
o
p
o
o
r
-
w
e
l
l
1
2
-
2
0
+
5
.
0
-
6
.
5
H
,
R
,
D
1
-
2
D
e
c
i
d
u
o
u
s
.
N
E
,
M
,
S
,
W
N
o
r
w
a
y
S
p
r
u
c
e
P
i
c
e
a
a
b
i
e
s
p
e
r
e
n
.
/
t
r
e
e
w
a
r
m
n
o
m
e
d
.
-
p
o
o
r
1
6
-
2
0
+
5
.
0
-
6
.
0
R
,
D
E
v
e
r
g
r
e
e
n
.
C
o
l
d
s
i
t
e
s
.
P
l
a
n
t
s
e
e
d
l
i
n
g
s
.
A
l
l
L
o
m
b
a
r
d
y
P
o
p
l
a
r
P
o
p
l
u
s
n
i
g
r
a
I
t
a
l
i
c
a
p
e
r
e
n
.
/
t
r
e
e
w
a
r
m
n
o
w
e
l
l
1
2
-
2
0
5
.
0
-
7
.
0
R
,
D
D
e
c
i
d
u
o
u
s
,
s
h
o
r
t
-
l
i
v
e
d
.
P
l
a
n
t
s
e
e
d
l
i
n
g
s
.
S
C
,
S
L
i
v
e
O
a
k
Q
u
e
r
c
u
s
v
i
r
g
i
n
i
a
n
a
p
e
r
e
n
.
/
t
r
e
e
w
a
r
m
y
e
s
m
e
d
.
1
5
-
2
0
+
6
.
0
-
7
.
0
H
,
R
,
D
1
-
2
E
v
e
r
g
r
e
e
n
.
S
l
o
w
g
r
o
w
i
n
g
.
N
C
,
S
C
,
M
B
u
r
O
a
k
Q
u
e
r
c
u
s
m
a
c
r
o
c
a
r
p
a
p
e
r
e
n
.
/
t
r
e
e
w
a
r
m
y
e
s
m
e
d
.
1
0
-
2
0
4
.
0
-
7
.
0
H
,
R
,
D
1
-
2
D
e
c
i
d
u
o
u
s
.
P
l
a
n
t
a
c
o
r
n
s
o
r
s
e
e
d
l
i
n
g
s
.
M
,
N
E
,
S
B
r
i
s
t
l
y
L
o
c
u
s
t
R
o
b
i
n
a
f
e
r
t
i
l
i
s
A
r
n
o
t
p
e
r
e
n
.
/
s
h
r
u
b
w
a
r
m
y
e
s
w
e
l
l
1
5
+
3
.
5
-
7
.
5
H
,
R
,
D
2
-
5
D
e
c
i
d
u
o
u
s
.
F
o
r
m
s
t
h
i
c
k
e
t
s
.
N
-
f
x
e
r
.
S
c
a
r
i
f
y
s
e
e
d
.
< PREVIOUS VIEW
10-38
Chapter 10
MAY 2009
T
A
B
L
E
1
0
.
1
3
C
H
A
R
A
C
T
E
R
I
S
T
I
C
S
O
F
C
O
M
M
O
N
S
P
E
C
I
E
S
P
O
T
E
N
T
I
A
L
L
Y
S
U
I
T
A
B
L
E
F
O
R
R
E
C
L
A
M
A
T
I
O
N
(
C
o
n
t
i
n
u
e
d
)
R
e
g
i
o
n
s
o
f
U
.
S
.
(
1
)
C
o
m
m
o
n
N
a
m
e
S
c
i
e
n
t
i
f
c
N
a
m
e
a
n
d
C
u
l
t
i
v
a
r
(
2
)
G
r
o
w
t
h
H
a
b
i
t
/
F
o
r
m
s
(
3
,
4
)
G
r
o
w
t
h
S
e
a
s
o
n
(
5
)
N
a
t
i
v
e
S
p
e
c
i
e
s
(
6
)
D
r
a
i
n
-
a
g
e
P
r
e
c
i
p
.
R
a
n
g
e
(
7
)
(
i
n
)
p
H
R
a
n
g
e
(
7
)
L
a
n
d
U
s
e
s
(
8
)
S
e
e
d
i
n
g
R
a
t
e
(
9
)
(
l
b
/
a
c
)
M
i
s
c
e
l
l
a
n
e
o
u
s
N
o
t
e
s
A
l
l
B
l
a
c
k
L
o
c
u
s
t
R
o
b
i
n
i
a
p
s
e
u
d
o
a
c
a
c
i
a
p
e
r
e
n
.
/
t
r
e
e
w
a
r
m
y
e
s
p
o
o
r
-
w
e
l
l
1
5
+
4
.
0
-
7
.
5
R
,
D
1
-
3
D
e
c
i
d
u
o
u
s
.
N
-
f
x
e
r
.
S
c
a
r
i
f
y
s
e
e
d
.
M
o
s
t
c
o
m
m
o
n
r
e
c
l
a
m
a
t
i
o
n
t
r
e
e
.
S
C
,
W
T
a
m
a
r
i
s
k
s
p
e
c
i
e
s
T
a
m
a
r
i
x
s
p
p
.
p
e
r
e
n
.
/
s
h
r
u
b
w
a
r
m
n
o
p
o
o
r
-
w
e
l
l
8
-
2
0
6
.
5
-
7
.
5
H
,
R
1
-
2
A
b
o
r
e
a
l
,
e
v
e
r
g
r
e
e
n
.
N
o
t
e
:
1
.
R
e
g
i
o
n
s
o
f
t
h
e
U
.
S
.
:
M
=
m
i
d
-
w
e
s
t
e
r
n
s
t
a
t
e
s
(
I
A
,
I
L
,
I
N
,
K
Y
,
M
I
,
M
N
,
M
O
,
O
H
,
W
I
)
N
C
=
n
o
r
t
h
-
c
e
n
t
r
a
l
s
t
a
t
e
s
(
N
D
,
N
E
,
S
D
)
N
E
=
n
o
r
t
h
e
a
s
t
e
r
n
s
t
a
t
e
s
(
C
T
,
D
E
,
M
A
,
M
D
,
M
E
,
N
H
,
N
J
,
N
Y
,
P
A
,
R
I
,
W
T
,
W
V
)
S
=
s
o
u
t
h
e
r
n
s
t
a
t
e
s
(
A
R
,
A
L
,
F
L
,
G
A
,
L
A
,
M
S
,
N
C
,
S
C
,
T
N
,
V
A
)
S
C
=
s
o
u
t
h
-
c
e
n
t
r
a
l
s
t
a
t
e
s
(
K
S
,
O
K
,
T
X
)
W
=
w
e
s
t
e
r
n
s
t
a
t
e
s
(
A
Z
,
C
A
,
C
O
,
I
D
,
M
T
,
N
M
,
N
V
,
O
R
,
U
T
,
W
A
,
W
Y
)
2
.
S
p
e
c
i
e
s
a
n
d
c
u
l
t
i
v
a
r
s
m
a
y
v
a
r
y
f
o
r
s
i
t
e
-
s
p
e
c
i
f
c
c
o
n
d
i
t
i
o
n
s
.
C
o
n
s
u
l
t
l
o
c
a
l
a
g
r
o
n
o
m
y
e
x
t
e
n
s
i
o
n
a
g
e
n
t
s
.
3
.
A
n
n
u
a
l
s
p
e
c
i
e
s
a
r
e
b
e
s
t
s
u
i
t
e
d
a
s
c
o
v
e
r
c
r
o
p
s
.
P
e
r
e
n
n
i
a
l
s
p
e
c
i
e
s
a
r
e
b
e
s
t
f
o
r
p
e
r
s
i
s
t
e
n
t
c
o
v
e
r
.
4
.
G
r
a
s
s
e
s
s
h
o
u
l
d
b
e
p
l
a
n
t
e
d
i
n
c
o
m
b
i
n
a
t
i
o
n
w
i
t
h
l
e
g
u
m
e
s
a
n
d
o
t
h
e
r
f
o
r
b
s
.
L
e
g
u
m
e
s
s
h
o
u
l
d
b
e
i
n
o
c
u
l
a
t
e
d
w
i
t
h
a
p
p
r
o
p
r
i
a
t
e
r
h
i
z
o
b
i
u
m
a
t
f
v
e
t
i
m
e
s
t
h
e
r
e
c
o
m
m
e
n
d
e
d
r
a
t
e
.
M
a
n
y
a
d
d
i
t
i
o
n
a
l
f
o
r
b
s
,
s
h
r
u
b
s
,
a
n
d
t
r
e
e
s
a
r
e
a
v
a
i
l
a
b
l
e
a
n
d
s
u
i
t
a
b
l
e
f
o
r
r
e
c
l
a
m
a
t
i
o
n
.
5
.
W
a
r
m
s
e
a
s
o
n
g
r
a
s
s
e
s
g
e
n
e
r
a
l
l
y
a
r
e
m
o
r
e
t
o
l
e
r
a
n
t
o
f
d
r
o
u
g
h
t
a
n
d
w
a
t
e
r
s
t
r
e
s
s
.
6
.
N
a
t
i
v
e
d
i
s
t
r
i
b
u
t
i
o
n
n
o
t
e
d
a
s
o
c
c
u
r
s
i
n
t
h
e
p
a
r
t
i
c
u
l
a
r
r
e
g
i
o
n
.
7
.
R
a
n
g
e
s
a
r
e
i
n
d
i
c
a
t
e
d
f
o
r
o
p
t
i
m
a
l
g
r
o
w
t
h
c
o
n
d
i
t
i
o
n
s
a
n
d
m
a
y
b
e
w
i
d
e
r
f
o
r
s
o
m
e
s
p
e
c
i
e
s
.
8
.
P
o
s
t
-
c
l
o
s
u
r
e
l
a
n
d
u
s
e
s
.
M
u
c
h
o
v
e
r
l
a
p
i
s
p
o
s
s
i
b
l
e
:
H
=
w
i
l
d
l
i
f
e
h
a
b
i
t
a
t
a
n
d
o
p
e
n
s
p
a
c
e
A
=
a
g
r
i
c
u
l
t
u
r
a
l
R
=
r
e
c
r
e
a
t
i
o
n
a
l
D
=
s
i
t
e
d
e
v
e
l
o
p
m
e
n
t
9
.
R
a
t
e
i
n
d
i
c
a
t
e
d
f
o
r
s
e
e
d
m
i
x
t
u
r
e
s
.
H
i
g
h
e
r
r
a
t
e
s
a
p
p
l
y
f
o
r
m
o
n
o
t
y
p
i
c
s
t
a
n
d
s
.
S
e
e
d
i
n
g
r
a
t
e
s
a
n
d
m
i
x
e
s
s
h
o
u
l
d
b
e
d
e
t
e
r
m
i
n
e
d
b
y
s
i
t
e
-
s
p
e
c
i
f
c
c
o
n
d
i
t
i
o
n
s
.
C
o
n
s
u
l
t
l
o
c
a
l
a
g
r
o
n
o
m
y
a
g
e
n
t
s
.
(
D
I
G
I
O
I
A
E
T
A
L
.
,
1
9
9
5
)
< PREVIOUS VIEW
10-39
Environmental Considerations
MAY 2009
the quick-cover species may choke out and prevent successful long-term establishment of perennial
species. A balance between short-term erosion control from quick-cover annuals and long-term self-
sustaining perennials can be achieved by two-step seeding. This involves an initial dense planting of
quick-cover annuals, allowing them to be winter-killed, and then seeding perennials into the stubble
remaining from the annuals the following spring. Plants and recommended cultivation practices for
humid regions are discussed in Vogel (1987) and Bennet et al. (1978).
In arid and semi-arid areas, exceptionally drought- and salt-tolerant species should be selected (Packer
and Aldon, 1978). Even if adaptable species are used, high seeding and planting densities without
supplementary irrigation can lead to excessive water stress and failure. Supplementary irrigation
may be necessary until root systems are developed. Furrowing along contour lines and planting in
the furrows will generally result in efcient use of irrigation water and natural precipitation.
Selection of vegetation for impounding embankments should take into account potential impacts
on dam safety inspection and performance. Inspection of vegetated surfaces of dams and adjacent
areas, particularly the crest, downstream slope, toe, and adjacent foundation areas is important, as
discussed in Section 12.3. Trees and woody vegetation are detrimental to both inspection and the
long-term durability of the embankment. Grasses and shallow rooted native vegetation are the most
desirable surface cover for an active impounding embankment and dam. Guidance on this issue is
presented in Marks and Tschantz (2002).
< PREVIOUS VIEW
Chapter 11
CONSTRUCTION AND
DISPOSAL OPERATIONS
11-1 MAY 2009
The preparation of clear and concise construction drawings, specifcations and an operation and
maintenance plan to guide contractors and mine personnel is an essential part of coal refuse facility
development and operation. Property procurement, selection of hauling and transport equipment,
determination of haul road locations, determination of refuse disposal procedures, and scheduling
of construction for the various facility components are among the frst steps of coal refuse disposal
facility planning. These planning steps must refect long-term operational needs as well as short-term
start-up requirements.
Communication of this information to the personnel responsible for implementing the plans is essen-
tial for efective operations. A basic understanding of the general design concepts and potential safety
hazards associated with a disposal facility will increase the likelihood that a program of monitoring
and construction control will be conscientiously followed. Regular monitoring and on-site testing
will facilitate construction and refuse disposal. Monitoring data related to disposal facility develop-
ment will aid in compliance with regulatory agency requirements and will facilitate modifcations
to facility development should changes occur in mine production. Table 11.1 presents a summary of
refuse facility operational considerations and related sections in this Manual, as well as references by
other authors.
11.1 OPERATIONAL PLANS AND CONSTRUCTION DOCUMENTS
Constructing and operating a coal refuse disposal facility in accordance with regulatory requirements
while meeting the mines operational needs requires careful development of construction plans, speci-
fcations and an operation and maintenance plan. Development of these construction documents
requires that the designer collect, interpret and analyze data related to design requirements, regula-
tory requirements, site conditions and the mines operational needs. Input from the mine operator is
critical throughout the design and permiting process so that mine refuse production, particularly
start-up rates, are accommodated. Important aspects of refuse disposal facility design should be docu-
mented in a design report accompanying the construction documents. This report should present the
rationale for facility design, important assumptions and constraints, and engineering calculations sup-
porting facility design.
The plans and specifcations should provide essential information and should be organized and pre-
sented in such a way that they can be easily interpreted by mine personnel, contractors, regulatory
agencies, and monitoring personnel. The operation and maintenance plan should provide important
< PREVIOUS VIEW
11-2
Chapter 11
MAY 2009
TABLE 11.1 REFUSE DISPOSAL OPERATIONS
Operational Considerations
Manual
Sections
Supplemental References
I. Operational Plans and Scheduling
Determine the scope of activities 11.1 to 11.2
Review equipment capabilities 11.3 to11.5 OBrien et al. (1996), Hartman (1992)
Schedule development 11.2 Clough et al. (2008), Day and Benjamin (1991)
II. Equipment Selection and Use
For transport of refuse from preparation
plant to disposal area
11.3 OBrien et al. (1996), Day and Benjamin (1991),
Hartmann (1992)
For spreading and compacting refuse 11.4 OBrien et al. (1996), Day and Benjamin (1991),
Hartmann (1992)
For multiple use in related earth work
activities
11.4 USBR (1998), Sherard et al. (1963), Church (1981)
III. Refuse Disposal
Disposal as structural fll 11.5 USBR (1991, 1998), Sherard et al. (1963),
Leonards (1962), MSHA (2007)
Fine refuse disposal 11.5 Hartmann (1992), MSHA (2007)
Combined refuse disposal 11.5 Hartmann (1992), MSHA (2007)
IV. Related Activities
Haul road construction and maintenance 11.6 Leonards (1962), Oglesby and Hicks (1982)
Liner systems (if required) 11.6 USBR (1998)
Embankment foundation preparation 11.6 Sherard et al. (1963), USBR (1998), Leonards (1962)
Control of water 11.6 USBR (1998)
General excavation 11.6 Sherard et al. (1963), Day and Benjamin (1991),
OBrien et al. (1996)
Placement of earth borrow as structural fll 11.6 Sherard et al. (1963), USBR (1998)
V. Materials Selection for Facility
Appurtenances
Filters and drains 11.7 MSHA (2007)
Culverts and decants 11.7 MSHA (2007), FEMA (2005a)
Concrete structures 11.7 USBR (1987a)
Gates, valves and other metal works 11.7 USBR (1987a)
Mine opening and auger-hole seals and
drains
11.7 MSHA (2007)
VI. Quality Control and Field Testing
Testing of compacted earth and refuse fll 11.8 USBR (1998), Leonards (1962), Sherard et al. (1963)
Materials testing 11.8 USBR (1998), USBR (1991)
< PREVIOUS VIEW
11-3
Construction and Disposal Operations
MAY 2009
information in a concise and easily understandable format so that it can quickly be used as an ef-
cient reference guide for mine personnel performing construction and conducting construction and
performance monitoring activities.
Ambiguous details, terms and statements in the specifcations and drawings should be avoided. The
extent of information and details provided in the plans and specifcations will vary depending upon
the size and complexity of the refuse disposal facility. The following sections of this chapter present
guidance as to the content that should be provided in plan drawings, specifcations and operation
and maintenance plans for refuse disposal facilities.
An expansion or modifcation plan is sometimes prepared for an existing coal refuse disposal facility
to address changes in facility construction associated with revisions to production rates or changed
design criteria. Such plans typically include construction drawings, specifcations, and a design report
with supporting engineering calculations. For expansion or modifcation plans, it is important that an
index of drawings citing all remaining original and new plan drawings be provided. Revised draw-
ings should clearly reference appropriate specifcations, and a revised set of specifcations should be
included in the associated submitals. If facility drawings are modifed or revised as part of the new
submission, it is recommended that a current index and cover sheet be prepared with a complete set
of drawings to prevent confusion during subsequent review and construction.
11.1.1 Design Report
The design report should present the background data upon which the design is based, along with
critical aspects of the design related to construction, operation, and abandonment. The design report
generally includes a discussion of the following:
Project requirements (disposal facility requirements such as refuse generation rates
and anticipated life of mine)
Existing conditions and history of the site including general geologic seting, envi-
ronmental seting (wetlands, streams or other sensitive areas) previous mining
activities, and climate/weather
Field investigations such as surfcial reconnaissance and geotechnical and environ-
mental explorations
Laboratory testing of refuse, water, soil, rock, geosynthetic materials, and structural
materials
Geologic and geotechnical site conditions, including assessment of impact of previ-
ous and planned mining activities
Site limitations based upon reconnaissance, explorations and testing results
Facility development and staging
Special considerations (e.g., amendments, liner systems, mine bulkheads and barri-
ers, construction issues such as dewatering)
Equipment considerations (e.g., pumping equipment if needed for meeting
impoundment drawdown requirements, equipment for spreading and compaction
consistent with coal refuse production when conveyor systems are employed for
hauling)
Geotechnical engineering analyses including slope stability, setlement and seepage
analyses (and pillar stability and subsidence analyses if underground mining is a
factor)
Hydrologic and hydraulic analyses (including design storm, dam breach and inun-
dation mapping) and hydraulic structure design
< PREVIOUS VIEW
11-4
Chapter 11
MAY 2009
Structural design and detailed monitoring provisions for buried pipelines and inlet
structures
Environmental issues
Abandonment requirements
Monitoring and maintenance requirements (schedules for implementation and maxi-
mum recommended readings for instrumentation)
Design recommendations and limitations
11.1.2 Construction Drawings
Construction drawings for a coal refuse disposal facility are probably the most important component of
the construction documents. Construction drawings will likely be referred to more ofen by mine per-
sonnel, inspectors, engineers and regulatory agencies than any other construction documents. Accurate
and detailed depiction of refuse facility design will clearly indicate the intent of the design and will
facilitate construction of the facility. Construction drawings must provide information as required in
30 CFR 77.215 and 77.216 and must satisfy applicable state regulatory agency requirements. 30 CFR
77.216 also indicates specifc information that must be presented, including the locations of surface
and underground coal mine workings and the depth and extent of such workings within a distance of
500 feet from the perimeter of the facility. Construction drawings generally show initial site develop-
ment and intermediate construction steps with enlargements or details of critical construction features
and delineation of dimensions and materials. The following sections provide guidance as to the type of
information and level of detail that should be provided on construction drawings.
11.1.2.1 Title Sheet and Existing Conditions Plans
Construction drawings should include a Title Sheet listing drawings and other project and site refer-
ence information. A plan showing existing conditions at the site should be prepared, and each refuse
facility structure should be identifed on an USGS 7.5-minute or 15-minute topographic quadrangle
map as a general location reference. This is a requirement for all impounding facilities per 30 CFR
77.216. Plan drawings depicting existing conditions at a proposed refuse disposal site should gener-
ally be prepared at a scale of 1 inch = 100 feet (state regulatory requirements may be more stringent).
Alternate scales (either smaller or larger) may be appropriate provided that existing features can be
accurately located and represented and regulatory requirements are satisfed. Existing conditions
plans should provide coverage of the proposed refuse disposal area and support areas and an addi-
tional 500-foot-wide area around the perimeter of the site.
In general, existing features that impact site development and construction of the refuse disposal
facility should be shown on existing conditions plans. The following guideline describes the general
level and type of information that should be included:
Topographic contours of existing ground surface at 5-foot intervals or less (contours
at 10-foot intervals may be appropriate for steep slopes)
Footprint of proposed refuse facility for all proposed embankment stages
Existing and proposed new underground mine workings within 500 feet from the
perimeter of the proposed facility
Extent and location of spoil piles and surface, auger, or highwall mining beneath the
dam and impoundment areas
Location of coal seam outcrops
Reported geologic features or observed structures (e.g., joints, hillseams, etc.)
Surface and subsurface utilities
Existing refuse disposal facilities
< PREVIOUS VIEW
11-5
Construction and Disposal Operations
MAY 2009
Property boundaries
Prospective borrow areas
Existing dams and embankments
Delineation of forests/woods and heavy vegetation
Existing buildings and structures
Oil, gas and water wells (active and abandoned)
Watershed limits, streams and wetlands
Springs, seeps and mine discharges
Landslide areas, mine subsidence features, and other ground disturbance
Exploratory boring and test pit locations
Public roads and mine access/haul roads
Mine shafs, boreholes, vent holes and other mine openings
Other identifable natural or man-made features (e.g., cemeteries, buildings, etc.) that
could afect the operation of the refuse facility
11.1.2.2 Initial Site Development Plans
As explained throughout this Manual, extensive site development may be required for preparing a
site to receive coal refuse material. To provide a stable embankment foundation and safe working
area for the placement of refuse and efcient operation of heavy equipment, site development con-
struction may be necessary. The following items should be considered for presentation on initial site
development plans at the same scale or larger than the existing conditions plans:
Areas requiring clearing, grubbing and topsoil stripping
Topsoil and excess material stockpile areas
Utility line modifcations or relocations
Oil, gas and water well modifcations or abandonment details
Areas requiring subsurface drainage such as spring/seep collection systems and
other underdrain systems
Prospective borrow areas
Stream diversions
Access/haul road alignments and proposed grading contours
Location(s) of survey control points
Foundation and abutment areas requiring special subgrade preparations or treatment
Barriers or backfll areas for treatment of mine openings (shafs, boreholes, auger
holes, drif mine openings) or underground mine workings
Initial erosion and sedimentation control measures
Diversion and collection ditch alignments
Environmental barriers or bufer zones
Cutof and key trench location (for impounding facilities)
Seepage barrier locations and requirements
General subgrade preparations including benching (keying) requirements
Subsurface drainage (spring, seep or stream drains)
Limits of liner installation (clay or synthetic) if applicable per state regulations
Decant pipe (with riser locations and elevations), principal spillway, and emergency
spillway alignments
< PREVIOUS VIEW
11-6
Chapter 11
MAY 2009
Some of the items listed above will probably require depiction with additional plans and details.
To meet both operational and environmental requirements, some larger coal refuse disposal sites
may need to have phased initial site development plans. These plans should be organized such that
site development requirements are depicted sequentially. Sections 11.1.2.5 and 11.1.2.6 address addi-
tional drawings that may be required and the level of detail that should be provided to supplement
plan view drawings.
11.1.2.3 Refuse Embankment Construction Plan Views
Proper presentation of refuse embankment designs, whether for an impounding or non-impound-
ing facility, will normally require more than one plan view drawing. As indicated in Chapter 5,
coal refuse embankments are typically constructed in stages. Each stage of a coal refuse embank-
ment should be depicted by at least one plan drawing at the same scale or smaller than used in the
existing conditions or initial site development plan drawings. Detailed plan views, cross sections,
profles and larger-scale details should be used to depict the design of refuse embankments and
appurtenances. The required drawings for impounding facilities (including slurry cell facilities)
will generally be greater in number and detail than for non-impounding facilities. Detailed draw-
ings of impounding facilities should normally be prepared at a scale of 1 inch = 100 feet, although a
smaller scale may be appropriate for larger sites if sufcient detail is provided in cross sections and
enlargements. As a guideline, the following items should be presented on embankment construc-
tion drawings:
Footprint and limits of each embankment stage and associated haul road or conveyor
belt system
Delineation of the embankment and material zones
Proposed topographic contours of each stage at intervals at 5 feet or less
Footprint, grades and elevations, and fow direction of internal drains
Collection and diversion ditch alignments, grades and elevations
Decant pipe alignment and riser locations
Decant riser schedule with allowable inlet elevations for each embankment stage
Spillway location and alignment
Drainage requirements for the working surface
Piezometer and other instrument locations
Tabulation of allowable phreatic surface levels measured at piezometers
Tabulation of maximum pool level for each stage under design storm conditions
Anticipated fne refuse level at end of each stage, especially for impounding facilities
with upstream construction
Work area sequence within each stage of construction where multiple zones of place-
ment are specifed (upstream, downstream or centerline)
Construction items and sequence for each stage, including features such as seepage
barriers, mine barriers and seals, etc.
Haul road and conveyor belt locations and planned extensions
The plan view drawings for each construction stage should include references to cross sections and
details for complex features of the design.
11.1.2.4 Cross Sections and Profles
Cross sections and profles are normally prepared to provide details not depicted on plan views.
Cross sections are very useful for delineating the various stages of facility development. Cross sec-
< PREVIOUS VIEW
11-7
Construction and Disposal Operations
MAY 2009
tions are also useful for illustrating complex features of refuse facility design such as cutof trenches,
subsurface drainage/underdrains, decant pipe installations, and subgrade preparations.
Consistent with 30 CFR 77.216, at least two cross sections including longitudinal and lateral cross
sections through the highest and lowest elevations of a refuse embankment must be provided. Ad-
ditional cross sections may be needed for complex refuse embankments, especially impounding em-
bankments. The number of additional cross sections is dependent on the complexity of the facility
design and the construction required for tying into previously constructed features. The following
items should be considered for inclusion on refuse embankment cross sections:
Soil and rock units with meaningful descriptions below and at foundation grade
Elevations and locations of any underground mines (this may require a separate
cross section, as the depth to the mine workings may require too large a scale to
depict other items)
Original ground surface (and existing ground surface if diferent from original
ground surface)
Foundation improvements and special subgrade preparations
General subgrade preparations
Embankment surfaces and slopes depicting material zones, if applicable
Upstream and downstream slope and abutment protection measures
Delineation of each embankment stage
Maximum and minimum crest elevations for each embankment stage
Anticipated fne refuse level at the end of construction for each embankment stage
For upstream stages, the approximate mixing zone of coarse and fne refuse associ-
ated with pushout and development of the embankment stage
Maximum normal and design storm pool level for each stage
Final surface grade upon reclamation including provisions for drainage control
Underdrains/subsurface drainage features and internal drainage systems
Decant pipeline and risers
Spillways, including dimension, peak water surface elevation and erosion protection
Terraces/benches on embankments
Piezometer locations with sensing zone elevations
Proposed support equipment or structures (conveyor belts, load out bins, haul roads, etc.)
Liner systems
Design phreatic surface
Stability analysis results depicting subsurface conditions/properties, critical failure
surfaces (circular, block and wedge surfaces) and minimum factors of safety
Profles are useful when presenting the design of haul/access roads, decant pipe alignments, ditch
alignments, and underdrains and other components that extend a great distance laterally. As a guide-
line, profles should include the following information:
Original ground surface
Proposed/fnal ground surface
Grade breaks with corresponding slope designations
Geometry and layout data (such as radii, point of vertical intersection, etc.)
Bedding and backfll details for pipes and spillways
< PREVIOUS VIEW
11-8
Chapter 11
MAY 2009
Erosion protection measures for ditches and spillways (if not shown on detailed
cross sections)
Critical subsurface soil and rock conditions and site development requirements
The scale for cross sections and profles should be the same for both the vertical and horizontal directions,
although exaggerated cross section segments may be appropriate at large sites where there is signifcant
relief. The scale should be sufcient for clear and accurate representation of design features. If necessary,
more detailed, larger-scaled cross sections should be prepared for critical facility components.
11.1.2.5 Detailed Plan Views and Cross Sections
For instances when the plan view and cross-section drawings for individual stages of a facility do not
convey the level of detail for critical design features desired by the designers, larger-scale plan views
and cross sections should be prepared. The following are facility components or features for which
detailed plan views and cross sections may be needed:
Cutof trench/keyway cuts, butresses and other stabilization structures
Subsurface drains and underdrain systems including spring/seep drains
Soil liners and graded flters
Collection and diversion ditches
Terraces/benches on the refuse embankment
Subgrade and foundation preparations
Treatment provisions (backfll, barrier, drainage, etc.) for mine workings and openings
Spillways
Decant pipes and risers, including bedding/backfll zones, thrust blocks, and trashracks
Stilling basins
Filter diaphragms
Haul/access roads
Piezometers, weirs and other instrumentation
Embankment and impoundment capping details
Stream diversions
Erosion and sedimentation controls
Culverts and other piping systems
11.1.2.6 Miscellaneous Details and Information
Documentation for impounding refuse disposal facilities should include a stage-area curve and a
stage-volume curve for the impoundment and a stage-volume curve for the embankment. For facili-
ties that rely on storage of all or part of the design storm runof, notation of the maximum decant
level inlet, open channel spillway inlet (if present), and the design storm volume (or portion thereof
that must be stored) should be included on the appropriate curves to demonstrate that sufcient stor-
age is available without overtopping the embankment. Additionally, the head-discharge curve for the
decant and spillway should be presented, if applicable. These ploted data should be provided in the
operation and maintenance plan, as discussed in Section 11.1.4.
If special construction methods or items are required, they should be detailed in the drawings.
Facility components/details such as berms, pipe beddings, piezometers, V-notched weirs, staf and
rain gauges, clear water cells, and sealing of mine openings may require additional drawing details
and information related to their installation or construction.
< PREVIOUS VIEW
11-9
Construction and Disposal Operations
MAY 2009
11.1.3 Technical Specifcations
A complete set of technical specifcations that corresponds directly to the construction drawings
should be prepared for the construction of a refuse disposal facility. Similar to the construction draw-
ings, the level of detail required in the specifcations is a function of the type and complexity of the
refuse facility being constructed. At a minimum, the critical construction requirements that impact
dam safety and facility operation, as cited in Section 11.2.2, should be clearly addressed in the speci-
fcations. Any information related to construction sequencing or methods that are recommendations,
but not requirements, should be cited as such in appropriate locations in the plan.
There are several standardized specifcation systems ranging from basic to sophisticated. Specifcations
for refuse disposal facilities must provide sufcient detail that contractors and/or mine personnel can
easily and clearly understand the facility design requirements. It is recommended that specifcations
follow a consistent format. The following are industry accepted standard specifcation systems that
may be applicable and are regularly used for other civil engineering projects:
CSI by the Construction Specifcations Institute
SPECTEXT
r
o
c
k
f
l
l
s
6
t
o
1
2
(
s
o
i
l
)
t
o
3
6
(
r
o
c
k
)
3
t
o
5
4
t
o
6
V
i
b
r
a
t
i
n
g
B
a
s
e
p
l
a
t
e
C
o
m
p
a
c
t
o
r
s
F
o
r
c
o
a
r
s
e
-
g
r
a
i
n
e
d
s
o
i
l
s
w
i
t
h
l
e
s
s
t
h
a
n
a
b
o
u
t
1
2
p
e
r
c
e
n
t
p
a
s
s
i
n
g
N
o
.
2
0
0
s
i
e
v
e
.
B
e
s
t
s
u
i
t
e
d
f
o
r
m
a
t
e
r
i
a
l
s
w
i
t
h
4
t
o
8
p
e
r
c
e
n
t
p
a
s
s
i
n
g
N
o
.
2
0
0
s
i
e
v
e
,
p
l
a
c
e
d
t
h
o
r
o
u
g
h
l
y
w
e
t
.
8
t
o
1
0
3
c
o
v
e
r
a
g
e
s
S
i
n
g
l
e
p
a
d
s
o
r
p
l
a
t
e
s
s
h
o
u
l
d
w
e
i
g
h
n
o
l
e
s
s
t
h
a
n
2
0
0
l
b
.
M
a
y
b
e
u
s
e
d
i
n
t
a
n
d
e
m
w
h
e
r
e
w
o
r
k
i
n
g
s
p
a
c
e
i
s
a
v
a
i
l
a
b
l
e
.
F
o
r
c
l
e
a
n
c
o
a
r
s
e
-
g
r
a
i
n
e
d
s
o
i
l
,
v
i
b
r
a
t
i
o
n
f
r
e
q
u
e
n
c
y
s
h
o
u
l
d
b
e
n
o
l
e
s
s
t
h
a
n
1
,
6
0
0
c
y
c
l
e
s
p
e
r
m
i
n
u
t
e
.
V
i
b
r
a
t
i
n
g
p
a
d
s
o
r
p
l
a
t
e
s
a
r
e
a
v
a
i
l
-
a
b
l
e
,
h
a
n
d
-
p
r
o
p
e
l
l
e
d
,
s
i
n
g
l
e
o
r
i
n
g
a
n
g
s
w
i
t
h
w
i
d
t
h
o
f
c
o
v
e
r
a
g
e
f
r
o
m
1
-
t
o
1
5
f
t
.
V
a
r
i
o
u
s
t
y
p
e
s
o
f
v
i
b
r
a
t
i
n
g
-
d
r
u
m
e
q
u
i
p
m
e
n
t
s
h
o
u
l
d
b
e
c
o
n
s
i
d
e
r
e
d
f
o
r
c
o
m
p
a
c
t
i
o
n
i
n
l
a
r
g
e
a
r
e
a
s
.
C
r
a
w
l
e
r
T
r
a
c
t
o
r
B
e
s
t
s
u
i
t
e
d
f
o
r
c
o
a
r
s
e
-
g
r
a
i
n
e
d
s
o
i
l
s
w
i
t
h
l
e
s
s
t
h
a
n
4
t
o
8
p
e
r
c
e
n
t
p
a
s
s
i
n
g
N
o
.
2
0
0
s
i
e
v
e
,
p
l
a
c
e
d
t
h
o
r
o
u
g
h
l
y
w
e
t
.
6
t
o
1
0
3
t
o
4
c
o
v
e
r
a
g
e
s
V
e
h
i
c
l
e
w
i
t
h
s
t
a
n
d
a
r
d
t
r
a
c
k
s
h
a
v
i
n
g
c
o
n
t
a
c
t
p
r
e
s
s
u
r
e
n
o
t
l
e
s
s
t
h
a
n
1
0
p
s
i
.
T
r
a
c
t
o
r
w
e
i
g
h
t
u
p
t
o
8
5
t
o
n
s
.
P
o
w
e
r
T
a
m
p
e
r
o
r
R
a
m
m
e
r
F
o
r
d
i
f
f
c
u
l
t
a
c
c
e
s
s
a
n
d
t
r
e
n
c
h
b
a
c
k
f
l
l
.
S
u
i
t
a
b
l
e
f
o
r
a
l
l
i
n
o
r
g
a
n
i
c
s
o
i
l
s
.
4
t
o
6
f
o
r
s
i
l
t
o
r
c
l
a
y
;
6
f
o
r
c
o
a
r
s
e
-
g
r
a
i
n
e
d
s
o
i
l
s
2
c
o
v
e
r
a
g
e
s
M
i
n
i
m
u
m
w
e
i
g
h
t
3
0
l
b
.
C
o
n
s
i
d
e
r
a
b
l
e
r
a
n
g
e
i
s
t
o
l
e
r
a
b
l
e
,
d
e
p
e
n
d
i
n
g
o
n
m
a
t
e
r
i
a
l
s
a
n
d
c
o
n
d
i
t
i
o
n
s
.
W
e
i
g
h
t
s
u
p
t
o
2
5
0
l
b
s
,
f
o
o
t
d
i
a
m
e
t
e
r
4
t
o
1
0
i
n
.
(
A
D
A
P
T
E
D
F
R
O
M
D
O
D
,
2
0
0
5
)
T
A
B
L
E
1
1
.
5
S
U
G
G
E
S
T
E
D
C
O
M
P
A
C
T
I
O
N
E
Q
U
I
P
M
E
N
T
A
N
D
M
E
T
H
O
D
S
(
C
o
n
t
i
n
u
e
d
)
< PREVIOUS VIEW
11-32
Chapter 11
MAY 2009
percent of the maximum dry density. For a refuse pile, MSHA regulations do not permit lif thick-
nesses in excess of 2 feet unless an adequate factor of safety has been demonstrated. However, lif
thicknesses of two feet (or greater) may preclude uniform compaction. From a technical standpoint,
coarse refuse at non-impounding facilities that contains predominantly large particles similar to rock
fll can be placed and compacted in larger lifs, if adequately-sized compaction equipment is used.
Since coarse coal refuse typically consists of both rock and soil particles, one to two-foot lifs can typi-
cally be used if suitable compaction equipment is employed.
Appropriately-sized equipment does not guarantee that placement and compaction of refuse will be
as desired. The coordination of equipment operations and sequencing of refuse placement greatly
afects the level and quality of compaction achieved. Refuse should be placed in piles on the working
surface and spaced so as to allow spreading equipment to achieve the specifed lif thickness with
minimal efort prior to compaction. Refuse should be spread away from a conveyor discharge point
to the specifed lif thickness. A sequence or progression of refuse placement should be established.
When handling equipment of sufcient number and size is employed using an efcient operating
system, refuse should be evenly distributed on the working surface and can be spread in relatively
thin horizontal lifs, allowing compaction equipment to perform under favorable conditions.
To achieve consistent and uniform compaction, successive lifs should be knited together by scari-
fying smooth compacted surfaces prior to placement of subsequent lifs. This is particularly impor-
tant where concentrated haul trafc has resulted in additional breakdown of the material or where
smooth-drum rollers are used. Generally, litle scarifying is necessary where padded or sheepsfoot
rollers are employed unless the working surface has been dormant for a long period of time or where
concentrated haul trafc has occurred. A rock ripping atachment for a tractor dozer may be required
to scarify the working surface properly.
Where an intermediate crest elevation has existed for a period of time, the surface may become highly
compacted from the combination of trafc and breakdown of the surface material due to weathering.
Additionally, freezing conditions can result in frost heave and formation of ice lenses. If this condi-
tion is not addressed when the next lif is placed, then a layer of difering hydraulic conductivity may
cause seepage to run horizontally and exit at the face of the downstream slope. Such a condition,
which will lead to concerns about seepage and stability, can be addressed by scarifying or removing
the top surface material prior to placing the next lif.
Special care must be taken to achieve adequate compaction at the sloped edges of each lif where,
due to the lack of natural confnement, the refuse tends to move away from the equipment without
densifcation. Although the total stability of an embankment is not signifcantly afected by refuse
density at the slope face, loose material is susceptible to erosion and creep. To achieve adequate sur-
face density, compaction should extend several feet down the slope. If this is not possible, the loose
material should be shaved of with a dozer or excavator and pushed back onto the working surface
for compaction.
When structural fll is placed against a hillside or against an existing embankment, the existing mate-
rial should be keyed or benched. This can be accomplished by using a dozer to cut a sufcient dis-
tance into the slope (e.g., approximately three or four feet horizontally where the terrain is steep) as
the new refuse embankment is advanced in height. Figure 11.3 illustrates the process of benching to
tie structural fll into a natural slope. This process removes surface material that may not be at the
required density, permits compaction at the construction interface, and reduces the tendency for a
natural slip surface to develop at this critical location.
Compaction tests, as discussed in Section 11.8, should be performed relatively ofen during initial
structural fll refuse construction when material characteristics and equipment efciency are being
< PREVIOUS VIEW
11-33
Construction and Disposal Operations
MAY 2009
evaluated. Thereafer, the testing frequency can be reduced, provided that a history of successful
equipment performance and density testing indicates that the desired compaction is being accom-
plished. A typical criterion for the frequency of compaction testing for embankment fll is: at least
one density test per 2,000 cubic yards of material placed or one test for every lif placed, whichever is
greater (USBR, 1998). Tests should also be performed when it is suspected that adequate compaction
is not being achieved.
11.5.2 Upstream Construction Implementation Procedure
Incidents have occurred while establishing pushouts where bulldozers have sunk into the underly-
ing fnes and not been recovered. Even in situations where the fne refuse surface is dry, dessicated
and vegetated, the underlying materials may remain sof and can exhibit sudden shear failure under
equipment operation.
As indicated previously, upstream construction of an impounding embankment poses technical and
construction challenges. Since upstream construction involves the development of an embankment over
saturated, unconsolidated fne coal refuse, the stability of the embankment under both static and dynamic
loading must be carefully evaluated. The potential for sudden failure of the pushout surface and the
underlying fne refuse during upstream construction requires that special techniques and equipment be
used and that the work be performed by experienced and properly trained equipment operators.
Placement of material on top of saturated, hydraulically-placed fne coal refuse results in the com-
pression of fne materials. Since the material is initially loose and saturated, as the particles move to
a closer packing, the water in the pores is placed under pressure. This excess pore-water pressure
reduces the efective stress and can make the material unstable. Fill material must be placed on top of
the fnes at a slow enough rate that the pore-water pressure can dissipate without causing signifcant
instability. If too thick a layer is placed too quickly, instability will occur that can adversely afect both
the immediate safety of the equipment operators and the overall safety of the embankment.
It is important that equipment operators understand the potential for instability during upstream
construction and the general concept that the rate of material placement during upstream construc-
tion must be controlled. Placing material thicker and/or faster on hydraulically-placed fnes can be
detrimental. When excess pore-water pressures are created in one area, construction activity should
be moved to another area to allow pore-water pressures to dissipate.
Task-specifc training should be provided to the equipment operators that will be performing
upstream pushout embankment construction and to workers who will be in the vicinity of the
3-4 FT
REFUSE EMBANKMENT
DOZER CUT FOR BENCHING
FILL INTO NATURAL SLOPE
NATURAL SLOPE
FIGURE 11.3 BENCHING OF EMBANKMENT FILL INTO NATURAL SLOPE
FIGURE 11.3 BENCHING OF EMBANKMENT FILL INTO NATURAL SLOPE
FIGURE 11.3 BENCHING OF EMBANKMENT FILL INTO NATURAL SLOPE
< PREVIOUS VIEW
11-34
Chapter 11
MAY 2009
upstream construction. The training should familiarize operators and workers with the risks asso-
ciated with upstream construction and should include specifc instructions for developing access
to pushout areas, along with specifc construction methods for performing upstream construction.
Information describing the risks associated with upstream construction and features that are indica-
tive of unstable working surfaces should be provided as part of the training program. Records docu-
menting the training should be kept.
Certain precautions are essential for minimizing the potential for failure of coarse coal refuse placed
over fne refuse. These precautions will also help to minimize the potential for occurrence of acci-
dents associated with the upstream construction activities. These precautions include the following:
Impoundment construction and discharge of fne refuse should be managed such
that a sufcient fne refuse delta on which to initiate upstream construction is creat-
ed. The delta should be as uniform as is practical, which can be facilitated by routine-
ly moving the slurry discharge point along the upstream slope of the embankment.
The normal pool elevation should be lowered via pumping or other means (if practi-
cal) to the lowest practical level and away from the fnes delta.
A bufer should be established from the edge of proposed pushout where high-
ground-pressure vehicles such as haul trucks are excluded from travel. The width
of the bufer should be established based on site-specifc conditions and equipment
(e.g., 50 feet has been satisfactorily used).
Only low-ground-pressure equipment should be used to perform upstream con-
struction within the bufer area.
Two-way radio communication for equipment operators and mine personnel should
be provided during upstream construction activities, and it is recommended that
equipment operations during the initial pushout be within sight of mine personnel
and that operators be provided with foatation devices (e.g., life jackets).
Work should only be performed during daylight hours until a stable working surface
is established.
The placement of coarse refuse for upstream construction should initiate with ad-
vancement of a thick layer (typically 4 to 6 feet thick) of coarse coal refuse onto the
exposed fne refuse delta. Placement of the initial lif of coarse refuse should begin
along the embankment upstream slope and gradually advance upstream over the
fne refuse. A portion of the advancing lif may sink into the fne refuse in sof areas
or areas where the surface is saturated. It may be possible to minimize this efect by
reducing the lif thickness or lowering the impoundment water level.
Equipment working near the upstream edge of the pushout should be oriented per-
pendicular to the face of the active edge (i.e., no equipment should travel near and
parallel to the upstream edge of the pushout).
Pushout construction should be sequenced so that haul trucks do not travel adjacent
to pushout areas until a stable working surface is established.
Pushouts should be constructed utilizing a bufer consisting of at least one pile of
coarse refuse. The bufer pile of coarse refuse should remain between the dozer and
impoundment as the refuse is pushed onto the fne refuse delta. Use of this method
will always keep the dozer in a safer position away from the edge of the fne refuse.
Pushouts should be performed perpendicular to the upstream face of the embank-
ment and/or impoundment and should be limited to a prescribed length onto the
delta (e.g., 25 feet measured from the upstream edge of the embankment or stable
working surface). It is recommended that the initial lif for upstream construction be
< PREVIOUS VIEW
11-35
Construction and Disposal Operations
MAY 2009
spread to a width of at least two times the push out length (e.g., 2 times 25 feet or 50
feet) before further advance of the lif upstream over the delta.
The surface of the upstream pushout embankment should be graded to drain toward
the impoundment.
Monitoring and inspection of the upstream construction area should be performed by a qualifed
person who is familiar with upstream construction methods and risks. Prior to and during initial
pushout construction, this person should inspect the refuse embankment and the area of the upstream
construction. The inspection should focus on identifying conditions that could afect the safety of the
equipment operators as well as conditions that could afect the safety of the embankment. These
include the following conditions:
Development of cracks with vertical displacement or scarps in the vicinity of the
pushout (the orientation and shape of the cracks may indicate shearing rather than
diferential setlement)
Excessive pumping of the pushout surface
Excessive bulging of the fne refuse delta (e.g., bulge or displacement height in excess
of the pushout lif level) where work is being performed
A situation posing a threat that the embankment could be overtopped by water or slurry
Sudden or major subsidence of the embankment crest
Longitudinal or transverse cracking of the embankment crest
Major sliding/failure of upstream or downstream embankment slopes or abutment
slopes adjacent to the embankments
Unusual seepage from areas of the downstream face or from the toe of the embankment
Unusual conditions on the embankment downstream slopes that develop during
upstream construction
Signifcant landslides within the impoundment area
In addition to the above conditions, embankment piezometers should be monitored before and during
the upstream construction. Where development of signifcant pore pressures are or remain a concern for
the initial pushout, new piezometers can be installed within the fne refuse to aid in monitoring of the
upstream construction process. The location for piezometers should refect the potential interference
from construction activities and the likely displacements associated with upstream construction.
If any of the above listed conditions are observed, or if piezometers indicate unacceptable levels of
pore pressure, the information should be reported to the Engineer and Operator, and equipment and
personnel should be moved to another work area until the cause of the problem is identifed and cor-
rected. The results of the inspection should be documented.
11.5.3 Excess Coarse Refuse/Inclement Weather Disposal
Some coal preparation plants may generate more coarse refuse than required for impounding
embankment construction (excess coarse refuse), with the result that specifc embankment zones
or separate embankments are designed with diferent material placement and compaction crite-
ria (e.g., thicker lifs). This situation may also result in the designation of a location for inclement
weather disposal, when compaction to the normal embankment specifcations is precluded due to
moisture or freezing conditions.
To mitigate concerns for combustion, excess coarse refuse should generally be spread in layers or
lifs less than 2 feet thick. Additionally, lif placement should result in a working surface capable
< PREVIOUS VIEW
11-36
Chapter 11
MAY 2009
of supporting equipment trafc associated with subsequent disposal operations. Other geotechni-
cal factors that must be considered before constructing embankment zones with thick lifs include
setlement and hydraulic conductivity.
Placement of refuse in thick lifs may also be acceptable at non-impounding embankments and in
surface mine backstack areas where such materials do not infuence the stability and drainage control
of the site. Note that 30 CFR 77.215 addresses construction requirements for refuse piles. Section
77.215(h) requires that refuse piles be constructed in compacted layers not exceeding 2 feet in thick-
ness except that the MSHA District Manager may approve construction of a refuse pile in compacted
layers exceeding 2 feet in thickness where engineering data substantiate that a minimum static safety
factor of 1.5 will be atained.
Sealing of any disposal surface that will be exposed for a long period of time is always important.
Sealing requires use of smooth-wheeled vehicles (hauling units or smooth-drummed compaction
equipment) rather than tamping-foot rollers, with appropriate grades provided for drainage. For
fnal surfaces, preparation for abandonment and revegetation should follow procedures discussed
in Chapter 10.
11.5.4 Disposal of Fine Refuse Slurry
The fne coal refuse slurry disposal system is a distinct part of a coal refuse disposal facility. Slurry
pipe size, composition, anchor/bracing details, discharge point locations and construction methods
should be included in facility plans and specifcations for implementation as part of overall refuse
disposal operations. Disposal of fne coal refuse slurry is discussed here with respect to site-related
planning, scheduling and the relationship to other construction and disposal operations. If possible,
the disposal of fne refuse slurry should result in a fne coal refuse delta above the impoundment pool
to: (1) provide the best possible foundation for upstream advancement of the embankment whether or
not the facility is initially planned to be developed by the upstream construction method, (2) protect
the embankment face from wave erosion by minimizing the impounded water depth at the upstream
embankment slope, and (3) reduce seepage into the embankment and keep the phreatic surface asso-
ciated with embankment seepage as low as possible (assuming the fne coal refuse hydraulic conduc-
tivity is lower than the embankment hydraulic conductivity). At some facilities clarifed water from
the impoundment is pumped to a separate cell, typically at the upstream end of the reservoir, further
isolating the pool from the embankment face.
As illustrated in Figure 11.4, slurry disposal functions most efectively if the discharge point of the
transport pipe is located at a relatively low elevation on the impoundment side (upstream slope)
of the embankment and is periodically moved to adjacent locations along the upstream slope. The
coarsest material in the slurry will be deposited in the immediate vicinity of the discharge point,
leading to maximum material strength near the embankment and maximum water depth near
the upstream end of the impoundment. Periodic movement of the discharge point back and forth
along the upstream edge of the embankment will create a more uniform delta deposit. The slurry
may be discharged at other locations to achieve slurry deposition at specifc features such as mine
barriers or to more fully utilize available impoundment capacity. The refuse disposal plan should
discuss the type and size of equipment required to move the slurry discharge pipe around the
impoundment.
Pipes transporting the slurry to the impoundment should not be placed through the embankment
unless seepage and structural provisions are provided. Pipes through embankments tend to be nat-
ural paths for seepage and could result in embankment failure from internal erosion. Also, pipes
installed within the embankment or along the downstream face of the embankment could cause ero-
sion and are an environmental hazard if they leak or fail (Figure 11.4a).
< PREVIOUS VIEW
11-37
Construction and Disposal Operations
MAY 2009
11.4a INCORRECT SLURRY PIPE INSTALLATION
FIGURE 11.4 SLURRY DISPOSAL
SLURRY
IMPOUNDMENT
NOTE: DISCHARGE SOLELY FROM THE UPPER
END IS GENERALLY NOT RECOMMENDED
UNLESS THE FACILITY IS DESIGNED TO
ACCOMMODATE IT.
11.4b CORRECT SLURRY PIPE INSTALLATION
SLURRY
IMPOUNDMENT
NOTE: DISCHARGE FROM THE UPPER END OF THE
IMPOUNDMENT SHOULD GENERALLY ONLY
BE IN CONJUNCTION WITH OTHER DISCHARGES
UNLESS THE FACILITY IS DESIGNED TO
ACCOMMODATE IT.
FIGURE 11.4 SLURRY DISPOSAL
FIGURE 11.4 SLURRY DISPOSAL
< PREVIOUS VIEW
11-38
Chapter 11
MAY 2009
11.5.5 Disposal of Combined Refuse
As discussed in Chapter 2, fne coal refuse slurry can be processed by dewatering with a vacuum
flter, belt flter press or centrifuge. A flter cake of fnes that in some instances is at a water content
too high for easy transport and disposal as a solid waste material can result, particularly during
inclement weather periods. Under such adverse conditions, the flter cake may be difcult to confne
in haulage units. To minimize spillage and resulting degradation of haul roads, conventional trucks
fted with sealing tailgates may be necessary. As a separate waste material, flter cake typically cant
be spread in lifs and compacted and will not support construction equipment.
To resolve this materials transport and disposal problem, the flter cake can be mixed with coarse
refuse at the preparation plant, resulting in a combined refuse that normally can be transported in
conventional trucks or by conveyor. However, watertightness of transport unit bodies is desirable,
because the material is typically very wet.
Combined refuse disposal facilities are typically designed as non-impounding embankments or
refuse piles. In some instances where the combined refuse is very wet, the material may not be suit-
able for normal embankment construction. A large area may be needed for spreading and drying the
material. During wet or winter periods, movement of spreading or compacting equipment over this
area may be difcult or impossible. This can lead to operational inefciencies and larger disposal
areas than normally required. If confned disposal is necessary, a structural zone or embankment
may be needed for retaining the combined refuse. Such embankments must be constructed from bor-
rowed soil and rock materials, or some of the coarse refuse must be diverted for this purpose.
The mine operator or designer should carefully evaluate the efects of combined refuse disposal
prior to implementation of this type of total disposal system so that all relevant requirements and
limitations are factored into the planning. The procedure for disposing of coarse refuse and dewa-
tered fne refuse flter cake must be planned and designed specifcally for conditions found at the
disposal site.
11.5.6 Use of Amendments
Coal refuse may require amendments for neutralization and stabilization. The relative quantities
and methods for combining refuse and materials such as lime, combustion ash, and kiln dust should
be detailed in the facility plans, specifcations and the operation and maintenance plan. Rich and
Hutchinson (1990) discuss the operational implications of using lime kiln dust to neutralize and
stabilize combined refuse at a West Virginia mine site using the following practices: (1) mixing the
kiln dust with flter cake at the preparation plant to absorb moisture, (2) refuse cell development
with elevated roadways constructed with rock for equipment access for dumping of the combined
refuse, and (3) methods to shed excess water by separating the dumped piles to allow drainage
prior to grading, and (4) maintaining drainage control within the refuse cell. For the cited case, an
application of 2 percent by weight of kiln dust provided sufcient improvement in the combined
refuse to enable efective disposal. Section 10.3.2.3 provides additional discussion of amendments.
11.6 RELATED ACTIVITIES
Many activities besides refuse disposal are involved in the development of a disposal facility. Activities
associated with coal refuse disposal include:
Construction and maintenance of haul or access roads
Construction of liner systems prior to refuse placement
Development of underdrain and spring collection systems
< PREVIOUS VIEW
11-39
Construction and Disposal Operations
MAY 2009
Foundation preparation prior to disposal of coal refuse
Placement of earth borrow material as structural fll and impervious cores and/or
seepage control blankets
Control of water while constructing facility appurtenances and developing the em-
bankment foundation
Construction of internal drainage systems within the embankment
Excavation for spillways, ditches, drainage installations, repair work, etc.
Construction of embankments for mine seals and barriers (e.g., flls for protecting
impoundments from the infuence of potential underground mine subsidence)
In general, construction procedures for these related activities are the same in mining as for heavy
construction. Therefore, discussion of these topics herein is limited and is related specifcally to coal
refuse disposal operations.
The activities listed above and the construction of appurtenant structures subsequently discussed
in Section 11.7 are similar in the sense that neither pertain directly to the primary facility purpose
of refuse disposal, yet both are needed for refuse disposal to proceed. The distinction made herein
is: (1) the above activities will generally be performed with the same equipment used for refuse dis-
posal, and the materials required are basically available at the site, and (2) the structures discussed
in Section 11.7 generally require special equipment and procedures, and the materials involved must
be purchased.
11.6.1 Haul and Access Road Construction and Maintenance
Haul and access roads are an integral part of a site refuse transport system and have been a source of
safety concerns. Surfacing, width, signage, runaway vehicle provisions, drainage control and berms
are important design issues. Roadway layout, design and maintenance impacts the operations of
the refuse disposal facility. Haul and access road locations must be determined in the initial stages
of planning and design of the refuse disposal facility. Consideration must also be given to tempo-
rary haul roads required during various stages of facility construction for safe access to the working
surface of the embankment. Because of the importance of disposal facility haul roads, MSHA (1999)
published The Haul Road Inspection Handbook, which focuses on the safety aspects of haul road design.
Guidance for the layout and design of haul and access roads is also provided in the U.S. Bureau of
Mines publication, Design of Surface Mine Haulage Roads A Manual (Kaufman and Ault, 1977), and
other references.
11.6.2 Liner Systems
State regulations require protection of the groundwater. This may in some states necessitate the
installation of a liner system to reduce the potential for impacts to the environment. The extent
and type of liner system required for groundwater protection generally depends upon the envi-
ronmental seting of the refuse disposal site and the potential acidity of the refuse. Liner systems
are normally constructed of clay soils from on-site borrow areas or from geosynthetic materials.
Geosynthetic liners may consist of a geomembrane or a clay-impregnated geotextile (geosynthetic
clay liner or GCL). Clay soil liners are used when the acid potential of the refuse is low to moder-
ate, and geosynthetic clay liners may be considered for higher acid potential refuse or when clay
is unavailable at the site. Liner systems are normally installed afer foundation preparations have
taken place and prior to the placement of refuse. Extensive underdrain and spring collection sys-
tems are sometimes installed beneath liner systems to prevent sloughing and/or slides that could
compromise liner integrity. Because of the potential for a liner to introduce a slip plane, the efect of
a liner system must be evaluated in the stability analyses. Section 10.4 provides additional discus-
sion of liner systems.
< PREVIOUS VIEW
11-40
Chapter 11
MAY 2009
11.6.3 Embankment Foundation Preparation
As discussed in Section 6.3, the foundation of a coal refuse embankment will afect embankment
stability, setlement and seepage. To achieve stability or to minimize potential setlement, removal of
sof cohesive materials may be desirable. However, for seepage and leachate control, these materi-
als can provide an efective seal if lef in place. Thus, cost-efective foundation design for a disposal
facility embankment requires that important decisions be made relative to removal or use of on-site
materials.
11.6.3.1 Clearing and Topsoil Stockpiling
Cuting and removal of trees, brush and other vegetation within the footprint of the embankment
should be specifed. Vegetative mater, if not removed, can decay and cause setlement and the forma-
tion of slip planes. Vegetative mater may also contribute to spontaneous combustion and burning.
Additionally, foating trees and branches can plug hydraulic structures, particularly culverts and
decants. Cleared material should be removed from the construction area or burned.
Topsoil present in areas where refuse will be placed should be removed and stockpiled for reclama-
tion. Stockpiling should be performed in a controlled fashion in areas away from natural drainage
courses and areas of planned future development. State mining regulations typically provide guid-
ance for topsoil removal and stockpile requirements.
Within impoundment areas, trees and heavy brush should be cleared and removed so that foating
debris that could impact operation of spillways and decants is not present. Soil and topsoil should be
stockpiled for future use, as required. Clearing and removal of soils within the impoundment area
should be sequenced so as to minimize excessive disturbance to hillsides that could cause erosion
and lead to potential slope instabilities. Typically, clearing should be limited to areas that will poten-
tially be afected by the impoundment within one or two years.
11.6.3.2 Soft Soil Foundations
An efort should be made during the geotechnical subsurface exploration program to identify sof
soil foundation areas. If sof soils are present where a refuse embankment is planned, removal of the
sof materials may be necessary. Such removal can generally be accomplished with normal excavating
and hauling equipment. Temporary access roads for construction equipment should be constructed
as needed to facilitate safe operations. Depressions associated with removal of sof soils can be back-
flled with either on-site borrow material or coarse refuse, as available. Some types of sof soils may
be suitable for clay liner construction or for use as fnal reclamation cover, in which case they should
be stockpiled on the site and reused.
11.6.3.3 Competent Soil Foundations
If generally competent foundation materials are present in the area where embankment construction
is planned, it may be possible to utilize these in-situ materials. Some over-excavation and recompac-
tion of the material may be required if the soil type is acceptable but the in-situ density is too low. Prior
to such work, the area should be cleared of vegetation and organic topsoil. Clearing and excavation/
recompaction should not be performed very far in advance of refuse disposal since prolonged expo-
sure of the foundation soil could lead to weathering and deterioration of important physical proper-
ties and/or cause environmental damage. Construction of temporary access roads and stockpiling of
topsoil or other recoverable material should be performed in the manner previously described.
11.6.3.4 Rock Foundations
From the standpoint of strength, rock is usually an adequate foundation material for a coal refuse
embankment. However, an evaluation of foundation rock conditions should be performed, since trou-
< PREVIOUS VIEW
11-41
Construction and Disposal Operations
MAY 2009
blesome layering or discontinuities and localized weaknesses may be present. Seepage and uncon-
trolled discharge of leachates through fractures in foundation or abutment rock can be a signifcant
problem, especially for impounding facilities. Sections 6.6.5 and 8.9 discuss methods for preparing
rock foundations and abutments for construction of embankments.
Not all rock characteristics may be apparent following the geotechnical exploratory drilling and labo-
ratory testing performed as part of facility design. Thus, it is highly desirable that the foundation
surface be inspected by a geologist or geotechnical engineer afer the foundation has been exposed
by excavation. This will allow fnal confrmation of the embankment design prior to construction. If
special provisions associated with rock foundations are needed, they will typically involve control
features such as impervious soil blankets, localized grouting, shotcrete placement, or preparation of
a planned contact between clayey soils and rock. Section 11.6.6.1 discusses the placement of impervi-
ous blankets. Foundation grouting of rock formations to cut of or minimize seepage through frac-
tures is discussed in Fell et al. (2005) and USACE (1984a).
11.6.3.5 Mine Spoil
Mine spoil may be present in the foundation of the area planned for coal refuse disposal, either as
remnant materials from past mine operations at the site or as a result of planned surface mining and
refuse disposal. Evaluation of the mine spoil characteristics and associated foundation preparation
requirements should be addressed in the refuse disposal facility design. Characterization of the mine
spoil properties and the potential for use of mine spoil in construction along with the possible need
for removal, densifcation, and seepage/internal erosion control should also be addressed. Section 8.8
presents design and construction considerations associated with mine spoil. State regulations may
limit the use of mine spoil at a refuse disposal facility.
If unanticipated spoil materials are encountered, they should be thoroughly characterized to determine if
they are suitable for incorporation into the foundation at an impounding coal refuse disposal facility.
11.6.4 Water Control
Control of water in relation to the development of coal refuse disposal facilities falls into four basic
categories:
1. Control of food water by storage capacity combined with spillway and decant
structures for impoundments and by diversion ditches for non-impounding embank-
ments
2. Control of seepage through and under the refuse embankment (from springs/
groundwater and from impounded water)
3. Control of natural stream fow and storm runof during construction when perma-
nent facility hydraulic structures are not completed
4. Control of site drainage from storm runof
The frst category of water control structures refers to hydraulic structures associated with convey-
ing watershed runof through or around the coal refuse disposal facility. Design requirements for
these structures are presented in Section 9.7, and construction requirements for these features are
discussed in Section 11.7.
The second category of water control structures refers to internal drains for collection and control of
springs in the foundation area, seepage from an impoundment, interception of leachates, and control
of the phreatic level in the embankment. Design requirements for flters and drains are presented in
Section 6.6.2.3.1, and construction requirements are discussed in Section 11.7.
< PREVIOUS VIEW
11-42
Chapter 11
MAY 2009
The third category of water control is ofen referred to as a diversion and may be required during
facility construction. A diversion can be as simple as installing a small drainage culvert under a
haul road or as complex as construction for passing a stream around a site prior to completion of
permanent facility hydraulic structures. For coal refuse disposal, diversions may be required on
several occasions during the operational period of the facility as, for example, the perimeter ditch
system at a valley fll is upgraded to match the advancing refuse embankment. Temporary diver-
sions should be designed and constructed consistent with the risk associated with failure during
the time of use (i.e., the shorter the period of time required for diversion, the smaller the design
storm requirement, consistent with regulatory requirements). Permanent diversion ditches should
be designed as long-term drainage channels. Typical design criteria for stream diversions are pre-
sented in Chapter 9.
Diversion requirements discussed in Chapter 9 include:
Inlet and outlet sections sized to handle expected fows and designed without ob-
structions or sharp angles
Transport section alignment and grade
Freeboard design for preventing overfow to areas that must be protected
Design for prevention of erosive damage through material selection and energy dis-
sipation control
Planning, scheduling and construction of diversions should refect requirements for continued refuse
disposal and other site related activities. Access across open diversions may be difcult, especially
where permanent stream fows are being conveyed. Light-gauge, temporary culverts may be capa-
ble of passing the required fow. However, heavy equipment passing over such culverts can cause
damage if they are inadequately designed and constructed.
Although material selection for temporary diversions is not as important as for permanent hydraulic
structures, strength, erosion and corrosion issues must be considered. Further discussion of these
topics is provided in Section 9.7.3 and Section 11.7.
The fourth category associated with water control is related to on-site drainage for handling storm
runof. This type of water control generally includes grading of embankment surfaces and con-
struction of conveyance and collection structures. Temporary and fnal embankment surfaces,
including the active work area for refuse placement, must be graded such that runof is directed
toward conveyance channels and that ponding and saturation of the fll is minimized. This is typi-
cally accomplished by specifying surface grades of at least 1 percent and controlling the placement
of lifs so as to maintain a positive gradient toward the perimeter of the embankment where the
runof is collected and conveyed from the area in channels. The channels then convey site runof
to sedimentation traps or ponds for clarifcation before release. These drainage control features
should be incorporated into the facility plans and specifcations. The design of site drainage struc-
tures is addressed in Chapter 9.
A critical element in control of on-site drainage is the actual process of coarse refuse placement. Lif
placement (typically end-dumped piles placed by trucks) and spreading to a uniform thickness for
compaction is dependent upon a working surface that can support construction trafc and is not
excessively wet or does not have standing water that would interfere with the compaction efort. If
each lif is advanced uniformly across the entire active working surface, the surface gradient to drain-
age channels is maintained. Also, when periods of sustained inclement weather occur that adversely
impact the working surface, it is useful to grade and seal the working surface so that it sheds water
and to have other non-critical areas for short-term disposal.
< PREVIOUS VIEW
11-43
Construction and Disposal Operations
MAY 2009
Mobile conveyor systems are sometimes employed to deliver coarse refuse directly to the working
surface, with dozers, loaders and trucks used to spread the lif materials. Mobile conveyor locations
and equipment selection should refect the need for advancing lifs uniformly across the working
surface to maintain the drainage gradient.
11.6.5 General Excavation
Excavation is required for construction of diversions, minor drainage ditches and channels, trenches
for internal drains and control of natural springs, major decants or spillways, embankment founda-
tion development, liner construction, haul and access road construction, and other site operations.
Many of these earth-moving activities have already been discussed. However, certain precautions
should be taken for excavations at coal refuse disposal sites.
Many coal refuse disposal facilities are located in mountainous areas. It follows that some exca-
vations associated with facility development at these sites will result in steep, side-hill cuts. In
making such excavations, care should be taken to disturb as litle vegetation as possible, particu-
larly on the uphill side of drainage ditches. This is especially important if the cut being made is
for the construction of a permanent drainage channel. The stability of slopes above drainage chan-
nels is crucial to their performance during storms. Vegetation, especially in the form of trees with
extensive root systems, helps to maintain slope stability of steep natural hillsides and also aids in
minimizing the potential for erosive movement of soil into the channel and consequent blocking
or reduction of fow capacity.
Stability of excavated slopes, whether the excavation is permanent or temporary, can be achieved by
decreasing the angle of the cut or by bracing the exposed surface. Only if excavations are in sound,
competent rock should vertical cuts be made. Steeply-sloped excavations in clayey soils may appear
stable at the time of excavation, but sudden collapse, especially during rainy weather, can occur.
Safety regulations such as those developed by OSHA for providing safe conditions in and around
open excavations should be strictly followed. Slope stability issues are discussed in more detail in
Section 6.6.4.
11.6.6 Embankment Fill Constructed from On-Site Borrow Materials
Starter embankments for impounding facilities are frequently constructed from on-site borrow mate-
rials. At some coal refuse disposal facilities, coarse coal refuse may not be available in sufcient
quantities or with adequate properties for all embankment fll requirements. Combined coarse and
fne refuse may be difcult to compact to the density and strength required for structural fll. Also, as
discussed in Chapter 6, an impervious earth core or cutof may be needed within the refuse embank-
ment or as an upstream blanket for controlling seepage and water pressure. These considerations may
necessitate construction of a zoned embankment with design implications as discussed in Sections
5.1 and 6.3.1.2. Borrow materials may also be needed for flls and embankments associated with mine
seals and barriers. For any of these situations, use of on-site sources may be necessary. While borrow
materials generally can be characterized as clayey and silty or rock and granular, mine spoil
materials may exhibit both types of characteristics.
11.6.6.1 Clayey and Silty Borrow Materials
If fne-grained borrow materials such as clays or silts are to be placed in zones as part of embank-
ments or for seepage control (impervious blankets and cores), major considerations during construc-
tion include:
Material properties such as grain-size distribution, water content, densities, strength,
hydraulic conductivity, etc.
< PREVIOUS VIEW
11-44
Chapter 11
MAY 2009
Material availability in sufcient quantities, preferably located near the construction site
Equipment required for loading, transporting and placing the material
Lif thickness during placement
Equipment required for compaction and appropriate compaction procedures
Detailed discussions of earth-fll operations can be found in Sherard et al. (1963), Church (1981),
USACE (1995b) and USBR (1998). Some general comments regarding the distribution and placement
of clayey and silty borrow materials (for structural fll and impervious blankets and cores) are:
The material may be difcult to remove from the borrow areas. If scrapers are used
to excavate earth borrow of this type, an additional tractor may be required.
The material should be placed in relatively thin lifs. Sherard et al. (1963) recom-
mended approximately 6 to 9 inches. If thicker lifs are used, the required density
may not be achieved throughout the full depth of the lif.
Compaction equipment used for clayey or silty soils is typically of the segmented-
pad or tamping-foot type, commonly referred to as a sheepsfoot roller. Vibration
does not signifcantly improve the compaction of these soils.
The compactability of clayey and silty soil is very sensitive to moisture conditions.
In most regions, borrow areas should be graded to shed surface water, and fll areas
should be sloped to drain. In dry regions, moisture may have to be added to the
borrow material. Sealing of fll areas by rolling with smooth-wheeled vehicles or
non-vibrating, smooth-drum rollers is recommended if precipitation is anticipated.
USBR (1998) and Church (1981) provide discussions of embankment construction
and moisture control.
If the borrow material water content is too high to permit adequate compaction, dry-
ing may be required. Drying can sometimes be accomplished by disking and allow-
ing the sun and wind to remove the moisture.
Placement and compaction of fll for impervious blankets and cores in winter
weather is always difcult and is essentially impossible if the soil is frozen. Opera-
tions should be scheduled to avoid fll placement during periods of adverse weather
conditions.
The above discussion relates primarily to embankment zones where strength is a major factor. If
borrow material is being used for seepage control such as an impervious core within the embank-
ment or an imperious blanket in the valley botom, it is essential to obtain a uniform, non-layered
clayey or silty soil consistency. In such applications, the borrow material should be placed a few per-
centage points weter than optimum water content if the material can meet the associated strength
requirements and is in accordance with the construction specifcations. This should result in a homo-
geneous mass fexible enough to accept deformation without cracking. Further discussion of this
subject is provided in Section 6.5.3.
11.6.6.2 Rock and Granular Borrow Materials
Many considerations for use of rock or granular borrow material for structural fll are similar to those
mentioned above for clayey or silty borrow material. However, the handling and compaction equip-
ment needed, acceptable lif thicknesses, and material properties are diferent.
Rock-fll material is generally handled by large excavators and haul trucks, and granular borrow
material can be handled by scrapers if large boulders are not present. Quarrying hard rock may
< PREVIOUS VIEW
11-45
Construction and Disposal Operations
MAY 2009
require explosives and blasting, but removing sof weathered rock may only require a powerful
dozer with ripper teeth. Church (1981) discusses the various dozer atachments available for ripping
and their applications.
The wear on equipment used for handling rock and granular materials is signifcant. Steel tracks on
dozers require frequent replacement. Rubber tires are easily cut by sharp rock edges and require fre-
quent replacement, although replaceable chain guards can be used to reduce damage. Special armor
plating is needed for the bodies of trucks that haul rock and coarse granular materials.
Acceptable lif thicknesses for rock and granular material are much greater than for fne-grained soils,
provided that increased compactive efort is applied. USBR (1998), USBR (1987a) and Sherard et al.
(1963) recommend a lif thickness of one to two times the maximum rock diameter for rock flls that
contain fnes, while USACE (1995b) recommends that lif thicknesses for rockfll dams be no greater
that 24 inches unless test flls show that adequate compaction can be obtained using thicker lifs. This
guidance may require modifcation for mine spoil conditions where large rock on the order of the lif
thickness may inhibit densifcation of surrounding spoil or where uniform gradation or maximum
hydraulic conductivity characteristics are critical. In situations where the largest dimension of rock
in the fll approaches the desired lif thickness, the interstices around the rock should be flled with
fner materials and compacted until there is no visible evidence of consolidation of the material being
compacted. The recommended lif thickness for gravel is about one foot, but this can be increased
to two feet or more if the gravel is coarse. Vibratory compactors are very efective on rock and clean
granular fll and may allow even higher lif thicknesses. The type and size of vibratory compactors
required depends on the borrow material characteristics and lif thickness.
The water content for rock and granular fll is relatively unimportant. However, weting of rock and
clean gravels used for compacted fll will minimize friction between contact points of adjacent par-
ticles, allowing them to approach maximum density with reduced compactive efort.
Density controls for rock and granular flls are not as precisely defned as for fne-grained soils, and
use of a method specifcation (i.e., minimum number of passes by construction equipment) based on
site-specifc observations may be appropriate. It is relatively difcult to take accurate in-place density
tests for rock flls, as discussed in Section 11.8. Careful observation of the action of compaction equip-
ment is an important part of evaluation of the efectiveness of the placement and compaction process.
Key points to watch for include the following:
Compaction eforts should continue until no further decrease in the lif thickness is
observed during a pass of the compaction equipment.
When compaction is complete, loose fne particles on the surface near the compac-
tion equipment bounce due to vibrations transmited through dense fll.
When compaction is complete, the edges of adjacent rock particles will be crushed
from being wedged into a tight position. Further compactive efort will only increase
the amount of crushing rather than signifcantly increase the density. If the presence
of crushed rock particles reduces desired drainage, compaction should be closely
observed and terminated when crushing begins to occur.
11.6.6.3 Use of Mine Spoil
Mine spoil may be an acceptable borrow material for use in refuse disposal embankments, including
structural zones, provided that the material characteristics are identifed and accommodated in the
design and construction. Mine spoil may include substantial portions of fne silt and clay materials,
but typically is predominantly granular materials and large rock. Accordingly, in a zoned embank-
ment, more pervious mine spoil would be placed downstream of less pervious embankment mate-
< PREVIOUS VIEW
11-46
Chapter 11
MAY 2009
rials such as coarse refuse. Key issues are the consistency of the spoil borrow source material and
measures that need to be implemented so that it meets design requirements. Internal erosion of fner
material into mine spoil is a particular concern. Characterization, design, and construction consider-
ations of mine spoil are presented in Section 8.8.
Characterization of mine spoil can be accomplished by: (1) review of the mine overburden geology,
(2) use of exploration test pits and borings to obtain samples for testing, and (3) performance of geo-
physical surveys to determine density and shear-wave velocity. The durability of the spoil material
should be evaluated; in particular drainage and grain-size characteristics are important. Use of larger
specimens for laboratory testing to obtain a representative sample may be needed. Measures that
may be necessary during construction include segregation of large rock (using screens or in some
instances during lif placement using dozers and graders) and rigorous quality control testing to
verify material properties. Additionally, it may be desirable during mining operations that will pro-
vide borrow material, to strip and segregate soil and some rock strata (e.g., shales) from more durable
rock strata for subsequent use.
11.7 FACILITY APPURTENANT STRUCTURES
Facility appurtenant structures generally require materials and construction procedures that are
diferent from those for the disposal of refuse. These structures generally include, but are not lim-
ited to:
Internal drainage systems in the refuse embankment such as internal drains, drain-
age trenches, etc.
Spillways, decant structures and other major hydraulic control structures
Minor surface drainage structures such as collection ditches, erosion control devices
and culverts
Haul and access road bridges and miscellaneous buildings associated with refuse
disposal operations
Mine-opening/auger-hole seals and drains
Discussion of material requirements and construction procedures for various types of appurtenant
structures is provided in the following pages.
11.7.1 Material Requirements
The potentially corrosive nature of some coal refuse or leachates from the coal refuse makes material
selection important if the appurtenant structures are to function as intended for long periods and into
abandonment. Table 11.6 presents the most common construction materials used for facility appur-
tenant structures, shows the corrosive or deterioration mechanism for each material, and indicates
protective measures that may be taken.
Coal refuse, and particularly leachate water therefrom, should be chemically tested for corrosive
components. Typically, leachates will be acidic (low pH value) and will contain sulfates that cause
corrosion of many materials. Coal refuse varies signifcantly, and testing is advisable prior to identi-
fying appropriate protection measures at a particular disposal facility.
11.7.2 Filters and Drains
Control of water fowing across a disposal site, or seeping through an embankment or its foun-
dation, is an important factor in satisfying both safety and environmental requirements. Design
requirements for limiting, controlling and collecting these fows are presented in Section 6.6.2.3.
< PREVIOUS VIEW
11-47
Construction and Disposal Operations
MAY 2009
TABLE 11.6 EFFECTS OF CORROSION AND COUNTERMEASURES
Facility
Component
Materials and their Corrosive or
Deterioration Mechanism
Protective Measures
I. Internal Drainage Systems
A. Bank run gravels
and sands or
crushed rock
Sandstones: If the bonding agent between
sand grains is silt particles or clay fakes
consisting of calcite (CaCO
2
), sandstones will
be attacked by acidic leachates.
Use only corrosive resistant sandstone with
silica bond (SiO
2
).
Limestones: Mineral calcite (CaCO
3
) will be
dissolved by acidic leachates.
Do not use in internal drainage system.
Shale: Silt and clay composition breaks down
when subjected to weathering or moisture.
Do not use in internal drainage system.
B. Geosysnthetics
Nylon, Polypropylene, Polyester, HDPE, PVC:
Good corrosive resistance.
Most materials are sensitive to excessive
heat. Protection from ultraviolet light is
necessary prior to installation.
Steel and Corrugated Metal: See
discussion of metal products,
Item III below.
Galvanizing and/or coating with asphaltic
or epoxy protective material or coal
tar. Asphaltic and coal tar coatings are
susceptible to damage. Asphalt and coal
tar pipes are not recommended in critical
situations. Properly designed epoxy-coated
steel pipes can be considered for critical
situations.
Use thick-walled pipe to account for
corrosion; applicable if rate of corrosion is
relatively low.
Cathodic protection in vulnerable areas.
C. Drainage and
transport pipes
Use special alloy steels, stainless or other,
(See discussion of Metal Products under
Item III).
Cast Iron: Same as steel, but corrosion may
form protective coating on metal under ideal
conditions.
Coating with asphaltic protective material
or coal tar. Coating is susceptible to
damage and is not recommended in critical
situations.
Use thick-walled pipe to account for cor-
rosion; applicable if rate of corrosion is
relatively low.
Cathodic protection is required in
vulnerable areas.
Plastic (General): Excellent corrosion
resistance.
Consideration should be given to the
compatibility of plastic pipe materials with
bedding preparations and backfll.
< PREVIOUS VIEW
11-48
Chapter 11
MAY 2009
Facility
Component
Materials and their Corrosive or
Deterioration Mechanism
Protective Measures
Polyethylene (PE): Excellent for corrosive
liquid transport and backfll.
Check characteristics and strength for
the density PE pipe considered. Outside
diameter controlled HDPE, when properly
designed, can be used in critical situations.
ABS: Good corrosion resistance.
Check characteristics and strength of
ABS pipe.
C. Drainage and
transport pipes
(Continued)
PVC: Good for use with corrosive solutions
below 140F, brittle at low temperatures.
Do not use when temperature above
140F may be contacted. Check allowable
application of specifc PVC type used.
Concrete: See discussion on Concrete Structures, Item II below.
Thermosetting: Epoxy and polyester resin
reinforced (fberglass and flament wound)
and non-reinforced pipe.
Check properties of specifc resin used for
strength, temperature effects and cor rosion
resistance.
D. Gates and
miscellaneous
metal work
See discussion of Metal Products under Item III.
II. Concrete Structures (spillways, decant structures, channels, outfow conduits, etc.)
A. Cement
The bonding agent in ordinary concrete
(hydrated Portland cement) leaches when
exposed to acidic waters.
Use Type II or IIa cement for contact with
water containing (100 to 150 ppm) sulfates
as SO
4
.
Sulfates in water or refuse react with
tricalcium aluminate in cement causing
swelling and disintegration of concrete.
Use Type V cement for contact with water
containing greater than 1000 ppm of
sulfates as SO
4
.
Provide coatings for concrete structures as
discussed in ACI (2005) Committee 515
report.
Low water-cement mix ratios result in less
permeable concrete thus reducing pene-
tration of water-containing, deleterious
compounds. Air entrainment and other
additives are also recommended for
increasing resistance to water penetration
and improving workability of low-water
cement mixes.
Replace concrete materials with materials
that are less susceptible to deterioration
(e. g., coated cast-iron pipe in place of
concrete pipe).
TABLE 11.6 EFFECTS OF CORROSION AND COUNTERMEASURES
(Continued)
< PREVIOUS VIEW
11-49
Construction and Disposal Operations
MAY 2009
Facility
Component
Materials and their Corrosive or
Deterioration Mechanism
Protective Measures
B. Aggregate
See discussion of bank-run gravels or crushed rock, Item I.A.
See also ACI (2005), ACI Committee 201 for reactive aggregates.
C. Reinforcing
steel
See also discussion of metal products under Item III.
Porous concrete or inadequate cover
permits rapid attack on reinforcing steel.
Corroding steel causes expansion,
disintegrating the concrete.
Typical concrete cover may protect reinforcing
steel if corrosive environment is not severe.
For severe environments and long service life,
epoxy-coated rebar may be considered.
Provide coatings for concrete structures as
discussed in ACI (2005).
Low water-cement mix ratios result in less
permeable concrete and thus provide greater
protection against corrosion. Air-entrained
concrete is recommended as an aid to
placement and for increased resistance to
water penetration.
III. Metal Products (gates, valves, outfow conduits, decant pipes, etc.)
A. Steel
Electrochemical corrosion caused by
oxidation in an ionizing medium (water).
Rate of corrosion is governed by environ-
mental factors such as hydrogen ion
concentration, oxygen concentration, and
temperature. For pH between 6.5 and
8.0, corrosion protection is probably not
required.
Use internal and external coatings such
as asphalt, coal tar, epoxies or other non-
reactive materials.
Provide cathodic protection. Use sacrifcial
anodes to reduce corrosion.
Use thicker material in construction to account
for deterioration; applicable if rate of corrosion
as determined from bench or feld studies is
relatively low.
Use special-alloy steels, stainless or other.
Corrosion caused by reaction with
dissolved, ionized or colloidal substances in
the corroding medium (water) such as: H
+
,
0
2
, C0
2
, NH
3
(bacteria), S0
4
-2
, S
-2
, Cl.
Use of internal and external coatings such
as epoxy, asphalt, coal tar, epoxies or other
nonreactive materials.
Provide cathodic protection. Use sacrifcial
anodes to reduce corrosion of metal product
to be protected.
Use thicker material in construction to account
for deterioration; applicable if rate of corrosion
as determined from bench or feld studies is
relatively low.
Use special-alloy steels, stainless or other.
TABLE 11.6 EFFECTS OF CORROSION AND COUNTERMEASURES
(Continued)
< PREVIOUS VIEW
11-50
Chapter 11
MAY 2009
The material and construction issues associated with design implementation are discussed in the
following subsections.
11.7.2.1 Granular and Geotextile Filters
Filters consist of granular materials placed against disposed coal refuse or earth fll where seepage is
likely to exit, or between materials of greatly diferent particle size (e.g., between clayey silt and coarse
coal refuse) where seepage fows from the smaller-particle-sized material to the larger-particle-sized
material. The primary purpose of a flter is to allow movement of water without allowing fne par-
ticles to exit or to move into the coarser material void spaces. Such movement could lead to internal
erosion and weaken the fne material. Internal erosion can create voids in an embankment, increase
rates of seepage and cause eventual failure. Movement of fner particles can also obstruct the fow
within the coarse material. Granular flters may consist of multiple layers. The gradation of each layer
should be sized such that fner granular material is unable to enter coarser layers and cause clogging.
Geosynthetic materials such as geotextiles can be used in single layer applications or sometimes in
combination with granular materials to create multiple layers. Layered granular and/or geotextile
systems are commonly used in flter diaphragms and internal drains in slurry impoundment embank-
ments. The National Dam Safety Review Board (NDSRB), which comprises federal agencies that deal
with dam safety, recommends that a geotextile not be used as a flter for a critical internal drain in a
water impounding dam unless it is accessible in the event that it does not perform as intended.
Granular flters should be composed of durable free-draining granular materials. Section 6.6.2.3
discusses design requirements for grain-size distribution. Filter materials typically should not
contain more than fve percent clay- and silt-size particles, either before or afer placement.
Similarly, geotextiles should be designed based on the grain-size distribution of the soil or refuse
and apparent opening size of the geosynthetic material. Design requirements and limitations for
geotextiles, including guidance for evaluating the potential for chemical and biological clogging,
are presented in Section 6.6.2.3.2. Construction-related guidance for installation of geotextiles for
flters includes:
In preparing surfaces for geotextile placement, depressions, holes, and voids should
be flled so that the geotextile sheet is continuously supported. The geotextile should
be placed to loosely drape the surface with no sagging between surface contact
points. Continuous contact between the geotextile and the support material is con-
sidered critical in addressing clogging potential (Talbot et al., 2000). Geotextile
should not be placed over sharp or angular rocks that could tear or puncture it; an
intermediate layer of compatible fner material should be placed over such rock as a
bedding layer or protection bufer.
Geotextiles should be secured by sewing, pins, staples, or weights so that specifed
overlaps are maintained during construction.
Adjacent geotextile sheets should generally be placed with upstream layers overlap-
ping downstream layers.
In placing material on a geotextile, care must be taken to avoid punctures or tears.
The construction specifcations should limit the size and drop height of rocks to be
placed on the geotextile. Generally, stones weighing more than 250 pounds should
be placed with no free fall. Field trials should be made to verify that no damage will
occur due to the rock placement procedures, or a cushion layer of fner material may
be utilized to protect the geotextile.
Geotextiles should be covered or protected as soon as possible to prevent degradation
or damage from exposure or equipment trafc and to prevent fnes from accumulating
on the fabric. Manufacturer recommendations relative to exposure should be followed.
< PREVIOUS VIEW
11-51
Construction and Disposal Operations
MAY 2009
Geosynthetic materials are commonly used as flter layers for drainage systems at a refuse embank-
ment and as separating and reinforcing layers for channel linings and haul roads. Similar to granu-
lar flters, a geosynthetic material will allow the movement of water while preventing fne particles
from entering the voids of a coarser drainage medium. Properties of fltering geosynthetics such as
permitivity, hydraulic conductivity, percent open area and apparent opening size must be evalu-
ated as part of drainage system design. Where geosynthetics are employed at internal drainage
structures for impounding embankments, performance tests (e.g., gradient ratio test) should be
conducted using actual embankment materials. Separating and reinforcing geosynthetics are nor-
mally chosen based upon their strength properties such as puncture strength, grab tensile strength,
wide width tensile strength, grab tensile elongation, and trapezoid tear strength. Manufacturers
construction guidance for geotextile flters should also be followed when geotextiles are used as a
separation layer.
Care should be taken during construction to prevent runof from washing fne particles into an unpro-
tected flter. Filters should be constructed to an elevation slightly above that of adjacent materials. If
this is not practical, diversion ditches should be constructed or temporary coverings should be placed
over the flter.
Filter materials should be stockpiled, handled and placed in a manner similar to material used for
underdrains, as discussed in the following paragraphs.
11.7.2.2 Granular Underdrains
Underdrains typically consist of trenches flled with granular material, granular blankets, or granular
collection zones along stream botoms in valleys or at isolated seeps or springs. These drains can be
constructed with or without perforated pipes for conveying seepage from the collection zone to a safe
exit location. The granular zones are generally thin, and typically there are no compaction specifca-
tions for the granular material; however, where more extensive granular zones are present, such as
with chimney drains, provisions for placement should be provided.
Granular drains or collection zones must remain free-draining. Therefore, drain materials should
be selected to: (1) allow fow and act as flter protection for adjacent material (Section 6.6.2.3), (2)
resist deterioration over time, and (3) resist deterioration due to chemical atack. Table 11.6 lists
possible efects of chemical atack or weathering on gravels and crushed rock that might be used
as granular drainage materials. Granular aggregate used for underdrains and flters should be
obtained from reputable suppliers or verifed borrow sources. To minimize the potential for deg-
radation, rock used in granular drains should have high durability; applicable rock testing for
durability is discussed in Section 6.5.9.4. It is particularly important to avoid the use of most lime-
stone because of the possibility for generation of acidic leachates that can react chemically causing
dissolution or degradation. As a quick feld test, the presence of limestone in granular material can
be determined by placing a dilute solution of hydrochloric acid on the material. If an efervescent
or fzzing reaction occurs, the material should not be used. Laboratory test data for both the refuse
and the potential drain material should be carefully reviewed prior to the purchase or use of any
drain material.
Granular material must not be allowed to segregate prior to placement. Figure 11.5 shows cor-
rect and incorrect ways of stockpiling and handling granular materials. The stockpile location
and materials handling procedures should be such as to prevent siltation contamination by the
loading or unloading equipment or by wind- or water-carried particles. Well-graded materials
should not be dropped or allowed to roll down a slope for any signifcant distance, because the
larger particles will separate from other materials by moving to the outside of the pile as they
roll down the slope.
< PREVIOUS VIEW
11-52
Chapter 11
MAY 2009
FIGURE 11.5 STOCKPILING AND HANDLING OF GRANULAR MATERIALS
CRANE OR OTHER MEANS OF PLACING
MATERIAL IN PILE IN UNITS (NOT LARGER
THAN A TRUCK LOAD) THAT REMAIN WHERE
PLACED AND DO NOT RUN DOWN SLOPES.
PREFERABLE OBJECTIONABLE
METHODS THAT PERMIT THE MATERIAL TO ROLL
DOWN THE SLOPE AS IT IS ADDED TO THE PILE
OR PERMIT THE HAULING EQUIPMENT TO OPER-
ATE OVER THE SAME SURFACE REPEATEDLY.
FIGURE 11.5 STOCKPILING AND HANDLING OF GRANULAR MATERIALS
INCORRECT
FREE FALL OF MATERIAL FROM END OF
STACKER PERMITTING WIND TO SEPARATE
FINE FROM COARSE MATERIAL.
CORRECT
CHIMNEY SURROUNDING MATERIAL FALLING
FROM END OF CONVEYOR BELT TO PREVENT
WIND FROM SEPARATING FINE AND COARSE
MATERIALS. OPENINGS PROVIDED AS REQUIRED
TO DISCHARGE MATERIALS AT VARIOUS
ELEVATIONS ON PILE.
WIND
SEGREGRATION
LIMITED ACCEPTABILITY
PILE BUILT RADIALLY IN HORIZONTAL LAYERS
BY BULLDOZER WORKING WITH MATERIALS
DROPPED FROM A CONVEYOR BELT.
GENERALLY OBJECTIONABLE
UNLESS MATERIALS STRONGLY RESIST
BREAKAGE, STACKING PROGRESSIVE
LAYERS ON SLOPE NOT FLATTER THAN 3:1.
FIGURE 11.5 STOCKPILING AND HANDLING OF GRANULAR MATERIALS
< PREVIOUS VIEW
11-53
Construction and Disposal Operations
MAY 2009
If granular drains must be placed in thick layers, or are critical to the stability of the embankment because
of their position, the guidance in Section 11.6.6.2 relative to the placement and compaction of rock and
granular fll should be reviewed prior to granular drain construction. Should compaction be necessary,
the construction specifcation should account for potential breakdown of the granular material.
11.7.2.3 Drainage Pipe
Drainage pipes are ofen placed in portions of the refuse embankment to convey water from the
granular drains (internal or surface drains) to an acceptable exit point. Within the collection zone,
drainage pipes are normally perforated to facilitate infow of water that collects in the drain. The
remainder of the drainage pipe may or may not be perforated. Drainage pipes should be constructed
from a material resistant to chemical atack by refuse leachates and should be sized according to fow
capacity and structural requirements. They should generally be designed and installed such that they
are self-cleaning and the potential for accumulation of fne material in the pipe is minimized.
Although corrugated metal pipe (CMP) can ofen be protected by coatings, the potential for coating
damage during installation is a potential problem. Since plastic pipe is not normally susceptible to
corrosion from acids, it is ofen more suitable than CMP for drainage pipes in coal refuse.
When perforated pipe is placed, the perforations should normally be placed facing downward. This
causes the water to fow up into the pipe, minimizing the chance for transport of fnes into the pipe
and consequent reduction in fow capacity. Instructions from the pipe manufacturer are ofen explicit
in this respect and should be followed. For most perforated pipe, several rows of perforations are
normally provided around the circumference of the pipe. For this type of pipe it will not be possible
to orient all of the perforations facing downward. For installation in very fne materials, a double
layer flter and granular drain may be needed to prevent fne material from entering the perforations.
Such requirements should be detailed in the facility plans and specifcations.
Drainage pipes beneath embankments should be designed to withstand the loading of the embank-
ment, with specifcation of pipe support and backfll that prevents point loading on the pipe. Design
and construction of conduits for dams was published by FEMA (2005a), and additional guidance by
FEMA related to plastic pipe has also been published (FEMA, 2007). During construction, sufcient
fll must be placed over the pipe before traversing it with heavy equipment. Manufacturers recom-
mendations should be checked.
11.7.3 Culverts and Decants
Conduits are ofen used for transporting decanted impoundment water and food water beneath
embankments and also for culvert-type spillways. Conduit materials should be selected on the basis
of strength and resistance to corrosion. Potential corrosion and deterioration mechanisms for conduit
materials are discussed under Item I-C (Drainage and transport pipes) in Table 11.6. Progressive
deterioration of large conduits that are critical to refuse facility hydraulic systems is not acceptable,
and efort should be made to prevent loss of conduit fow capacity, leakage that can lead to material
infltration or exfltration, and structural distress.
Conduits require special bedding or thrust blocks at turns. Uniform contact around and under the
conduit is essential, particularly at thrust- block areas. When conduits associated with decant sys-
tems and culvert spillways are installed in critical areas, such as through impounding embankments,
special provisions are required for performing compaction beneath and around the conduit to mini-
mize the potential for seepage along the exterior of the pipe. Compaction beneath the haunches of
the pipe is difcult and normally can not be accomplished unless special materials and designs are
employed. Concrete cradles and sand bedding have been used to improve the support beneath and
around conduits.
< PREVIOUS VIEW
11-54
Chapter 11
MAY 2009
Other more specialized materials and methods such as use of fowable backfll or the cut-earth
cradle method can be employed for conduit installation. Flowable fll, or controlled low-strength
material (CLSM), is a low-strength, fne-aggregate concrete with unconfned compression strengths
between 50 and 200 psi. The use of fowable fll to backfll a conduit trench and serve as bedding for
the conduit provides frm support extending from the spring line around the base of the conduit.
Flowable fll has the added beneft of having low hydraulic conductivity. The mix must be designed
to prevent shrinkage and to have satisfactory strength/deformation characteristics, as discussed in
Sections 6.5.10.2 and 6.6.6.3.3. Either shrinkage or cracking of fowable fll will allow a fow path for
seepage.
The cut-earth cradle method involves compaction of select material at the base of the conduit location
and subsequent excavation using a specialized atachment on an excavator resulting in a semi-cir-
cular cradle excavation. The cradle excavation is sized and shaped so that it closely fts the conduit.
A material such as bentonite powder has been used to compensate for small, discontinuous gaps
between the cradle and the pipe. The depth of the cradle should be approximately one-half the diam-
eter of the pipe. This method requires a high level of quality control in order to prevent a fow path
from occurring along the pipe.
To collect and discharge any possible seepage along the conduit in a controlled manner, a flter dia-
phragm is recommended regardless of the method employed for conduit backfll and support.
11.7.4 Concrete Structures
Coal refuse disposal facility appurtenant structures that may be totally or partially constructed of
concrete include spillways, decant structures, culverts and drainage channels, as well as foundations
for various small structures. Such concrete work is not unique to coal refuse disposal facilities, and
most applications are extensively discussed by the ACI (2007), the USBR (1992b) and in numerous
concrete design textbooks.
Chemical atack and deterioration and related preventative measures are key considerations in the
use of concrete for any facility structure that may be in contact with coal refuse or leachates from coal
refuse. Table 11.6 presents a summary of major considerations. Although acidic leachates are an obvi-
ous corrosive environment, it is emphasized that for concrete durability it is also important to guard
against the sulfates common to coal refuse and coal refuse leachates. The USBR (1987a) shows typical
examples of sulfate atack and discusses methods for retarding it.
Before specifc protective measures for concrete exposed to coal refuse and leachates are selected, the
acidic potential and sulfate contents of the refuse should be evaluated. Chemical testing is recom-
mended. Concrete structures exposed to natural ground adjacent to coal refuse disposal facilities
should also be protected from deterioration by chemical atack. Groundwater in and around coal
refuse may contain corrosive elements and can cause concrete deterioration even if the concrete is not
in direct contact with coal refuse.
Minimum practices and protective measures for concrete used in or near coal refuse disposal facili-
ties should include:
Use of sulfate-resistant cement (preferably Type V, but in some mine environments
Type II cement will sufce). Depending on cement availability, Type II or Type I ce-
ment in combination with pozzolans (e.g., fy ash, slag cement) may provide ad-
equate sulfate resistance.
Low water-cement ratios and relatively rich mixes should be used so that dense,
high-strength concrete results.
< PREVIOUS VIEW
11-55
Construction and Disposal Operations
MAY 2009
Air entrainment should be used to increase the workability of low water-cement
ratio mixes and to decrease the hydraulic conductivity of the hardened concrete.
The use of other water reducing chemical additives to allow further reduction of the
water-cement ratio and to provide increased strength should also be considered.
Extra cover concrete should generally be provided over all reinforcing steel.
Fresh concrete should be thoroughly vibrated to maximize densifcation and to
eliminate honeycombing.
Concrete should be thoroughly cured to maximize strength gain.
Many protective coatings discussed in the following section are also applicable to concrete structures,
although application and maintenance may be more difcult than for application to metals.
11.7.5 Gates, Valves, Pipelines and Other Metal Work
Metal work in and around a coal refuse disposal facility must be protected against corrosion from
chemical atack. Special alloy steels may ofer good protection against corrosion, but their selection
should be made with care because some types of alloy steels are actually more susceptible to atack
from certain chemicals than regular construction grade carbon steels. The cost of many alloy steels
is similar to carbon steel, but cost must be evaluated on a case-by-case basis. Table 11.6 summarizes
various types of chemical atack and corresponding measures for protection.
Installation of metal products with protective coatings is common, but installation should be accomplished
in strict accordance with the recommendations of the coating manufacturer. Various coatings are available
that will provide adequate protection if the integrity of the coatings is maintained during construction
and operation. Coatings should be: (1) continuous, (2) not support bacterial growths or absorb water, and
(3) not deteriorate with time. Epoxy coatings are very durable and provide excellent protection against
corrosion and bacterial growth. However, feld application of epoxy coatings is difcult and requires spe-
cial surface preparation per the manufacturers recommendations. Coal tars meet all of the coating criteria
and are one of the best bituminous materials available for protection of steel surfaces. Asphalt coatings
are an alternate low-cost bituminous coating. However, if bituminous coatings are scratched or damaged
during installation, they must be repaired prior to completion of the installation.
Special vinyl chloride or chlorinated rubber coatings are also available for corrosive protection.
Application of these coatings must be made in strict accordance with recommendations of the manu-
facturer, as there are signifcant diferences in methods of application.
Imperfect coating of metal may actually lead to more rapid deterioration than no coating at all. This
can occur where galvanic action is very strong. If the predominant portion of the metal is protected
by materials that resist electrical currents, galvanic action will concentrate at points of imperfection,
leading to very rapid removal of the metal at these locations. Protection against this type of corrosion
requires installation of a sacrifcial anode designed by a professional experienced in corrosion protec-
tion. USACE (1985) and USACE (2004) provide background information and guidance on testing and
design for cathodic protection.
11.7.6 Support Structures
Ofces, maintenance buildings, loading bins, etc. should be built in accordance with applicable
building codes and generally accepted construction practices. Coal refuse can ofen be used efec-
tively for structural flls on which such structures are constructed. The same precautions against
material deterioration due to acid or sulfate atack should be taken. Drainage and utility pipes,
foundations and electrical conduits are all subject to corrosive action if exposed to coal refuse or
coal refuse leachates.
< PREVIOUS VIEW
11-56
Chapter 11
MAY 2009
11.7.7 Mine-Opening and Auger-Hole Seals and Drains
Depending upon their location within an embankment or impoundment area, the presence of
mine openings and auger holes in coal seams should be addressed with backflling, barriers,
bulkheads, seals or drains. The construction materials for these structures can range from struc-
tural materials such as cast-in-place concrete or block to aggregate materials for construction of
flters and drains. On-site borrow materials may be specifed for backfll or for compacted fll as
part of a barrier.
11.8 QUALITY CONTROL AND FIELD TESTING
Quality control and feld testing are required for verifcation that a refuse facility is constructed
according to plans, specifcations and the operation and maintenance plan. The principal concerns
for construction of a coal refuse disposal facility include:
Compaction control where refuse is to be used as structural fll.
Compaction control where borrow materials are to be used as structural fll.
Verifcation that all areas that require structural fll are so constructed.
Verifcation that material properties that afect construction operations, such as
grain size, water content and compaction characteristics are in accordance with
specifcations.
Verifcation that facility appurtenant structures, especially conduits and open chan-
nel spillways, are being constructed in accordance with design requirements, manu-
facturers' recommendations and good construction practice.
Activities that require quality control can be identifed through review of the design report, plans,
specifcations, operation and maintenance plan for the facility, and technical chapters of this Manual,
particularly Chapter 6.
11.8.1 Compaction Control
The monitoring of compaction in the feld is essential for verifcation that structural fll refuse is
compacted to a density meeting design requirements. The density of fll material should be tested
in order to verify that the pertinent engineering properties of the fll that are a function of density,
such as shear strength and hydraulic conductivity, are consistent with the values used in the design.
Compaction control entails feld testing during the construction of embankments and at other areas
requiring structural fll refuse. Compaction test locations should be selected so as to include areas
where routine compaction activities are performed, as well as areas that are more difcult for com-
paction equipment to access with repeated passes or where successful compaction is doubtful. At
some sites, a grid system has been employed to aid in test location selection, and for each lif random
grid points are chosen for testing. USBR publications (1987a, 1992a, 1998) discuss methods for testing
compacted fll materials.
The in-situ density testing of structural fll has become much easier and more efcient with advances
in nuclear testing methods. Unless special conditions arise during fll construction, most structural
fll refuse, fne-grained soils, and fne gravels can be density tested using a nuclear density-moisture
gauge. USACE (1995b) provides guidance on the use of these types of gauges, which should be oper-
ated in the direct transmission mode. Nuclear density-moisture gauges provide direct readings of in-
situ density and the moisture content of fll. The instrument operates by projecting radiation (gamma
rays) into the fll and measuring returns with a Geiger-Mueller detector. The instrument is easy to
use and requires only observation of a digital read-out. However, the presence of varying amounts
of hydrocarbons in coal refuse can cause the gauge to indicate higher than actual moisture contents.
< PREVIOUS VIEW
11-57
Construction and Disposal Operations
MAY 2009
Therefore, samples should be collected from the tested fll for laboratory moisture analysis so that
accurate calculation of dry density can be performed.
For evaluation of the degree of compaction, measured feld densities can be compared to Proctor den-
sity curves (Chapter 6) developed from laboratory testing. All feld test results should be recorded. If
test results indicate a lack of compliance with fll density requirements, compaction procedures should
be modifed. Areas of low density should receive additional compaction. If necessary, unacceptable fll
should be removed and replaced, and recompaction should be performed. If it is determined that the
fll is being compacted to densities greater than the maximum value from the Proctor test, an evalua-
tion of the material (grain size and specifc gravity) and the associated Proctor density curves may be
necessary. In some cases, the presence of oversized particles, that is particles too large to be included
in the Proctor mold (ASTM D 698), may be causing overestimation of density. If this is suspected, a
sieve analysis should be performed and a rock-correction factor should be developed in accordance
with ASTM D 4718, Standard Practice for Correction of Unit Weight and Water Content for Soils
Containing Oversize Particles, and used to recalculate the maximum dry density of the refuse.
Application of the ASTM D 4718 may be difcult or precluded when coarse refuse contains more than
25 to 30 percent of plus--inch particles, even afer the oversize correction provisions using labora-
tory determined specifc gravities are employed. Sometimes the oversize correction yields unreason-
able target densities (usually higher than achievable), and in this case other compaction tests or feld
procedures should be considered. While determination of relative density in accordance with ASTM
D 4253 and ASTM D 4254 is a possibility, this is generally not recommended because of the fne con-
tent in coal refuse. Another possible method for compaction control is the roller pass test (WVDOT,
1999). This method is discussed in Section 11.5.1.
Alternative material testing methods may be employed when more than 25 to 30 percent of plus-
-inch particles are encountered. One such testing method is specifed in Appendix VIA of the
USACE Publication EM 1110-2-1906 titled, Laboratory Soils Testing (USACE, 1986). This method
employs a 12-inch-diameter mold that accommodates larger particle sizes. Disadvantages to this
method are that it is more costly than the ASTM D698 method, and many geotechnical laboratories
are not equipped to perform the test. However, if refuse with a signifcant proportion of oversize
particles is routinely generated, it may be efective to use the USACE test to supplement or replace
the roller pass method.
In addition to periodic feld density tests, visual observation of refuse placement and compaction
should be performed regularly. If compacted material tests as having an acceptable density, but the
surface is not frm, then the compaction specifcation should be re-evaluated. Regular observations
by feld personnel can ofen lead to acceptable modifcation of compaction procedures. Such observa-
tions are discussed in Section 11.6. Key observations for control of earth-fll operations are discussed
by Church (1981) and USBR (1998). These observations, which typically apply to refuse disposal
operations, include:
Changes in lif thickness with subsequent passes of compaction equipment (includ-
ing controlled routing of haul units) should be noted.
Compaction equipment should not excessively weave or rut the fll. If this oc-
curs, the material is too wet to be properly compacted.
The compacted fll surface exhibits pumping under construction equipment trafc,
which may indicate that lifs below the surface were compacted at too high a mois-
ture content, and need to be uncovered and reworked.
The feet of a sheepsfoot or tamper-type roller will frst penetrate the loose material
of a new lif, but should then begin to ride up on the material with each successive
pass, as compaction is performed.
< PREVIOUS VIEW
11-58
Chapter 11
MAY 2009
A consistent and reliable compaction procedure based upon a sufcient number of passes with
appropriate equipment should be established. The USACE (1995b) provides general guidance rela-
tive to the number of passes required for various earthfll materials. Site-specifc conditions must be
considered in establishing compaction parameters.
It is recommended that structural fll refuse be compacted to at least 95 percent of the maximum dry
density determined by the ASTM D 698. The refuse should be placed at a moisture content in the
range indicated in the plans and specifcations, typically in a range of 2 to +3 percent of optimum.
The moisture content should be uniform throughout each lif. At least one feld density test should
be conducted for every 2,000 cubic yards of compacted structural fll with at least one test per lif. A
common specifcation for compaction of pipe backfll and around structures is: at least one density
test for every 200 cubic yards with at least one test per lif. Because of the importance of achieving
adequate and consistent compaction of pipe backfll, more frequent density testing should be con-
ducted. It is recommended that multiple tests be performed so that the testing frequency is sigmif-
cantly lower than every 200 cubic yards. Testing should also be performed whenever it is suspected
that adequate compaction is not being achieved.
If failing density tests occur, it is the responsibility of feld personnel to determine the cause of the
failed tests. Generally, failed density tests are a result of either insufcient compaction or a change of
material from that tested in accordance with ASTM D 698. If it suspected that the fll was not placed
and compacted adequately, the limits of the lif and/or area should be marked, and mine personnel
should be notifed that the lif/area requires reworking and additional compaction. No additional
refuse should be placed in the area with failed density tests until it has been adequately compacted.
Additional density tests should be taken afer the lif/area is reworked to verify the efectiveness of
the additional compaction efort. If it is believed that the lif/area has been placed and compacted
adequately, and a variation in grain-size distribution of the refuse being placed is suspected, a new
sample should be collected for the ASTM D 698 testing. The Proctor curve(s) should be verifed on a
regular basis, typically between quarterly and annually.
11.8.2 Material Testing
The material properties of coal refuse, borrow materials, flter or drain gravels, and concrete can be
determined by on-site feld testing and/or by submital of test samples to a qualifed testing labo-
ratory. Laboratories should be selected based on qualifcations. The AASHTO Materials Reference
Laboratory (AMRL) performs evaluation and accreditation programs for many construction materi-
als, and the Geosynthetic Accreditation Institute Laboratory Accreditation Program (GAI- LAP)
performs similar evaluations for synthetic materials. Laboratories should be accredited through these
programs for the type of testing being performed. Water contents and fll densities are best tested in
the feld for real-time assessment of compaction efciency, but occasional test verifcation by an inde-
pendent source may be appropriate. Strength testing of refuse, soils, concrete and other construction
materials generally requires laboratory equipment and in some instances may be incorporated into
quality control programs cited in the construction specifcations.
Routine sampling and testing of construction materials is recommended during the development of
a refuse disposal facility. The resulting records can be used for future evaluation of material perfor-
mance and for submitals to regulatory agencies. Most importantly, material sampling and testing
during construction increases confdence that the work is being performed in compliance with design
requirements.
< PREVIOUS VIEW
Chapter 12
MONITORING, INSPECTIONS
AND FACILITY MAINTENANCE
12-1 MAY 2009
Routine monitoring and inspections are essential to the successful construction and operation of a
coal refuse disposal facility. Inspections, as described herein, include the components of observa-
tions, testing and instrumentation measurements, as performed by a representative of the Mine
Owner/Operator (Operator) in accordance with 30 CFR 77.216-3. Personnel responsible for moni-
toring and inspections require MSHA impoundment inspection training and/or qualifcations and
should work with a registered Professional Engineer who is familiar with the design of the disposal
facility (Engineer). It is recommended that the Operator designate an Engineer to provide assistance
with the implementation of design plans and specifcations, assess the performance of the facility,
report on the progress of the construction work, and render professional opinions regarding the
conformance of construction, operation, and maintenance activities with design plans and specif-
cations. Note that the guidance in this chapter is not intended to address work-site safety issues or
daily safety inspections.
Well executed monitoring and inspection programs during construction will facilitate develop-
ment according to plans and specifcations and aid in identifying modifcations to the design that
may need to be made. Routine inspections during operation of the facility allow for the identifca-
tion of potential problems and resolution in a timely manner. Observations made during monitor-
ing and inspections can provide a basis for either increasing or decreasing certain maintenance
activities. This chapter is generally focused upon monitoring and inspection programs and their
relationship to facility maintenance during development and operation of the refuse disposal facil-
ity. These programs may also be required during abandonment and reclamation of a coal refuse
embankment.
Monitoring and inspection requirements should be established during the design of a coal refuse
disposal facility and should be documented in the Operation and Maintenance Plan, which should
clearly indicate the roles and responsibilities of various parties in monitoring and inspection activi-
ties. Sufcient controls should be in place so that, when monitoring and inspection is performed
as specifed, the intent of the design is met and satisfactory performance of the facility is achieved.
Monitoring normally includes observation of work activities during facility construction, including
those activities required to initially prepare the site. Monitoring may encompass such activities as
visual observation, collection and analysis of quality control samples, performance of feld and labo-
ratory testing, evaluation of construction methods, collection and review of instrumentation readings
and evaluation of survey data.
< PREVIOUS VIEW
12-2
Chapter 12
MAY 2009
Inspections for evaluating performance are typically conducted for the overall facility and at areas
that have been recently constructed. These inspections normally include visual observations; collec-
tion of physical data, instrument readings and fow rate measurements (if applicable); and evalua-
tion of the assimilated data. Photographing of site conditions, during construction and operation, is
an excellent way to document site conditions and any changes in conditions and is a recommended
practice. Information from previous inspections and monitoring activities should be reviewed and
compared to current data and ploted instrumentation data so that an informed evaluation can
be made. In accordance with 30 CFR 77.216-3, facility inspections for impounding refuse facili-
ties must be performed on a 7-day basis unless otherwise approved by the MSHA District Man-
ager. Annual reporting and certifcation requirements for impoundments are provided in 30 CFR
77.216-4. Additionally, some states have inspection, reporting, and certifcation requirements that
difer from the federal requirements. Annual dam safety inspections may be required by some state
regulatory agencies, but these can generally be combined with the MSHA annual reporting and
certifcation requirements.
It is recommended that monitoring and inspections generally be performed under the guidance of
a registered Professional Engineer who is familiar with the design (including safety criteria for the
structure) and involved in the implementation of the construction plans and specifcations. Thus,
this individual can prepare required reports and certifcations and can assess: (1) conformity with
the approved design, (2) signifcance of the instrumentation records, (3) potential conditions that
may warrant notifcation of agencies including MSHA, and (4) design modifcations and remedial
construction activities that may be needed. Typically, the design engineer performs or contributes to
this activity. In describing the monitoring and inspection activities in this chapter, and particularly
actions that may be in response to observed conditions, test results, or instrumentation readings,
reference is made to the Engineer in situations that are typically important for such an individual to
assess and determine subsequent action. The Engineers involvement in other activities may also be
important, and the site-specifc Operation and Maintenance Plan should cite monitoring and inspec-
tion requirements. By maintaining the involvement of a registered Professional Engineer who is thor-
oughly familiar with the design during construction and operation, the mine operator is in position
to assess conformity with the approved plan and to respond to regulatory agencies regarding perfor-
mance or compliance concerns.
If potentially hazardous conditions develop, MSHA requires under 30 CFR 77.216-3 that the Opera-
tor take action to: (1) eliminate the condition, (2) notify the District Manager, (3) notify and prepare to
evacuate miners who may be afected, and (4) perform inspections on an eight-hour or more frequent
basis. The Engineer should be a source of technical advice in such situations. Potentially hazardous
conditions can be identifed based on criteria established for Emergency Action Plan preparation, as
discussed in Chapter 14.
Facility maintenance includes routine recurring actions for establishing vegetation and controlling ero-
sion, periodic actions afer unusual events such as clearing debris, and long-term actions in response
to deterioration of structures, such as repair of channel linings and maintenance of embankment
slopes. Monitoring and inspections may identify conditions that are inconsistent with the Operation
and Maintenance Plan because of difering site conditions, construction procedures, or performance.
In such situations, new maintenance requirements may need to be developed.
12.1 MONITORING AND INSPECTIONS
There is an inherent degree of uncertainty in the engineering design of a refuse disposal facility that
can normally be ofset by applying factors of safety in analytical procedures and by using conserva-
tive estimates of material properties. Engineers, particularly in the geotechnical felds, have devel-
oped observa tional/monitoring procedures to deal with these uncertainties without being overly
< PREVIOUS VIEW
12-3
Monitoring, Inspections and Facility Maintenance
MAY 2009
conservative. The following is a brief listing of the purposes for monitoring and inspection and the
basis for their incorporation into refuse disposal facility development:
Construction quality can vary depending on the specifc activity, site conditions,
equipment and personnel, all of which can afect facility performance. Monitoring
and inspection provide a means of controlling adherence to design requirements and
documenting construction quality.
Geologic conditions can vary signifcantly at a site. These variations can be accom-
modated by monitoring and inspection programs during construction and the use
of reasonable factors of safety in the design. Monitoring during construction enables
evaluation of the validity of design assumptions and analyses and provides an
opportunity to make changes as needed.
Because it is impractical to investigate all natural conditions at a site or anticipate
every aspect of facility operation, confrmation of design assumptions through moni-
toring and inspection can ofset limitations in site exploration.
The long operating life of a refuse disposal facility can lead to many changes in site
conditions. For example, changes in coal cleaning procedures and refuse characteristics
may occur due to mining of diferent coal seams during the operational period. An
efective monitoring and inspection program can identify when signifcant changes are
occurring or performance problems are developing, so that appropriate modifcations
can be made quickly when implementation is least difcult and expensive.
Monitoring and inspection programs form the basis for determining when routine main-
tenance is needed in order to limit the deterioration of important facility components.
Monitoring and inspection rely heavily on visual observation and are supported by quality control
testing, instrumentation, and engineering evaluation. The complexity of a monitoring program can
range from simple visual examination of surface conditions to sophisticated surveys using electronic
equipment. The simplest means are usually the most efective, especially for the constantly changing
conditions associated with coal refuse facilities. A more complicated monitoring and instrumentation
program may be justifable when the information obtained can result in an improved design, opera-
tion or signifcant savings in construction costs. Regardless of complexity, a monitoring program is
successful only if the information obtained is pertinent to the design and operation and the data are
accurate and measured, recorded and properly interpreted by qualifed persons in a timely manner.
Accordingly, any monitoring program must be based on knowledge of design requirements, and the
observations and measurements must be obtained and interpreted by staf cognizant of the purpose
for and limitations of the instrumentation.
12.1.1 Monitoring and Inspection Objectives
Important objectives for monitoring and inspection at coal refuse facility sites include the following:
Validating the assumptions upon which the design was based A valuable function
of monitoring programs is to provide data for verifying that the assumptions upon
which the facility design is based are valid. Therefore, it is critical that monitoring
data be evaluated by an engineer who is thoroughly familiar with the design and
associated engineering analyses and assumptions and who is responsible for certif-
cation of the work.
Providing data for improving construction methods As an example, the compac-
tion equip ment being used at a site may produce an undesirable particle compaction
arrangement that, in turn, afects properties such as strength or hydraulic conduc-
< PREVIOUS VIEW
12-4
Chapter 12
MAY 2009
tivity. Through observation of embankment construction, changes can be made in
equipment selection, trafc routing, drying or weting, or other factors such that the
desired results are achieved.
Providing a means for early detection of potential problems or areas requiring main-
tenance or remedial improvement A feld monitoring and inspection program can
identify problems before conditions deteriorate to the point that extensive repair is
required or failure can occur. The natural variability of earthen and coal refuse ma-
terials may result in unexpected deviations from the original design. Changes in the
rate of refuse placement could alter embankment elevations and pool levels, afect-
ing embankment stability.
A project-specifc failure modes analysis can be useful in establishing site-specifc monitoring and
inspection objectives, particularly for facilities with impounding embankments (Martin and Davies,
2000). Through an objective review of potential modes of failure of the disposal facility, considering
all loading conditions and design features, the most important monitoring, instrumentation, and in-
spection objectives can be identifed. Martin and Davies suggest the following process:
Identifcation of potential failure modes and corresponding warning signs.
Evaluation of how quickly failure could occur and how potential problems could be
detected before they develop into incidents.
Critical assessment of how likely it is that warning signs will be detected, recog-
nized, reported and acted upon.
Identifcation of instrumentation types and locations that provide the most relevant
monitoring data.
Evaluation of the signifcance of data trends as opposed to single measurements/ob-
servations.
Evaluation of the value of establishing acceptable parameter limits or other limiting
criteria.
Review of lines of responsibility and communication.
Review of data management, interpretation, action plans, and reporting.
Figure 12.1, as modifed from Martin and Davies (2000), illustrates the relationship of failure modes,
warning signs, and surveillance measures for tailings dams.
12.1.2 Visual Observations
The simplest, but ofen most valuable, monitoring method is visual observation by experienced per-
sonnel of important facility components and the surrounding area on a schedule compatible with site
activity and the rate at which the facility confguration is changing. The required frequency for obser-
vations and the level/knowledge of the individual performing the observations should be established
by the designer and should be consistent with federal and state requirements. The following should
be considered when developing a schedule for visual observations:
The type of refuse facility (impounding or non-impounding) and stage of construction.
Specifc critical construction items (e.g., internal drains, surface channels, decant
pipe, cut-of trench) that should be observed.
Specifc operational areas or items (e.g., spillway operation) required for evaluating
performance and potential design modifcation.
Identifcation of components/areas that require only periodic monitoring and inspec-
tions and their frequencies.
< PREVIOUS VIEW
12-5
Monitoring, Inspections and Facility Maintenance
MAY 2009
Components/items to be observed depend upon the facility confguration. Typical factors include:
Facility confguration and type of refuse disposal
Geotechnical design requirements
Site mining and foundation requirements
Appurtenant hydraulic facilities (spillways or decant pipes)
Environmental control provisions
Construction procedures
FIGURE 12.1 FAILURE MODES, WARNING SIGNS
AND SURVEILLANCE MEASURES
LOSS OF DIVERSIONS/
SPILLWAY FLOW
CAPACITY
LOSS OF CREST
ELEVATION
INTERNAL EROSION
OF DAMFILLS
INTERNAL EROSION
OF FOUNDATION
MATERIALS OR
PIPE BACKFILL
SURFACE EROSION
OF DAM SLOPES
SLIDE THROUGH
DAMFILL AND/OR
FOUNDATION
BREAKTHROUGH
OF TAILINGS
INTO MINE
BLOCKAGE/FAILURE OF
DIVERSIONS/SPILLWAY
GRADUAL INFILLING
OF DIVERSIONS
TAILINGS POND
LEVEL
SETTLEMENT
CHANGE IN CREST
GEOMETRY/CRACKING
CHANGE IN SLOPE
GEOMETRY (US/DS)/
CRACKING
CHANGE IN SEEPAGE
CONDITIONS/LOCATIONS/
VEGETATION
SINKHOLES, TURBID
SEEPAGE
DISCHARGE
GRADUAL TREND TO
INCREASED TURBIDITY
IN SEEPAGE
SEEPAGE INDUCED
SLOUGHING IN
EMBANKMENT
RAISED SATURATION
LEVELS, SOFT
GROUND CONDITIONS
LARGE, SUDDEN
CHANGES IN
PIEZOMETRIC LEVELS
LARGE, SUDDEN
CHANGES IN
SEEPAGE FLOWRATES
LONG-TERM
PIEZOMETRIC
CHANGES
LONG-TERM
INCREASE IN
SEEPAGE FLOWRATES
LARGE, SUDDEN
MINE DISCHARGES
INSPECTIONS OF
DIVERSION/SPILLWAYS
TAILINGS POND
LEVEL CHECKS AND
PERIODIC SOUNDINGS
REQUIRED CREST
ELEVATION VERSUS
DECANT ELEVATION
SURFACE MONUMENT
SURVEYS
PIEZOMETERS/
WELLS
INCLINOMETERS/
SETTLEMENT GAGES/
EXTENSOMETERS
SEEPAGE WEIRS
INSPECTION OF DAM
SLOPES/CREST/
DOWNSTREAM TOE
SEEPAGE WATER
QUALITY/TURBIDITY
CLOGGING OF TOE
DRAINS/OXIDATION
PRODUCTS
INSPECTION OF
ABUTMENTS
INSPECTION OF
MINE DISCHARGE
OVERTOPPING
MASS
MOVEMENT
INTERNAL
EROSION
SURFACE
EROSION
MINE
SUBSIDENCE
FREEBOARD
LESS THAN
DESIGN CRITERIA
TAILINGS DAM
FAILURE MODES
ADVERSE
CONDITIONS
RAPID WARNING
SIGNS
GRADUAL
WARNING SIGNS
SURVEILLANCE
MEASURES
DECANT PIPELINE
INSPECTIONS
MINE DISCHARGE
DOESNT CORRELATE
TO RAINFALL
TAILINGS LEVEL
RISING FASTER
THAN ANTICIPATED
LOSS OF UPSTREAM
EROSION
PROTECTION
(ADAPTED FROM MARTIN AND DAVIES, 2000)
FIGURE 12.1 FAILURE MODES, WARNING SIGNS AND SURVEILLANCE MEASURES
FIGURE 12.1 FAILURE MODES, WARNING SIGNS AND SURVEILLANCE MEASURES
< PREVIOUS VIEW
12-6
Chapter 12
MAY 2009
Checklists should be developed for each refuse disposal facility and should be incorporated into
the Operation and Maintenance Plan. These checklists should be used in conjunction with the
facility plans and specifcations. Figure 12.2 presents an example of a data collection form that
is incorporated into a site as-built drawing with topographic contours. The combination of the
data collection form and as-built drawing provides a good point of reference during monitoring
activities and allows areas of concern to be marked without the need for cumbersome descrip-
tions/text. Items such as density test locations, seeps, sloughs, erosion and slurry discharge loca-
tion can easily be indicated on the drawing.
FIGURE 12.2 EXAMPLE DATA COLLECTION FORM
SCALE
400 400 0
FIGURE 12.2 EXAMPLE DATA COLLECTION FORM
FIGURE 12.2 EXAMPLE DATA COLLECTION FORM
< PREVIOUS VIEW
12-7
Monitoring, Inspections and Facility Maintenance
MAY 2009
Table 12.1 is a generic list that covers most aspects of refuse disposal facility activity. This list can be
used as the basis for development of a site-specifc data collection form. Other available references
include the MSHA Coal Mine Impoundment Inspection and Plan Review Handbook prepared by MSHA
(2007) and the Guidance Document for Coal Waste Impoundment Facilities & Coal Waste Impoundment
Inspection Form prepared by the West Virginia Water Research Institute (2005).
Extensive digital photography is recommended for documenting visual inspections of a coal refuse
disposal facility. Photographs not only document the observations made during monitoring or
inspection activities, but they can also be used later to track the construction progress of the facility.
Photographs are also invaluable for comparing how conditions such as seepage areas change over
time. The location and/or orientation of the camera for each photograph should be documented on
the as-built drawing or data collection form.
Hand-held global positioning system (GPS) units are very useful for locating items of interest. Most
hand-held GPS units can provide position location to within about 10 feet. For most refuse facilities,
this accuracy is sufcient for locating sampling or testing locations, wet areas, minor seeps, or other
non-critical items. The use of a hand held GPS unit or altimeter unit for elevation determinations is
not currently recommended due to lack of accuracy. It should be noted that the GPS coordinates may
not conform to the coordinate system used by the mine. Sofware for performing coordinate conver-
sions is available form commercial vendors.
12.1.2.1 Schedule and Checklist for Visual (and Other) Observations
A schedule for observations should be provided in the Operational and Maintenance Plan. The sched-
ule of observations must be compatible with federal and state regulatory agency inspection require-
ments. In accordance with 30 CFR 77.216-3, inspections (not to exceed 7-day intervals) are required
for all impounding facilities meeting the requirements of 30 CFR 77.216(a). Section 12.3 provides
additional details relative to embankment/impoundment safety inspections. The recording of data or
collection of measurements from installed instrumentation should also be defned in the schedule of
observations. Chapter 13 provides additional details concerning the types of instruments available
and information related to their installation and use. While 7-day visual inspections are required,
continuous monitoring may be appropriate for some critical construction activities (e.g., structures
that are ultimately buried, such as decant pipe and internal drain installations, instrumentations,
etc.). Alternate frequencies (more or less frequent than every 7 days) for instrumentation monitoring
may be desirable.
Inspection of impounding embankments during and afer signifcant precipitation events is recom-
mended. This is a good time to observe the operation of decant and spillway facilities and the outlets
to internal drains and to verify that they are functioning properly. Special atention should be paid to
whether fow is unobstructed and erosion protection, especially at outlets, is functioning properly.
Another important time to observe the operation of a dam is during frst flling. The frst time the foun-
dation and embankment are subjected to seepage pressures and the frst time that fow occurs through,
and possibly along, conduits, is an important time to verify that performance is acceptable. Slurry
impoundments are constructed over a long period of time and are flled as they are constructed. Special
atention should be paid when the saturation level reaches the level of new features, such as conduits.
A schedule for observations should be developed for each major aspect of construction of a refuse
disposal facility, including:
Initial site development
Liner installation (if applicable)
Starter dam/embankment construction
Refuse embankment construction
< PREVIOUS VIEW
12-8
Chapter 12
MAY 2009
Instrumentation installation and data measurement
Appurtenant hydraulic facility construction
Backfll and abandonment of impoundments
Mine barrier, bulkhead and seal construction, if applicable
To be most useful, the checklist/data collection form presented in Table 12.1 should be tailored to
site conditions. The completed data collection form/checklist (e.g., Figure 12.2) should be regularly
reviewed by the Engineer. In addition to descriptions of the conditions observed, deviations from the
plans and specifcations should also be indicated on the form and should be reported to responsible
facility personnel (Engineer, Contractor, and Operator).
12.1.2.2 Facility Development
12.1.2.2.1 Site Preparation
Site preparation includes activities such as clearing and grubbing, topsoil/subsoil stockpiling, foun-
dation preparation, and liner installation. These items relate to both initial facility development and
facility expansion and are typically performed as individual construction activities rather than as a
part of ongoing operations. A brief summary of typical observations associated with site preparation
is provided in the following:
Clearing and grubbing Clearing and grubbing typically includes removal of
unsuitable materials such as trees, tree stumps and roots, brush and other vegetation.
Important items related to clearing and grubbing that should be observed and noted
include: location and extent, adequacy of vegetation removal, debris disposition,
and identifcation of unanticipated and/or suspect conditions (e.g., mine openings,
landslides).
Topsoil/subsoil stockpiling Important items associated with topsoil/subsoil
stockpiling that should be noted include: location and extent, drainage control
provisions, depths removed, location and confguration of stockpiles, identifca-
tion of unanticipated conditions (e.g., landslides, springs, potential for additional
soil removal). Areas where the planned embankment and impoundment are to be
located should be observed, particularly where soils have been excavated, to deter-
mine potential natural seepage conditions that may not have been evident before
disturbance.
Foundation preparation Important items associated with foundation preparation
that should be noted include: location and extent of work areas; preparation work
performed prior to placement of any refuse materials (removal of unsuitable mate-
rial, proof-rolling, slope modifcation and removal of rock overhangs, spring col-
lection and underdrain installation, etc.); installation of cutofs, keys or benches in
preparation for refuse placement; and construction of mine barriers or bulkheads.
Liners If liners are required, the important items to be noted include: materials
(natural or synthetic including properties), thickness placed, compaction (equipment
type, tests), performance observations (frm or pumping), and cover protection (soil,
vegetation, refuse).
12.1.2.2.2 Starter Embankment and Internal Drains
Following site preparation, construction of a starter embankment using borrow material (soil and
rock) or coarse refuse is the frst step in preparing a disposal facility site for ongoing refuse dis-
posal. Starter embankments may be zoned or homogeneous. Important items to be observed include
< PREVIOUS VIEW
12-9
Monitoring, Inspections and Facility Maintenance
MAY 2009
TABLE 12.1 TYPICAL OBSERVATION AND INSPECTION CHECKLIST
Item
General
Condition
Changes
from
Previous
Inspection
Item
General
Condition
Changes
from
Previous
Inspection
FACILITY DEVELOPMENT
DISPOSAL OPERATIONS
AND MAINTENANCE
1. Erosion and Sediment
Controls
a. Diversion Ditches
b. Silt Fences
c. Sediment Traps
d. Sediment Ponds
1. Stage of Construction
and Refuse Placement
(lift thicknesses, com-
paction, scarifcation,
benching)
2. Clearing & Grubbing 2. Working Surface and
Area Maintenance
(survey control,
grading and
drainage control).
Refuse Placement
(lifts, compaction,
scarifcation, benching)
3. Topsoil/Subsoil
Stockpiling
3. Embankment Crest
and Benches
a. Alignment and
Grade
b. Evidence of
Movement (cracks,
sloughs, scarps)
c. Abutments
(benching, drainage
control, erosion)
d. Crest low point
elevation
4. Foundation Preparation
a. Removal of
Unsuitable Material
b. Collection of Springs
c. Underdrain
Installation
d. Benching, Cut-off or
Keyway Const.
e. Sealing of Mine
Entry Bulkheads and
Barriers
4. Embankment Slopes
a. Alignment and
Slope
b. Seepage (fow
rates, clarity)
c. Slope Movement
(scarps, sloughs,
bulging)
d. Erosion
e. Vegetation
f. Burning or
combustion
5. Liner Construction
(material, thickness,
compaction, cover)
a. Soil
b. Geomembrane
c. Geosynthetic Clay
Liner
d. Other
5. Liner, Foundation and
Abutment Conditions
a. Drainage Control
b. Erosion
c. Concentrated
Seepage or Boils
< PREVIOUS VIEW
12-10
Chapter 12
MAY 2009
Item
General
Condition
Changes
from
Previous
Inspection
Item
General
Condition
Changes
from
Previous
Inspection
FACILITY DEVELOPMENT
DISPOSAL OPERATIONS
AND MAINTENANCE
6. Starter Dam
Construction (material,
lift thickness, com-
paction, scarifcation,
benching and keying,
6. Internal Drains
(discharge fow rate
and clarity, drainage
away from outlet)
7. Internal Drains
(material, flter,
placement, covering)
7. Decant Pipe
a. Inlet (elevation,
infow depth, debris,
clogging, alignment,
anti-vortex plate)
b. Outlet (discharge
fow rate and clarity,
debris, erosion,
drainage away from
outlet)
c. Structural Concerns
8. Decant Pipe or
Spillway
a. Installation (material,
joint connections,
trench preparation,
bedding, backfll,
reaction blocks,
pressure testing,
cover)
b. Inlet (riser material,
joint connection,
reaction block and
foundation, backfll,
trash rack and anti-
vortex plate)
c. Outlet (bank
stabilization, stilling
basin, animal guard)
8. Channel Conditions
a. Alignment and
Dimensions
b. Debris and Erosion
c. Lining System
d. Receiving Stream
Condition (erosion)
9. Channel Construction
a. Alignment and Grade
b. Dimensions
c. Lining (material, sub-
base flter, thickness)
d. Transition and
Superelevation
Sections (stationing,
dimensions)
9. Instrumentation
Monitoring
(piezometers, weirs,
other)
TABLE 12.1 TYPICAL OBSERVATION AND INSPECTION CHECKLIST
(Continued)
< PREVIOUS VIEW
12-11
Monitoring, Inspections and Facility Maintenance
MAY 2009
Item
General
Condition
Changes
from
Previous
Inspection
Item
General
Condition
Changes
from
Previous
Inspection
FACILITY DEVELOPMENT
DISPOSAL OPERATIONS
AND MAINTENANCE
10. Instrumentation
(survey monuments,
piezometers, weirs,
other)
10. Reclaimed Slopes
(vegetation
development and
control, drainage
control, erosion,
seepage)
11. Reclamation
a. Grading and
Stabilization
b. Soil and Topsoil
Cover
c. Revegetation
(fertilizing, seeding,
mulching)
d. Seepage Control
(collection,
conveyance,
treatment)
11. Impoundment and
Slurry Deposition
a. Abutment Slopes
(erosion, instability)
b. Water Level and
Control (decant
inlet, pumping,
other)
c. Slurry Line
Alignment and
Discharge Location
c. Fines Deposition
(location, distri-
bution, delta areas,
percent exposed/
above pool level)
ADJOINING AREAS
1. Foundation and
Abutment Conditions
2. Upstream Watershed
3. Downstream
Inundation Area
OTHER DEVELOPMENT (Work unrelated to the disposal facility)
1. Within Facility
Watershed
(potential for
increased runoff or
landslides)
2. Downstream of
Facility (potential
to change hazard
classifcation)
3. Within Facility
Note: Reference should be made to a design or as-built plan drawing.
TABLE 12.1 TYPICAL OBSERVATION AND INSPECTION CHECKLIST
(Continued)
< PREVIOUS VIEW
12-12
Chapter 12
MAY 2009
materials placed, lif thickness and compaction (including density and moisture testing), preparation
(scarifcation) of compacted surfaces before placement of subsequent lifs, and benching of new lifs
into adjoining abutments or embankment slopes, etc.
Internal drains are frequently part of starter embankments, and typical observations include: mate-
rials used for the drain and flter system, installation conditions (trenched into embankment or
constructed on completed lif), placement of drain materials (susceptibility to segregation and con-
tamination during placement and spreading) including geotextiles, drain dimensions and grade, and
details for covering the completed installation to protect the drain from erosion.
12.1.2.2.3 Decant Pipes
During installation of decant pipes, the following observations should be recorded: alignment and
grade, pipe material and joint connections, corrosion protection (where applicable), installation
details (trench confguration), bedding and backfll (materials used and compaction, including
density and moisture testing results), seepage control measures (location, dimensions and mate-
rials), location and confguration of reaction blocks at changes in alignment and grade, results
of pressure testing, and details related to covering of the completed installation and protecting
the conduit from damage by heavy equipment trafc. The adequacy of placement of backfll in
the haunch area should be given special atention. Details related to the following should be
observed for decant inlets: riser confguration and materials, joint connections, foundation and
reaction block, riser backfll and compaction, inlet trash rack and anti-vortex plate. For decant
outlets details related to the following should be noted: receiving channel bank stabilization or
stilling basin, drainage control to preclude backup of water into pipe, and animal guard.
12.1.2.2.4 Channel Construction
Channels may include emergency spillways, decant outlets, diversion ditches, groin or perimeter
ditches, haul and access road ditches, and bench guters. Prior to construction, the upland areas
should be observed for evidence of instability that may adversely impact planned excavations. Impor-
tant observations to be recorded for channels during construction include: alignment and grade,
dimensions (width, depth, side slopes), erosion-protection lining (sub-base, flter, lining thickness
and material), and details of transition and superelevation conditions at changes in alignment, slope
or section (stationing and dimensioning).
12.1.2.2.5 Instrumentation
Instrumentation is sometimes installed during starter embankment construction. Additional
instrumentation is also generally installed during facility expansion. This instrumentation may
include survey monuments, piezometers, weirs, and other devices, as discussed in Chapter 13.
Important observations to be recorded may include: instrument type, material, make and model;
installation details (materials, dimensions); the results of initial calibration and verifcation of
function; and the results of periodic testing to verify continued performance. Maintenance should
also be regularly performed as required by the type of instrument (Section 13.5). For example, peri-
odic fushing of piezometers, if subject to fnes accumulation, and regular cleaning of weir boxes
are important.
Instruments need to be read in a consistent fashion such that the results are not afected by the indi-
vidual taking the reading and are directly comparable. If necessary, specifc instructions should be
provided relative to collecting and recording instrumentation data.
12.1.2.2.6 Reclamation
Any areas that are covered with soil and/or vegetated should be noted by location and include
a description of the work undertaken. Where areas are covered, the source of the soil or topsoil,
< PREVIOUS VIEW
12-13
Monitoring, Inspections and Facility Maintenance
MAY 2009
the thickness of the cover, and the method used to place and/or compact the material should be
noted. Where areas are vegetated, the type of seeding used, the method of placing the seeding, any
amendments added, and procedures used to protect the vegetation, such as mulch or straw should
be noted.
12.1.2.3 Adjoining Area Conditions
Areas adjoining a refuse disposal embankment include downstream areas, embankment abutments
and upstream areas. Each of these areas should be carefully observed for the presence of conditions
that could impact embankment stability, such as concentrated seepage, and other aspects of facility
operation and safety, as discussed in the following subsections.
12.1.2.3.1 Sloughing or Sliding
Stability of an embankment is dependent not only upon conditions within the embankment itself, but
also upon the embankment foundation materials. Therefore, areas immediately upstream and down-
stream from any embank ment should be inspected for bulging conditions, open cracks with vertical
displacement or other signs, such as the orientation of cracks, that could indicate movement of the
soil or rock materials is occurring. In wooded areas, a possible means of identifying the presence of
such conditions is to observe the trunks of the trees to determine if a group of trees has an unusual
inclination. If such signs are evident immediately adjacent to an embankment, the embankment sur-
face should be examined very carefully to see if the movement has extended into the embankment.
12.1.2.3.2 Signs of Subsidence
Evidence of mine subsidence can include surface cracks (particularly when oriented with mine work-
ings and/or vertical displacement), depressions, and sinkholes. Ofen evidence of subsidence due to
underground mining is more readily apparent on natural ground surfaces than on an embankment
surface, because the embankment can arch over a subsided area for some period of time or the sub-
sidence features can be hidden by disposal operation. Subsidence that occurs on the natural landform
can cause rapid changes in contours where bedrock is shallow or sinkhole development where water
washes material into the mine. Also, in areas where subsidence is not uniform, the inclinations of tree
trunks can vary signifcantly over short distances. If possible signs of subsidence are observed, mine
maps should be reviewed to determine if there are mine works present that could be causing the
problem. Subsidence may also occur over mines that are unmapped or incompletely or inaccurately
mapped. In many cases, small punch mines have been developed for residential coal consumption
and may not be mapped, as is done for commercial coal mines. The lack of availability of maps
should not be considered as evidence of the absence of underground mines.
12.1.2.3.3 Evidence of Erosion
Conditions where erosion of the natural ground surface occurs downstream or downslope from
an embankment, or erosion of the embankment itself, can be critical and includes cases where: (1)
an adjacent stream rises during a food causing erosion of the embankment toe or (2) the drain-
age system from areas upstream of the embankment discharges onto or immediately adjacent to
the embankment causing downcuting and eventual instability. The locations of any areas of sig-
nifcant increase in erosion of the embankment or the adjacent ground surface should be noted.
12.1.2.3.4 Springs or Seepage Areas
Instability of an embankment slope due to high pore-water pressures or piping of soil materials
can occur in the embankment foundation and abutments as well as on the downstream face of the
embankment. Therefore, it is important to observe the downstream slope and area downstream
from the abutments and toe of an embankment and to note development of new seepage areas
or changes in areas where seepage was previously occurring. In particular, it should be deter-
< PREVIOUS VIEW
12-14
Chapter 12
MAY 2009
mined whether existing seepage zones have increased in fow volume since the last observation
and whether the seepage is carrying fnes, which is an indication of internal erosion.
The presence of boils (points where a concentrated upwelling of seepage is creating a deposit of
fne material) is an especially important indicator of internal erosion. Records of spring and seep-
age area fow volume should be kept for comparison purposes. Signifcant changes in seepage or
evidence of internal erosion should be brought to the atention of the Engineer, because corrective
measures to prevent instability, progressive internal erosion, or piping may be warranted. Seepage
and water quality monitoring (e.g., temperature, turbidity, and total dissolved solids) and interpre-
tation of trends are discussed in Pelton (2000).
12.1.2.3.5 Changes in Vegetation
As noted previously, changes in the inclination of tree trunks can indicate sliding and sloughing of
the natural ground materials or subsidence due to underground mining. Also, areas that have either
an abnormal amount of dying vegetation or unusually dense vegetation may indicate the presence of
seepage zones that are not immediately evident at the ground surface.
12.1.2.4 Other Development Work
12.1.2.4.1 Upstream Development
Any changes in the watershed area above the facility, such as tree cuting, strip mining, and residen-
tial or industrial development should be noted, since these activities could afect runof fows to the
site. Also, these activities may afect the quality of surface water passing through the site, which could
mistakenly be atributed to the refuse disposal facility. The location and extent of upstream changes
and the nature of such changes should be documented and brought to the atention of the Engineer.
12.1.2.4.2 Downstream Development
Inundation mapping associated with dam breach analysis for establishing the hazard potential classi-
fcation is based upon topography and stream conditions downstream from the refuse impoundment.
Future development in the downstream foodplain could increase the hazard potential rating. Also, a
new mine opening or coal preparation plant constructed downstream from the facility spillway might
be susceptible to fooding during large storms. Any such change in downstream land use, including
the location and description of the change, should be brought to the atention of the Engineer.
12.1.2.4.3 Disposal of Materials Other than Coal Refuse
The design of critical portions of refuse disposal facilities is typically based on the assumption that
uniform materials will be used. The use of materials from other sources (e.g., power plant waste) or
from site development work unrelated to the disposal facility, such as placement of coarse-grained
rock from slope excavations in the disposal facility embankment, could cause undesirable behavior
and should not be permited unless specifcally allowed in the approved design plan. Material that
could lead to combustion in the coal refuse embankment is not permited.
12.1.2.5 Refuse Disposal Operations and Maintenance
12.1.2.5.1 Refuse Placement and Compaction
Disposal operations comprise the placement of refuse materials within embankments in accordance
with the plans and specifcations. Important observations to be documented include: materials (gra-
dation and moisture), delivery to site and spreading (location and equipment), lif thickness, com-
paction (equipment, number of passes), surface conditions (frm, pumping), surface preparation or
scarifying (method or equipment, extent and frequency), and benching of new lifs into embank-
ment and abutment materials (dimensions and extent). During inclement weather, it can be difcult
< PREVIOUS VIEW
12-15
Monitoring, Inspections and Facility Maintenance
MAY 2009
to meet construction specifcations for refuse placement and compaction, particularly in structural
zones, and documentation of potential problem areas is important to the planning of future actions.
Where feasible and appropriate, other zones or areas that can accommodate coal refuse where place-
ment standards are lower, such as during inclement weather, may be provided as part of the design.
Monitoring and inspection of these areas typically include observation of lif thickness, grading,
access, and drainage control.
12.1.2.5.2 Working Surface Maintenance
The working surface where active refuse disposal is taking place should be maintained consistent
with the plans and specifcations and with careful control of drainage. Observations should include:
(1) survey control such as stakes or markers to defne placement limits and grades, (2) lif placement
and compaction, (3) scarifcation of the surface of layers that are too smooth to properly bond to
the next layer, and (4) drainage control measures for directing water away from active placement
areas and toward collection ditches. Density and moisture content testing must be fully documented
including the location of the tests (which should generally be random or focused on areas of suspect
compaction), the lif thickness, and the test results. Compaction conditions should be further evalu-
ated if the test results show that the compaction specifcations are being met, but the lif surfaces
afer compaction remain sof or spongy. If the compaction specifcations are not met, then additional
compaction will be required and, if this situation occurs frequently, the cause should be determined.
Adequate compaction requires the proper lif thickness and number of passes of the compaction
equipment be used. The type and number of passes required for achieving the minimum compaction
specifcations will depend on the compaction equipment and lif thickness. When new or diferent
compaction equipment is used, the number of passes needed to meet the compaction specifcations
needs to be established by a coordinated program of feld density testing. The compaction operation
should be observed to verify that lifs are being systematically compacted at the proper thickness
with overlapping passes of the compaction equipment.
12.1.2.5.3 Embankment Crest Conditions
The crest of a refuse embankment should be constructed and maintained in conformance with
the facility plans and specifcations and should be generally level (or with a slight slope towards
the impoundment) to provide adequate freeboard and drainage control. Important observations
include: (1) alignment and level, (2) indications of movement (e.g., cracks, sloughs, scarps), and
(3) abutment interface (including benching to tie refuse materials into natural hillsides, control of
drainage from hillsides, and control of erosion).
Coal refuse tends to break down when exposed to the elements and to equipment trafc. When a
crest elevation is maintained for a period of time, a layer of highly compacted and weathered mate-
rial may be present at the surface. Prior to raising the crest, this material should be disturbed by
thorough scarifcation or should be removed so that it does not create a continuous horizontal layer
of low hydraulic conductivity within the embankment. Such a layer can cause elevated seepage
levels in the embankment and defeat the purpose of internal drainage measures.
12.1.2.5.4 Embankment Slope Conditions
The slopes of a refuse embankment should be developed and maintained in conformance with the
facility plans and specifcations. Temporary slopes should be consistent with the geometry and over-
all development shown in the design report. Documentation should include alignment and steep-
ness, seepage, movement, and erosion. The following conditions should be noted, as applicable:
Sloughing and sliding All embankment slopes should be carefully inspected for the
possible presence of sloughing and/or sliding. Such conditions are usually accom-
< PREVIOUS VIEW
12-16
Chapter 12
MAY 2009
panied by bulging at or along the botom of the distressed area and possibly the
formation of a vertical displacement (scarp) and cracking at or along the top of the
distressed area. Special care is required in areas of heavy vegetation to avoid over-
looking these areas of distress. Heavy vegetation that hampers inspection activities
should be cut or removed, as should any trees that begin to grow on the embank-
ment. Particular atention should be paid to the downstream slope of an impounding
embankment. Any signs of sliding or sloughing should be reported immediately to
the Engineer and Operator.
Subsidence Subsidence of an embankment can occur as a result of three conditions:
(1) the collapse of voids in the embankment or foundation that form due to piping,
when seepage fows remove particles of refuse or soil, (2) consolidation of sof
foundation or embankment materials, and (3) collapse of mine workings. Subsidence
due to piping can usually be recognized by the presence of localized bowl-shaped
depressions at the surface. Subsidence due to setlement of sof foundation materi-
als usually results in low areas on the embankment surface and occasionally cracks.
Subsidence due to under ground mining can be identifed by cracking along straight
lines, typically parallel to the primary direction of mining, or as deep circular sink
holes. When vegetation is tall, care should be taken to avoid overlooking such signs
of distress. If subsidence is discovered, the size, shape and extent of the distress
should be determined and recorded. Any evidence of subsidence should be reported
immediately to the Engineer and Operator.
Erosion Erosion is loss of ground cover caused by a lack of proper drainage
control. Sheet erosion is a surfcial loss of soil or refuse in a more or less uniform
patern extending over wide areas. Gully erosion is the result of concentrated
overland fow and leads to the creation of narrow and possibly deep channels in
the ground surface. Any occurrence of erosion should be documented as to type,
location, and extent. Erosion should be corrected when it starts to develop and
before it becomes signifcant. Erosion that has progressed to deep gullies can afect
seepage and stability and is much more difcult to correct. If erosion is found to
concentrate or recur in particular areas, consideration should be given to improv-
ing surface drainage in those areas.
Moisture conditions, springs, seeps and wet areas Slopes should be monitored
for the presence of springs, seeps and wet or boggy areas. Evidence of such condi-
tions includes the presence of unusually wet and/or sof areas and areas where
snow melts rapidly in the winter. Also, vegetation may be thriving due to the moist
conditions, leading to unusually dense and taller growth as compared to surround-
ing vegetation, or vegetation may be adversely afected by acidic seepage and may
be discolored or dying. The location and the extent of such conditions should be
documented. If actual fow is observed, the rate of fow should be measured or
estimated, and the quantity and character of any suspended solids (silt, sand, etc.)
carried by fow should be noted. The Engineer and Operator should be notifed
immediately. The fow should be measured regularly, and the records should be
maintained for comparison purposes. Particular atention should be paid to the
downstream face of impounding embankments, as springs, seeps and wet areas can
lead to sloughing and sliding.
Condition of vegetation All seeded areas should have a well-developed vegetation
cover. The vegetation should be uniform and continuous. Any irregularities such as
diference in color, density, rate of growth, type of growth, or a diference in the char-
acter of the vegeta tion should be noted. For example, catails are a common indicator
< PREVIOUS VIEW
12-17
Monitoring, Inspections and Facility Maintenance
MAY 2009
of wet areas. Such irregularities could be evidence of other problems. The location
and nature of any irregularities in vegetation should be recorded.
Trees The roots of trees can provide paths for seepage and, if the roots decay,
voids may be created. Trees that are blown over during a storm can also damage an
embankment. For these reasons, trees should not be permited to become established
on the crest or slopes of a refuse embankment. Guidance on maintenance of vegeta-
tion and embankment repair following removal of trees is presented in Marks and
Tschantz (2002).
12.1.2.5.5 Liner, Foundation and Abutment Conditions
Liner, foundation and abutment conditions should be observed for signs of deterioration, inadequate
drainage control, erosion or instability prior to placement of refuse. Along rock abutments, rock over-
hangs should be removed so that fll can be adequately compacted against the abutment. The treat-
ment of open discontinuities in the foundation/abutments should be addressed in the specifcations.
If such foundation features are observed and are not addressed in the specifcations, they should
immediately be brought to the atention of the Engineer.
12.1.2.5.6 Internal Drains
The outlets of internal drains should be observed to: (1) determine discharge rate and clarity, (2)
verify that no blockage restricting discharge is present, and (3) verify that drainage is directed away
from the outlet and not backing up into the drain. In addition to this monitoring, fow rates should
be measured, recorded, ploted, and regularly reviewed by the Engineer and should be compared
to other facility data such as impoundment pool level, rainfall, and the presence or absence of a fne
refuse delta.
12.1.2.5.7 Decant Pipe
The inlet of the decant pipe should be observed relative to infow depth and rate, alignment, pres-
ence and condition of a trash rack and anti-vortex plate, and verifcation that there is no blockage by
debris. The outlet of the decant pipe should be observed relative to discharge rate and clarity, pres-
ence of debris, erosion conditions, seepage along the outside of the pipe, and verifcation that the
discharge is directed away from the outlet without backup into the pipe. Where exposed, the pipe
wall should be observed for distress. The frst time that decants fow and when they are subject to
signifcant increases in head are important times to observe performance for possible indications of
problems. During decant pipe observations, or as part of scheduled inspections, the following should
be evaluated:
Clogging A common mode of failure of decant systems that is very difcult to cor-
rect is clogging of the intake or pipe with miscellaneous debris. Most designs include
a trash rack around the intake to block material that cannot easily be carried through
the pipe. The buildup of debris on and adjacent to the trash rack should be cleared,
and the removal should be documented. Proper equipment and access should be
provided so that any debris can be safely removed. Any damage to the trash rack
or the presence of debris that cannot easily he removed from the intake should be
reported.
Characteristics of discharge Where possible, an evaluation should be made of the
character and volume of water fowing into the decant inlet and out of the decant
discharge. If the appearance of the water at these two locations difers, it is possible
that cracks or open joints in the decant pipe are allowing seepage fow to enter the
pipe or decant fow to exit the pipe. If this situation is observed, it should be reported
< PREVIOUS VIEW
12-18
Chapter 12
MAY 2009
immediately to the Engineer. The presence of seepage fowing along the outside of
the conduit at the downstream end of the decant should also be noted.
Corrosion, cracking or crushing Loss of any portion of the decant system can afect
continued operation of a refuse disposal facility. For this reason, areas of distress
to the decant system due to corrosion of materials, cracking or crushing should
be reported immediately so that appropriate repairs can be made. Special inspec-
tion procedures and equipment may be needed for observation of the condition of
a buried decant pipe. Typically, a decant should be treated as a confned space and
only entered with appropriate precautions, including:
Initially checking and then regularly monitoring of the atmosphere
inside the conduit for safe conditions and air movement (e.g., a respi-
rable, non-explosive atmosphere).
Providing for continuous communications with personnel inside the
conduit.
Ensuring that personnel entering the conduit have appropriate per-
sonal safety equipment.
Providing for safe ingress and egress, including harnesses and life-
lines where appropriate.
Providing workers with training with respect to potential hazards
and required safety measures.
Track-mounted video cameras that can travel inside the decant pipe are commonly
employed to avoid entry of personnel and in situations where pipe diameters are
small. Inside-diameter measurements can be obtained using optical/laser equipment
coupled to a track-mounted video camera.
Erosion conditions Erosion at the decant intake is not generally important to the
integrity of the decant system, but is an indication that fne refuse particles are
being transported into the decant pipe. These materials may setle out at low points
or bends and eventually reduce the decant pipe capacity. Erosion at the discharge
point can be more critical if it eventually undercuts the outlet causing a break in the
pipe or undercuting of the toe of the adjacent embankment creating local stability
problems. Seepage discharging along the exterior of the decant should be noted and
observed for evidence of fnes carried with the fow (an indication of internal erosion
along the conduit). Any of these conditions should be reported to the Engineer and
Operator.
12.1.2.5.8 Channel Conditions
Observation of channels associated with spillways, ditches and guters should include: (1) alignment
and dimensions (depth, width and side slopes), (2) evidence of debris, (3) erosion and condition of
erosion protection, and (4) the condition of receiving channels, ponds, or streams. The following
observations should be documented:
Erosion The banks and the beds of all channels should be examined to determine
if erosion or siltation is occurring. Channels should have a uniform confguration
consistent with construction plans. Erosion of banks is evidenced by side slopes that
are steeper than shown on the plans or by sloughing or localized irregularities in
confguration. The location of any eroded areas should be recorded. The location of
the spillway discharge should be checked to see that it is far enough downstream
that the embankment toe is not being eroded.
< PREVIOUS VIEW
12-19
Monitoring, Inspections and Facility Maintenance
MAY 2009
Undergrowth The channel cross section should be clear of undergrowth such as
brush and small trees from the top of one bank to the top of the opposite bank. Such
undergrowth decreases the efciency of the channel and traps foating debris during
high fows. Areas of excessive undergrowth should be cleared.
Flow obstructions Any obstructions to fow in a channel such as fallen trees,
sloughed-in soil or rock from adjacent excavated slopes, or other foreign objects
or excessive vegetative growth should be noted and removed at frst opportunity.
Where signifcant sloughing of a channel slope is occurring, engineering personnel
should be notifed immediately.
Structure conditions The condition of all structures that are part of, or appurtenant
to, the channels should be noted, including culverts, retaining walls, wing walls, rock
cuts, pavements, slabs, etc. Culverts should be maintained in good condition, free
of debris and with unobstructed inlets and outlets. Walls, slabs, rock cuts and pave-
ments should be checked for evidence of distress such as setlement or other move-
ments, cracks, scour and erosion. All areas requiring improvement or maintenance
should be documented.
Riprap is commonly used in channels and elsewhere for erosion protection. Observations of riprap
should include the following:
Irregularities and displacements The surface of riprap areas should be examined
for abrupt changes in slope or alignment and for gaps or cracks in the surface. Areas
where the riprap is missing or has otherwise been disturbed should be noted. Local-
ized changes in the character of the riprap material should also be recorded. If the
deterioration of riprap is extensive, the Engineer and Operator should be notifed.
Sizes The maximum and minimum sizes of the riprap should be examined to verify
that they are generally in accordance with the plans and specifcations. The distri-
bution of riprap sizes should be uniform. Any observed segregation of riprap with
respect to size should be noted. Irregularities in the size of the riprap could indicate
that deterioration due to weathering or scouring is occurring.
Weathering Weathering is the deterioration of the riprap due to freezing and thaw-
ing cycles or due to chemical action in the presence of water, air, sunlight, etc. Weath-
ering of riprap is usually indicated by a lack of larger-sized pieces, since weathering
acts to break the larger pieces into smaller ones. If weathering of the riprap has
occurred, the location and extent of the weathering and the largest remaining size
should be recorded.
Scouring Scouring of riprap is caused by the efects of fowing water or wave
action. The mechanical action of the water acts to segregate the riprap and carry
smaller-sized pieces away. Scouring is usually indicated by a lack of smaller pieces.
If scouring has occurred, the location and extent of the scouring and the size of the
smallest remaining pieces should be recorded.
Grouted riprap At some locations, the riprap surface may be grouted to a certain
depth with fne aggregate concrete. Disturbances to the riprap, such as setlement,
undermining, erosion and scour can cause deterioration of the concrete grout. Any
areas where the concrete grout has deteriorated, or areas where it is missing entirely,
should be recorded.
Concrete is frequently used in the construction of spillways, channels and other hydraulic appurte-
nances. Distress to concrete structures can result from many factors ranging from simple deteriora-
< PREVIOUS VIEW
12-20
Chapter 12
MAY 2009
tion with time to movements of the structure foundation. Since the type of repair is dependent upon
the cause, observed conditions should be recorded so that appropriate repair and maintenance pro-
cedures can be implemented.
Cracking The surface of all concrete structures and other concrete appurtenances
should be inspected for the presence of cracks. Care should be taken to not over-
look hairline crack paterns. The size, location and confguration of observed cracks
should be recorded. If cracks are noted on a particular structure, other similar parts
of the same structure should be carefully examined.
Spalling The surface of all concrete structures and other concrete appurtenances
should be examined for the presence of spalling. Spalling is indicated by the removal
of the concrete matrix at the surface. The remaining surface texture will be rough, and
the aggregate will be exposed. In general, two types of spalling can occur. The frst is
a surfcial or uniform deterioration of the concrete surface. The second type is a deep
spalling that exposes the reinforcing bars. If spalling is observed, the type should be
noted, and the location and the extent of all spalled surfaces should be recorded.
Condition of joints All construction joints, expansion joints and contraction joints
should be examined to determine if relative movement at the joints has occurred. If
there is evidence of movement at any joint, the size and uniformity of the gap should
be noted. Variation in size and width of the gap may indicate the relative movement
of the structure, and observation of changes in the condition with time will indicate
if further movement can he expected.
Leaks and seepage All drains and weep holes in concrete structures should be
examined to verify that they are free from debris and are functional. Ofen water
fowing through improperly functioning drains and weep holes will wash backfll
material or flter material from behind a wall and deposit it in front of a weep hole. If
this condition is observed, it should be noted. All concrete surfaces should be exam-
ined for the presence of damp or wet surface areas. If damp or wet surface areas
are found, their size, extent and location should be noted. Any water seeping from
cracks or joints in any concrete structure should be noted.
Setlement or heave The vertical alignment of all concrete structures should be
checked for irregularities. This can best be accomplished by sighting along horizontal
lines of the structure. All structural lines should be straight. Setlement is indicated by
a dip or depression. Heave is indicated by an elevated section or hump. If setlement
or heave occurs over a substantial distance, it may be difcult to detect. Such large-
scale movements are normally associated with relative movement at one or more of
the joints in the structure. If it is expected that setlement or heave is occurring over
an extended distance, the condition of the joints in the structure should be carefully
observed to determine where the movement is occurring. The location of and an esti-
mate of the magnitude of any setlement or heave should be recorded.
Defection and lateral movements The horizontal alignment of all concrete struc-
tures should be checked for irregularities. This can best be accomplished by sighting
along vertical lines of the structure. All vertical lines should be plumb. Localized
lateral movements will be indicated by bulges in the walls and/or misalignment of
the structural elements. Relative displacement at the vertical joints may also occur. If
such displacements are observed, the location and an estimate of the magnitude of
displacement should be recorded.
Tilting Tilting is indicated by general rotation of a structural element. Tilting may
ofen be combined with setlement or lateral movement. If tilting occurs, vertical ele-
< PREVIOUS VIEW
12-21
Monitoring, Inspections and Facility Maintenance
MAY 2009
ments such as walls will be out of plumb and horizontal elements such as slabs will be
out of level. Such a condition is probably most easily detected by viewing the structure
from a distance. Tilting of water conveying channels is indicated by a tendency for the
depth of fow to be greater at one side of a channel. Tilting of structures such as retain-
ing walls and chute walls may be accompanied by a gap between the structure and
backfll. Periodic surveying should be performed when tilt or movement is suspected.
Undermining Concentration of rainfall runof along, around and behind con-
crete structures can cause erosion that can lead to undermining. This is particularly
relevant for paved guters and channels. Major seepage or leaks under or behind
concrete structures can also cause undermining. Undermining may be indicated
by holes or voids under or behind the structures or by localized setlements or
depressed areas on the surface of the backfll around them. Care should be taken not
to miss evidence of undermining that may be obscured by vegetation.
Condition of backfll The condition of the backfll around and behind all concrete
structures should be examined for evidence of erosion, setlement and/or lateral
movement. If the backfll was seeded, the condition of the vegetation should be
noted. The surface of the backfll should be examined for the presence of depres-
sions or small holes. Lateral movement of the backfll may be accompanied by a
gap between the concrete structure and the soil, cracks in the surface of the backfll
running generally parallel to the main lines of the concrete structure, or vertical steps
or scarps in the surface of the backfll. Particular atention should be paid to backfll
along and behind spillway chute walls.
12.1.2.5.9 Instrumentation
The overall condition of instrumentation such as fow monitoring weirs, survey monuments and
piezometers should be determined by visual inspection. For example, weirs should be inspected to
verify that they are in good condition and that erosion has not created bypass channels; survey mon-
uments should be observed to verify that they are still in place and in good condition; and piezom-
eters should be checked to verify that they are in good physical condition and operable. Observed
damage to instrumentation should be immediately reported to the Engineer and Operator and the
instrumentation should be repaired or replaced as necessary. Malfunctioning or damaged instrumen-
tation needs to be repaired or replaced, unless there is a technical basis for its complete elimination.
Instrumentation that is included in the construction plans and specifcations should be maintained
unless removal is approved by the appropriate regulatory agencies. Should supplemental instrumen-
tation be recommended by the Engineer or implemented by the Operator, the purpose and planned
frequency and duration of monitoring should be documented and submited to MSHA.
12.1.2.5.10 Reclaimed Slopes
Slopes and other completed surfaces that have been reclaimed should be observed with respect to
movement, vegetation development, drainage control, erosion and seepage. In addition to safety of
the refuse disposal facility, they are important with respect to abandonment of the facility and envi-
ronmental control. A primary reason for noting conditions in these areas is to verify that the type of
covering and vegetation procedures being used are adequate and will result in the desired fnished
embankment. Observations should include: (1) notation of the uniformity of vegetation, (2) areas
where the vegetation cover is apparently not surviving or is very sparse, and (3) areas where crack-
ing or erosion of the soil cover is afecting its function of protecting the embankment or other surface.
12.1.3 Quality Control Sampling and Testing
In addition to visual observation, feld quality control sampling and testing provides a means to
verify that construction of refuse disposal facility components is being performed in conformance
< PREVIOUS VIEW
12-22
Chapter 12
MAY 2009
with the plans and specifcations. Table 12.2 presents a summary of typical quality control tests and
guidelines for testing frequency associated with refuse disposal facility construction. Designers may
wish to incorporate alternate testing requirements and/or frequencies to refect site- or design-spe-
cifc facility conditions. For example, more tests may be performed initially to establish the variability
of properties.
12.2 MONITORING INSTRUMENTATION
Measurements should be routinely recorded as part of the monitoring and inspection program for
each coal refuse disposal facility. Most instruments that require monitoring are specifed and installed
as part of the normal design and operation of the refuse disposal facility. Supplemental instrumenta-
tion may be installed in response to unexpected conditions. The exact types and location of instru-
ments should be specifed in the plans and specifcations and the Operation and Maintenance Plan
for the facility. For most facilities, monitoring instruments typically include survey monuments,
piezometers, monitoring wells, and weirs or fumes. A method for monitoring the pool/slurry level is
usually also provided. More complex instruments are sometimes required during construction and
operation of a coal refuse disposal facility for monitoring and more closely observing selected areas
or conditions. A thorough discussion of coal refuse facility instrumentation is provided in Chapter
13. A brief summary of typical instrumentation at coal refuse disposal facilities is provided in the fol-
lowing paragraphs:
Monument surveys The use of monument surveys along with as-built topographic
surveys based on aerial photography is an efective method for monitoring both
displacements and confguration of the facility. Monument surveys can detect both
horizontal and vertical movement of an embankment, and it is recommended that an
active facility should be surveyed by a licensed surveyor at least annually. Annual as-
built topographic surveys provide an efective basis for verifying that construction is
progressing in accordance with the design plans and for preparing reports to regula-
tory agencies concerning the amount and extent of construction. They also provide
a means of confrming refuse production rates and progress on individual embank-
ment stages. More frequent as-built surveys for facilities during initial development
may be important for verifcation that critical embankment heights and geometries
are being achieved. Both monument surveys and as-built topographic surveys are
limited in their accuracy, and smaller movements may not be detected. If displace-
ments are suspected, additional instrumentation may be needed.
Pool-level gauge For slurry impoundments, a method is needed to monitor the
level of the pool and the slurry delta. This information is needed in order to verify
that the design provides adequate freeboard and slurry disposal capacity. The pool
level can be determined based on visual observations at a staf gauge and delta
monitoring conducted by periodic elevation surveys relative to established bench-
marks near the upstream slope of the refuse embankment. Also, soundings can be
performed to determine setled fne refuse levels within the pool area.
Rainfall gauge A rainfall gauge should be located close to an impounding refuse
embankment so that daily precipitation can be monitored and compared to piezom-
eter records or seepage rates observed at weirs. This is particularly important where
underground mine workings are close to an impoundment and monitoring of mine
discharge is being performed. In such cases, rain gauge data allow increases in mine
discharge due to rainfall to be distinguished from increases caused by impoundment
seepage or leakage.
Piezometers Piezometers are normally installed at impounding facilities for moni-
toring the phreatic level or pore pressures in a refuse embankment for comparison
< PREVIOUS VIEW
12-23
Monitoring, Inspections and Facility Maintenance
MAY 2009
to projected design levels used in stability analyses. Piezometers are also used for
monitoring pore pressures during upstream embankment construction. Tolerable
excess pore pressures predicted in the design can then be compared with monitored
pore pressures. The three most commonly installed types of piezometers are stand-
pipe, pneumatic and vibrating-wire. Chapter 13 provides a detailed discussion of
these types of piezometers, along with others less commonly used. Consistent with
CFR 77.216-3, instruments should be measured/read every seven days and the data
should be recorded. The actual frequency of measurement should also be based on
the historical or expected rate of change in pore pressure and the disposal facility
design features. More frequent piezometer readings are generally obtained during
upstream embankment construction. Automated data acquisition systems coupled
with vibrating wire piezometers have been used at refuse facilities to obtain more
frequent pore pressure measurements; such systems are also employed because of
their cost savings.
Monitoring wells Monitoring wells, and to a lesser extent piezometers, are used
for monitoring the groundwater quality and levels at and adjacent to refuse disposal
facilities. The monitoring or sampling frequency varies depending upon environ-
mental conditions and state regulations.
Weirs and fumes Weirs and fumes are used to measure water fow rates in channels
and from seeps and internal drains. The monitoring and measurement of fow rates
from seeps and internal drains is an important means for evaluating the performance
of the internal drainage system of an impounding refuse embankment. At sites with
underground mining, fow from mine openings may also be measured as a means for
monitoring changes in the amount of seepage into the mine, and these data may pro-
vide an indication of the perform ance of the impoundment with respect to potential
breakthrough, provided that surface runof is diverted away from the weir. Accurate
measurement of weir fow rates is important, and the frequency and methods of mea-
surement should be specifed in the plans and specifcations and the Operation and
Maintenance Plan. Flow rates are typically measured on a 7-day basis and recorded,
or less frequently if records indicate consistent measurements and conditions permit
(consistent with the regulations and approved plan). If fow rates are relatively low,
they can be measured using a standard bucket or calibrated container and stop watch.
The time required for the water fow to fll the container is recorded, and a fow
rate is calculated. Evaluation of fow data to determine whether they are normal or
unusual will typically require correlations with pool level and rainfall. Flow-related
data should be ploted on a common time-line, along with pool and piezometer data,
so that trends in the data can be identifed and fow trends can be evaluated to verify
that they are consistent with expectations. To accurately correlate seepage fows with
rainfall, measurements may need to be performed on a daily basis.
Blast/vibration monitoring Drilling and blasting activities can cause shock vibra-
tions that can afect a structural component of a refuse disposal facility or cause
displacements due to liquefaction of embankment foundation materials. Common
instruments for monitoring shock vibrations are commonly referred to as accelerom-
eters or velocity transducers. The deter mination of the equipment needed for vibra-
tion measurement requires the input of an experienced engineer who is familiar with
the feld data needed for analysis of the possible efects of vibration and dynamic
displacement on facility components.
Conduit monitoring Deeply buried conduits may require monitoring during the life
of the facility. Small video cameras mounted on a mobile unit that can travel through
the conduits are commonly used to inspect the interior condition. Measurements of
< PREVIOUS VIEW
12-24
Chapter 12
MAY 2009
defection can be obtained with various measuring devices such as laser or optical
equipment using the laser ring method. This method employs a non-contact laser that
is projected onto the interior of the pipe and viewed by a video camera. Sofware is
then used to calculate the interior dimensions of the pipe. The need and frequency of
conduit monitoring is a function of the pipe design (including tolerable deformations)
and site-specifc parameters such as foundation conditions, depth of cover, observed
performance, conduit extension schedule, and other aspects of the facility.
Extensometers/setlement gauges/inclinometers These instruments, as described in
Chapter 13, may be used where there is potential for deformation of the foundation
or abutment, particularly where mine subsidence is present. The need and frequency
of monitoring of these instruments should be determined on a site-specifc basis,
and ofen is related to performance monitoring of other instruments such as surface
monuments.
Temperatures If necessary, instrumentation can be installed in a coal refuse
embankment to monitor the internal temperature. Temperature monitoring instru-
ments vary from conven tional thermometers to telethermometers. Telethermometers
are sometimes coupled with a vibrating wire piezometer in automated data collec-
tion systems.
Instrumentation data and measurements discussed above should be regularly reviewed and should
become part of the annual dam safety inspection report required by some states and the Annual
Report and Certifcation required by MSHA.
12.3 FACILITY AND DAM SAFETY INSPECTIONS
Facility and dam safety inspections provide a basis for identifying maintenance needs, for evaluating
conformance of refuse disposal facilities with the plans and specifcations developed for their con-
struction, and for detecting signs of instability or other safety-related problems with the performance
of the facility. It is recommended that impounding facilities have in place dam safety inspection pro-
grams including: (1) routine dam surveillance by the facility operator, (2) dam safety inspection by
the Engineer, and (3) dam safety review (including evaluation of continued disposal facility develop-
ment requirements) by an engineering team (Szymanski, 1999). Routine embankment surveillance
should be consistent with the 7-day impoundment inspection requirement by MSHA. Dam safety
inspections should be conducted by, or under the guidance of, the Engineer and should include at
least annual inspections for evaluation of the condition of the facility and may also include routine
inspections that supplement dam surveillance. Dam safety review includes a more comprehensive
inspection and evaluation of future refuse production and disposal plans that typically is performed
as part of stage completion and initiation of new disposal facility stages.
12.3.1 Routine Inspections
The qualifcations of inspection personnel and the frequency of facility inspections are related to the
type and design of the refuse disposal facility and federal and state regulatory requirements. Inspec-
tion records should be maintained for future reference. Each inspection should involve the review
of data collected from previous inspections and from construction monitoring activities since the
previous inspection. Routine inspections are typically performed on a weekly to quarterly basis or as
specifed in the Operation and Maintenance Plan.
12.3.1.1 Impoundments and Dams
MSHA requires inspections (at intervals not to exceed 7 days or as otherwise approved by the MSHA
District Manager) of impoundments in accordance with CFR 77.216-3 and provides guidance for
these inspections in the MSHA Coal Mine Impoundment Inspection and Plan Review Handbook (MSHA,
< PREVIOUS VIEW
12-25
Monitoring, Inspections and Facility Maintenance
MAY 2009
TABLE 12.2 FIELD QUALITY CONTROL TESTING SUMMARY
Item Test Suggested Frequency Reference
Topsoil/Subsoil for Reclamation
Material Composition Nutrients, organic matter
Each area or as required by state
regulatory agency
Foundation Preparation
Foundation Soils
USBR, 1998
Suitable Material
Proof-rolling and/or feld density
and moisture (strength and
gradation on suspect material)
> 1 test per foundation material
type
Cutoff/Key Backfll
Suitable Material
Index properties, hydrometer,
standard proctor, (hydraulic
conductivity, if part of liner)
> 1 test per 20,000 cy
Compaction Compaction > 1 test per 2,000 cy
Underdrain
Suitable Material Gradation, durability > 1 test per material source
Liner Construction
Soil
USBR, 1998
Suitable Material
Sieve analysis, index proper-
ties, hydrometer, standard proc-
tor, hydraulic conductivity
> 1 test per lift per acre and as
material varies or as required
by state regulatory agencies
Compaction Compaction > 1 test per 2,000 cy or each lift
Geosynthetic
Haxo, 1986
Daniel and
Koerner,
1995
GRI, 2005
Geomembrane Strength at seams
6 per 3,300 feet of seam, if
sampled randomly, or 1 every
500 feet of seam, if sampled
on a uniform basis
Geosynthetic Clay Clay mass, strength, index fux > 1 test per material source
Structural Fill Placement
Material
Index properties, sieve
analysis
Each material and > 1 test per
20,000 cy
MSHA, 2007
Compaction
Compaction (embankment fll)
Compaction (backfll around
structures)
1 test per 2,000 cy
(at least 1 test per lift)
1 test per 200 cy
(at least 1 test per lift)
Proctor
Each material and > 1 test per 20
feld compaction density tests
< PREVIOUS VIEW
12-26
Chapter 12
MAY 2009
Item Test Suggested Frequency Reference
Placed Refuse Fill
(1)
Material
Index properties and sieve
analysis
As material varies
Compaction Compaction Dependent on facility design
Internal Drain
Aggregate
Suitable Material
Gradation and durability,
calcium carbonate
> 1 test per material source
Geotextile
Material Permittivity, strength > 1 test per material type
Decant Pipe
Material Connections Leak (pressure) testing Completed sections or entire
length
Bedding and Backfll
Material
1. CLSM or Concrete
2. Soil
Strength, air content, tempera-
ture, slump, and potentially hy-
draulic conductivity.
Index properties, hydraulic
conductivity (if penetrating
liner), proctor compaction
1. Each pour
2. Each material; >> 1 compaction
test per 200 cy (min. 1 test per
lift)
Operational Performance
or Long-Term Performance
Camera survey with defection
measurements
With major increase in embank-
ment height for fexible pipe,
depending on pipe design for
defection and strength
(2)
Channels
Riprap Material
(3)
Gradation and durability Each material
Grout Material Strength, slump Each pour
Concrete
(3) Strength, air content,
temperature and slump
Each pour
Soil Backfll Material
Index properties, proctor
compaction, compaction
Each material > 1 test per backfll
segment
Reclaimed Slope
Soil and Topsoil
Material
Thickness, nutrients
Each material or as required by
appropriate state regulatory agency
Note: 1. Placed refuse fll may be designed zones where material placement and compaction standards are typically
lower than for structural fll.
2. A baseline camera and defection survey may need to be performed upon completion of backfll and
cover for fexible pipe installations, depending on the pipe design.
3. Aggregate subbase and/or geotextile flter may be subject to similar testing as for an internal drain.
TABLE 12.2 FIELD QUALITY CONTROL TESTING SUMMARY
(Continued)
< PREVIOUS VIEW
12-27
Monitoring, Inspections and Facility Maintenance
MAY 2009
2007). The 7-day inspection requirement normally applies to impoundments that are classifed as
having high or signifcant hazard potential; sites with low-hazard-potential impoundments may
be inspected on a less frequent schedule, if approved by the MSHA District Manager. Personnel
who perform these inspections are required to: (1) have current MSHA training, (2) complete inspec-
tion forms for documentation, and (3) maintain a record of the inspections. It is recommended that
7-day inspections be supplemented by periodic inspections by engineering professionals (engineers
or experienced technicians familiar with the design and construction requirements of the disposal
facility and the purpose and function of the instrumentation). The periodic inspections should be
documented in reports that are retained for reference. These engineering inspections and associated
reports may also be used for periodic certifcation of stage construction by a registered Professional
Engineer, as required by federal and state regulatory agencies. Monitoring data, such as seepage
rates and piezometer levels, which are collected during 7-day inspections, need to be routinely
reviewed by engineering personnel familiar with the design of the facility to verify that the facility
is performing as designed. If monitoring data suggest that the facility is not performing adequately,
the observed conditions should be investigated, and the Engineer should be advised and consulted.
12.3.1.2 Non-Impounding Facilities
Non-impounding coal refuse disposal facilities should be inspected periodically in compliance with
the Operation and Maintenance Plan and as required by state permits (e.g., monthly and quarterly).
In addition to specifc permit requirements, these inspections should cover diversion and perimeter
ditches, haul roads and guters, benches and guters, disposal work surfaces and embankment slopes,
refuse compaction, reclaimed surfaces and slopes, instrumentation, etc.
12.3.2 Event-Related Inspections
Inspections of impounding facilities should be performed afer events that may have negatively
impacted the integrity of the impounding embankment. Examples of such events include mining
subsidence, severe storm events, prolonged periods of cold weather, nearby blasting, earthquakes,
sudden increase in pore pressures in the dam, sudden increase or decrease of internal drain fows,
and landslides or major sloughing near or on the embankment. The scope of such inspections is
dependent upon the nature of the event.
12.3.3 Annual Dam Safety Inspections
It is recommended that a dam safety inspection be performed at least annually by an engineer
familiar with the design and construction requirements of the facility (preferably the design engi-
neer). Annual dam safety inspections and a formal report by a registered Professional Engineer are
required by some state agencies. Sometimes these inspections can be combined with other site data
to satisfy the MSHA annual report and certifcation requirement. The annual inspection should
include a review of critical aspects of the facility (such as the results of a potential failure modes
analysis, as discussed in Section 12.1.1, or other issues documented in the Operation and Mainte-
nance Plan) and thorough inspection of the facility including embankments and associated abut-
ments. Ploted instrumentation records and construction data should be reviewed. Preparation of
completed embankment cross sections refecting instrumentation data may be an efective way to
evaluate large or complicated geometries and the signifcance of piezometric levels. A tabulation of
other monitoring data collected during the year should be prepared and submited in the inspec-
tion report. Photographs should be taken at specifc locations during the inspection and compared
to earlier photographs taken from the same locations for evaluation of changes over time. Photo-
graphs should also be taken of areas that require repair or other special atention. The following
data should be compiled in graphs or drawings:
Piezometric levels with corresponding pool level elevations
Internal drain fow rates with corresponding pool level elevations
< PREVIOUS VIEW
12-28
Chapter 12
MAY 2009
Seepage fow rates with corresponding pool level elevations
Refuse embankment density test results
As-built conditions based upon most recent survey data
As-built data and drawings for critical items such internal drains, piezometer, decant
pipes and spillways
Site rainfall and temperature records, if appropriate
A review of upstream and downstream areas should be conducted, and changed conditions that
could impact the hydrology and hydraulics of the embankment should be identifed. The review of
the downstream area should concentrate on identifcation of new development and/or flls that could
impact downstream food levels in the event of embankment failure (Chapter 14) and the appropriate
hazard-potential classifcation for the site. For instance, a bridge or fll constructed in the downstream
foodplain could require a modifcation to the dam breach analysis to account for the potential chan-
nel constrictions.
Upstream areas should be inspected to determine if watershed conditions remain consistent with
the design assumptions. If the watershed area has efectively increased or disturbance has increased
runof potential, the storm routing analysis may have to be modifed and the embankment confgura-
tion possibly adjusted to provide adequate freeboard.
An evaluation based upon the visual observations and review of operating records should be per-
formed with respect to evaluation of the overall condition of the embankment. Should areas require
remedial atention, they should be identifed along with recommended repairs. Similarly, the Opera-
tion and Maintenance Plan should be reviewed and modifed as needed.
The annual inspection report is a good place to provide information related to the anticipated life of
a refuse disposal facility and to provide an evaluation of whether changes may be needed to meet
freeboard requirements. Thus, an analysis of observed coal refuse generation rates and the remaining
capacity of the facility should be performed and included in the annual inspection report.
12.4 CERTIFICATION AND CONFORMANCE REPORTING
An annual report that includes a description of changes in the confguration of the impoundment is
required under 30 CFR 77.216-4. As part of the annual report, MSHA requires a certifcation by a
registered Professional Engineer that all construction, operation and maintenance of an impound-
ment was performed in accordance with the approved plan for the facility. Some state agencies also
require certifcation of construction of refuse disposal facilities for specifc periods, structures or
stages. The registered Professional Engineer should be familiar with design, construction, operation
and maintenance of the structure, including dam safety criteria. The design engineer, if involved in
the monitoring and inspection of the refuse disposal facility, may provide support or fulfll the certi-
fcation responsibility.
Certifcation is generally performed concurrently with reporting of construction monitoring or inspec-
tion and is an expression of the Engineers professional judgment with respect to the consistency of
the completed work relative to the approved design plan. It is recommended that the certifcation
be supported by the results of the monitoring and inspection activities performed in response to
the designers specifcations and the Professional Engineers ongoing guidance, which may include
visual observation, testing, and instrumentation measurements. The certifcation is not a guarantee
or warranty of performance of the facility and does not relieve any responsible partys obligation to
monitor, inspect, construct or operate the facility in accordance with the design plans or to respond
to conditions or situations that may represent a hazard to site personnel, public safety, or the envi-
< PREVIOUS VIEW
12-29
Monitoring, Inspections and Facility Maintenance
MAY 2009
ronment. It is recommended that the annual report cite the specifc scope of activities to which the
certifcation applies and reference (or provide documentation of) the observations, testing, and instru-
mentation measurements providing support for professional judgments.
When observations, test results or instrumentation readings identify possible deviations from the
approved plan, the Engineer should evaluate the situation to determine if there are non-conformances
and should implement the actions necessary to address such non-conformances, which should be
documented in a report available to the regulatory agencies (e.g., operators 7-day inspection doc-
umentation, other regular inspection report such as monthly or quarterly documentation, or the
annual impoundment inspection report). MSHA should be notifed immediately in accordance with
regulatory requirements (30 CFR 77.216-3) when a potentially hazardous condition to site person-
nel or the public develops, or in particular circumstances where the approved disposal plan requires
that MSHA or state agencies be notifed immediately. The Operation and Maintenance Plan should
provide guidance regarding what actions need to be taken when observations, testing or instrumen-
tation measurements identify suspected adverse, potentially hazardous, or unsafe or unacceptable
operating conditions. When changes need to be made in the design plans or construction specifca-
tions, the modifcations should be submited to regulatory agencies for approval.
12.5 MAINTENANCE
Maintenance of a refuse disposal facility throughout its entire operating life is important for many
reasons, including:
Exposure to weather for long time periods results in the natural deterioration of
some components that require periodic repair and maintenance.
Neglecting to perform minor repair work can result in future major repairs or undis-
covered defciencies leading to major distress.
A primary aim of design is to eliminate maintenance afer abandonment. Regular rou-
tine maintenance during the operating life will facilitate abandonment of the facility.
The broad range of conditions that can be found at refuse disposal facilities makes it impractical to
present a detailed set of instructions for maintenance and repair of each condition that might arise.
Maintenance issues should be addressed in the facility Operation and Maintenance Plan. The check-
list presented in Table 12.1 and the related discussion in Section 12.1 can serve as a guide for devel-
oping a site-specifc Operation and Maintenance Plan. Some general maintenance requirements are
discussed briefy in the following paragraphs.
12.5.1 Routine Maintenance
Routine maintenance should be performed on a continuous basis in accordance with the Opera-
tion and Maintenance Plan. This typically includes reseeding of disturbed vegetated areas, removal
of debris and patching and repair of collection and diversion channels, removal of debris from the
critical areas of the facility, and repair and reseeding of eroded areas. Such maintenance should be
documented.
Routine maintenance should also include control of vegetation during facility operation. Periodic
cuting of vegetation on slopes and toe areas of impoundments is necessary to facilitate dam safety
inspections and will aid in preventing the development of heavy brush and trees. Trees should not
be allowed to develop on an impounding embankment. Should trees develop and mature, their
roots can provide seepage pathways leading to voids upon decay, and in some instances extensive
slope repair measures may eventually be required. Guidance related to this issue is presented in
Marks and Tschantz (2002).
< PREVIOUS VIEW
12-30
Chapter 12
MAY 2009
12.5.2 Maintenance after Unusual Events
Some portions of a coal refuse disposal facility and its appurtenant structures may be disturbed by
unusual events such as heavy rainfalls, extremely severe frost periods, or periods of very high winds.
If heavy rain is forecast, decants and spillways should be checked to ensure that they are clear of
debris. Normally, the bulk of maintenance work afer an unusual event is associated with repair of
diversion and collection channels, eroded zones, and vegetation. The degree of disturbance can be
minimized by a conscientious routine maintenance program that eliminates minor defciencies that
could become weak points during severe weather.
12.5.3 Long-Term Maintenance
The natural deterioration of materials exposed to weather requires that repair work beyond routine
maintenance be undertaken from time to time. Such work items can include major repairs to riprap
and replacement of concrete linings, valves, trash racks, monitoring instruments, etc. The frequency
of this work will depend upon the care taken during the site routine maintenance program and the
economics associated with performing numerous smaller maintenance eforts versus less frequent
major maintenance. The Operation and Maintenance Plan should provide schedules for all levels of
maintenance.
< PREVIOUS VIEW
Chapter 13
INSTRUMENTATION
AND PERFORMANCE
MONITORING
13-1 MAY 2009
The design of coal refuse disposal facilities is governed by criteria for achieving desired levels of
structural, geotechnical and hydraulic performance from initial startup through facility construction
and operation to abandonment. To verify that the desired performance levels are being achieved,
instrumentation should be installed and regularly monitored. Timely collection and review of instru-
mentation data can allow performance problems to be detected and addressed before unsafe condi-
tions develop. This chapter discusses the factors that should be considered in planning a site-specifc
instrumentation program and the types of devices used for monitoring. Supporting discussion
regarding the uncertainty associated with instrument measurements and types of instrument trans-
ducers and data acquisition systems is presented in Appendix 13A. Data typically monitored and the
types of instrumentation most commonly used to monitor embankment performance at coal refuse
facilities include:
Vertical and lateral displacements at the ground surface using surface monuments
Impoundment pool levels using staf gauges
Piezometric levels and pore-water pressures in embankments and foundations using
standpipe and vibrating-wire piezometers
Surface water fow primarily from seepage and mine discharges using weirs
Subsurface vertical and lateral displacements in situations of adverse conditions
(e.g., setlement and slope deformation) using extensometers and inclinometers
Meteorological conditions at or near the facility using weather stations (primarily
rainfall gauges)
These and other instrumentation are discussed herein along with the ability to remotely monitor and
efciently process data from instruments.
13.1 INSTRUMENTATION PROGRAM PLANNING
Similar to instrumentation programs for other civil works projects, instrumentation program plan-
ning for a coal refuse disposal facility should follow a systematic approach with defned objectives.
The process should follow a logical series of steps leading to preparation of construction plans and
specifcations that prescribe: (1) instrument types and installation methods, (2) performance and
maintenance requirements, (3) data acquisition methods, (4) sampling intervals and reporting, and
< PREVIOUS VIEW
13-2
Chapter 13
MAY 2009
(5) expected measurement ranges including appropriate action and hazard warning levels. This sys-
tematic approach, as defned by Dunniclif (1997), should follow the steps presented in Table 13.1.
The full beneft of an instrumentation program can best be realized if these steps are considered and
carefully implemented. In doing so, the following adage by Ralph Peck (1972) can be realized,
We need to carry out a vast amount of observational work, but what we do
should be done for a purpose and done well.
Monitoring the performance of coal refuse disposal facilities typically involves the measurement of
deformations, piezometric levels, impoundment levels, seepage fows, and other site-specifc param-
eters. The steps presented in Table 13.1 provide a rationale for determining the instrumentation and
stafng requirements, establishing provisions for instrument maintenance, and identifying the use
and beneft of monitoring results. Table 13.1 can be used as a design aid to focus atention on the
important parameters and locations to be monitored and to avoid the use of instrumentation that
may yield litle value. Among the critical steps is No. 2 defning performance questions that need to
be answered. Performance questions arise from project or geotechnical conditions that could lead to
instability or non-functionality of portions of the structure. Ofen, critical performance features such
as an internal drain or suspect conditions such as a weak subsurface foundation layer beneath the
facility or an observed seepage condition, may play a role. These critical or suspect conditions will
then be the focus of performance questions related to instrumentation.
Some performance questions that are typically relevant for coal refuse disposal impoundments are
summarized in Table 13.2. While this list includes common questions, each site will undoubtedly have
unique features or conditions that may introduce other performance questions. At new facilities, the
designer must rely on project information and engineering analyses and judgment to identify perfor-
mance questions and select instrumentation. At existing facilities, where some performance data are
available, the designer may identify suspect conditions such as observed seepage or deformation that
may lead to more refned answers to the performance questions.
While impoundments are generally facilities where the consequences of non-performance are great,
instrumentation should also be employed at non-impounding embankments and slurry cell facilities.
Similar performance questions and suspect conditions should be evaluated as part of the identifca-
tion of the need for and the types of instrumentation to be employed.
In selecting an instrument to measure a desired parameter within the project seting, consideration
should be given to: (1) instrument error and uncertainty, (2) instrument type reliability and sur-
vivability relative to its expected application, and (3) possible integration of the instrument into a
data acquisition system to facilitate data collection, processing, management and decision making.
Discussion of the uncertainty associated with instrument measurements and the types of instrument
transducers and data acquisition systems is presented in Appendix 13A.
13.2 MEASUREMENT TECHNIQUES
Instrumentation is typically installed at coal refuse disposal facilities to monitor movements and dis-
placements, piezometric levels and pore-water pressures, and surface- and seepage-water fow and
the hydrologic and operational factors that infuence these fows. Instruments may also be installed
to monitor miscellaneous factors such as soil pressure, vibration and shock, and internal tempera-
tures. The following subsections discuss the signifcance of these measurements and instrumentation
that can be used to provide the required data.
Selection of appropriate instrumentation depends on a variety of factors discussed in the previous
sections of this chapter. In addition, instrument selection relates to whether performance monitoring
< PREVIOUS VIEW
13-3
Instrumentation and Performance Monitoring
MAY 2009
TABLE 13.1 INSTRUMENTATION PLANNING AND DESIGN STEPS
Step 1
Defne project conditions and mechanisms that control expected behavior including subsur-
face conditions and stratigraphy, engineering properties of subsurface soil, rock and groundwater,
environmental conditions and other factors that may affect the planned construction.
Step 2
Defne the performance questions that need to be answered. For a coal refuse facility these
might include: (1) What are the initial or current site conditions? (2) Is performance satisfactory
during construction and facility operations? and (3) Is performance satisfactory during special load-
ing conditions such as rapid drawdown or upstream construction?
Step 3
Select the most important parameters to be monitored, magnitudes of change, and each
type of instrumentation considering parameter variations resulting from both cause and effect.
For example, lateral slope displacements may result from elevated groundwater levels, and specifc
hazard warning levels can be determined. If a clear purpose cannot be defned, the need for instru-
mentation cannot be defended.
Step 4
Identify staffng responsibilities for monitoring, interpreting and devising remedial action(s)
in terms of required labor and materials needed to respond to problem situations, including
resources and reporting requirements.
Step 5
Select instruments and locations considering reliability for measurement of the desired parameter
within the project site setting and the appropriateness of the location for measuring predicted behav-
ior, compatibility with methods of analysis to be used, and device survivability. Other features that
should be considered include in-service calibration, operator skill requirements, potential interfer-
ence during construction, and location access during installation and reading.
Step 6
Develop record of factors that may affect the measured data and establish procedures for
controlling data quality such as geologic conditions in the vicinity of the instrument, use of redun-
dant instruments, regular examination of data for consistency, and in-service calibration checks.
Step 7
Prepare instrumentation system report, budget and procurement specifcations summarizing
Steps 1 to 6 and including sections on the contracting method and basis for instrument procurement
and feld instrumentation services.
Step 8
Integrate instrumentation system report into the Operation and Maintenance Plan, including
data collection, processing, presentation, interpretation, reporting, calibration and maintenance.
(ADAPTED FROM DUNNICLIFF, 1993)
is required for a relatively short period of time up to a few years, or for a much longer time period up
to and including the expected service life of the refuse facility.
13.2.1 Movements and Displacements
The performance of embankments can generally be evaluated by monitoring the movement of the
embankment surface or confguration changes within the embankment and its foundation. For static
loading conditions (i.e., no earthquake or blasting loads), movement can be determined by: (1) visual
observation, (2) measurement of surface movements, and (3) measurement of internal movements.
Table 13.3 provides an evaluation of each approach relative to complexity and cost, applicability, and
limitations. Additional information is provided by Dunniclif (1981), Hanna (1985), Bartholomew et
al. (1987), Dunniclif (1993) and USACE (1995c).
13.2.1.1 General Observations
Periodic surveying of monuments and instruments provides an efective means for observing the
magnitude and rate of change of deformations at a coal refuse disposal facility during:
Initial facility development and initial disposal operations
Normal disposal operations
< PREVIOUS VIEW
13-4
Chapter 13
MAY 2009
TABLE 13.2 PERFORMANCE QUESTIONS FOR SLURRY IMPOUNDMENTS
Questions
Typical Instruments
for Monitoring
Other Typical
Measures for
Monitoring
Suspect
Conditions
Embankment Conditions
1. Where is the phreatic surface:
a. in the refuse embankment?
b. in the fne refuse deposit (e.g., when upstream construction is
employed)?
Piezometers
Closed-System
Piezometers
2. What is the seepage rate?
Weirs
3. What is the discharge rate from an internal drain?
Weirs
4. Are embankment grades and alignments conforming to plans?
Survey Monument Survey Methods
5. Are unacceptable deformations occurring:
a. at the embankment surface?
b. within the embankment?
Survey Monument
Survey Methods
Inclinometer
6. Could unacceptable pore-water pressure develop due to the
effects of the rate of fll placement, rapid drawdown, or earthquake
loading?
Closed-System
Piezometers
7. Could unacceptable deformations develop due to the effects of
the rate of fll placement, mine subsidence, rapid drawdown, or
earthquake loading?
Survey Monuments
Extensometers
Survey Methods
Inclinometer
Foundation Conditions
1. Could unacceptable deformations develop due to the presence of
compressible foundation materials, subsidence from mining below
the embankment or impoundment, or earthquake loading?
Survey Monuments
Extensometers
Geophysical
Surveys
2. Could unacceptable pore-water pressures develop due to the
effects of rapid embankment construction, upstream construction
over soft fne refuse, or earthquake loading?
Closed-System
Piezometers
3. Could unacceptable pore-water pressures develop due to the
effects of an elevated reservoir level or an undetected pervious
stratum?
Closed-System
Piezometers
Abutment Conditions
1. Will excavation or natural slopes be stable:
a. during construction?
b. long term?
Survey Monuments Survey Methods
< PREVIOUS VIEW
13-5
Instrumentation and Performance Monitoring
MAY 2009
Questions
Typical Instruments
for Monitoring
Other Typical
Measures for
Monitoring
Suspect
Conditions
2. Could unacceptable deformations or cracking develop due to
subsidence from mining beneath the abutment or impoundment?
Survey Monuments
Extensometers
3. Could unacceptable pore-water pressures develop due to the
effects of an elevated reservoir level or an undetected pervious
stratum?
Open-Standpipe
Piezometers
Reservoir Level
1. Is the water level in the impoundment at an acceptable level, and
is there adequate freeboard?
Staff Gage
Breakthrough Potential
1. Could an unusual increase in seepage quantity or change in
seepage quality from underground mine workings indicate a
potential problem related to impoundment leakage or possible
breakthrough?
Weir
2. Could the level of water in underground mine workings have an
effect on the impoundment or indicate possible breakthrough?
Observation Well
Decant Works, Primary or Emergency Spillways
1. Is the fow rate from the decant acceptable?
Weir
Pipe Discharge
Measurement
2. Is the strain within the fexible decant conduit acceptable
considering embankment loading and backfll support?
Dial Gage
Defectometer
Circumferential Survey
Camera
3. Is there movement affecting the primary conduit spillway joints (or
cracks) that could lead to separation?
Dial Gage
(Crack Monitor)
4. Is there movement affecting the lined emergency spillway joints (or
cracks) that could lead to separation?
Dial Gage
(Crack Monitor)
5. Are the excavation slopes for the emergency spillway stable? Survey Methods
Survey Monuments
Weather Conditions
1. How do changes in measured facility performance (e.g., phreatic
surface, fow from internal drains) relate to changes in weather
conditions such as precipitation?
Precipitation Gage Weather Station
Note: Instrumentation shown in italics is generally considered when adverse conditions are observed or suspected
(e.g., seepage or movement).
TABLE 13.2 PERFORMANCE QUESTIONS FOR SLURRY IMPOUNDMENTS
(Continued)
< PREVIOUS VIEW
13-6
Chapter 13
MAY 2009
Changes in the rate of disposal, in the physical characteristics of the fne and coarse
coal refuse components, or in the phreatic or piezometric levels in the embankment
Facility abandonment
Excessive movement or progressively increasing rates of movement can provide a warning of a
potential failure and may indicate that modifcation of a facility operation is necessary.
Long-term periodic monitoring of surface movements provides an inexpensive record of embank-
ment behavior. If changes due to natural phenomena or operations result in a reduction in stability
not anticipated in the design, long-term records can provide a forewarning of impending distress.
The potential value of surveyed records in relation to their cost justifes the installation of bench-
marks and monuments for most types of refuse facilities, even if the instrumentation must be regu-
larly monitored during the facility life and to abandonment.
TABLE 13.3 SUMMARY OF GENERAL METHODS FOR DETECTING EMBANKMENT
MOVEMENTS AND DISPLACEMENTS
Technique
Complexity
and Cost
Applicability Limitations
Visual Observations
Low to
Moderate
Normally used for monitoring
general conditions to identify areas
of potential distress where more
detailed evaluation is required.
Surface cracking indicates the
occurrence of displacements, but
does not provide quantitative data.
Measurements of
Surface Movements
Low to High
Reasonable costs and high accu-
racy provide desirable monitoring
procedure for practically any condi-
tion. Should be routine practice for
all facilities.
Surface movements can be very
local without being signifcant to
overall safety of facility.
Measurement of
Internal Movements
Moderate to
High
Necessary when cause of
movements is important and
surface measurements are not
suffcient for interpretation.
Programs requiring complex
instrument installation, monitoring
and interpretation must be
performed by an expert.
If a surface movement monitoring program is implemented primarily to develop long-term records
of embankment behavior, the measurement points are usually located along representative cross
sections perpendicular to the longitudinal axis of the embankment (e.g., at the toe, on benches, and
at the top of slope along each section) and at readily accessible points on other facility structures. A
limitation of conventional surveying techniques is that only movements at the surface are monitored.
Observations at the surface are not always sufcient for determining if the movements are shallow
and relatively unimportant or if they are associated with deeper, more signifcant conditions occur-
ring within the embankment or its foundation. In the later case, it may be necessary to install more
sophisticated subsurface instrumentation such as inclinometers to beter defne the mechanism caus-
ing the movement and its location.
Optical level surveys (land surveys) are ofen the best method for monitoring vertical movements
due to:
Setlement resulting from consolidation of an embankment foundation or constituent
materials
Impending stability failure due to embankment movement or foundation deforma-
tions
Surface subsidence from underground mining
< PREVIOUS VIEW
13-7
Instrumentation and Performance Monitoring
MAY 2009
The order of accuracy required for land surveys depends on the total potential magnitude of move-
ment, while the accuracy obtained is a function of the equipment and personnel employed and the
reliability of benchmark datums used for control.
Monitoring of horizontal surface movements is useful for detecting conditions that may indicate
impending instability or for verifying that such movements are not occurring. Preferably, identi-
cal reference points or monuments should be used for measuring both vertical and horizontal
movements.
Instruments for measuring vertical displacements within an embankment or its foundation are useful when:
Accurate setlement data are required for comparison to predicted embankment or
foundation setlement.
The integrity of an impervious core or internal drainage system is dependent on
limiting the amount of setlement.
Large setlements could afect drainage slopes and ditches afer abandonment.
A contractor is being paid for construction on a unit-price basis and the volume of
material placed could signifcantly change based on the magnitude of setlement.
A rigid structure is to be placed on an embankment at the completion of construction.
The embankment was constructed over setled fne refuse or the embankment foun-
dation is sof clay and potentially large setlements are a possibility.
The installation of internal setlement instrumentation is not generally recommended for routine
refuse disposal embankment construction. However, several of the techniques discussed in the fol-
lowing section are relatively simple and inexpensive, and they may be considered when embankment
setlement is important. The installation, monitoring and interpretation of data from these systems
should only be undertaken under the supervision of a knowledgeable engineer.
Durability of the installed instrumentation is particularly important for systems placed in an embank-
ment. Care in placement is critical, because instruments can ofen be irreparably damaged due to
improper installation techniques. The high potential for corrosion at coal refuse facilities should
always be considered, and plastic pipes are preferred over metal pipes unless it is known that high
temperatures that might afect the long-term behavior of the plastic pipe could occur.
For all types of instrumentation, erroneous data are much more likely when vertical setlement is
coupled with large horizontal movement. In such cases, instrumentation for measurement of both
vertical and horizontal movements should be installed. The purpose of measurement of horizontal
displacements within an embankment is normally to determine if there is instability and the depth
at which movement is occurring or to verify that surface distortions due to creep movements do not
have a deeper origin. Inclinometers are ofen used to determine the depth to the surface where move-
ment is occurring prior to implementation of remedial measures so that the costly remedial efort is
focused solely on the problem area.
Normally, sophisticated instrumentation is not required for newly-constructed refuse embankments
designed in accordance with current engineering practice and constructed in conformance with
detailed plans and specifcations. Possible exceptions may include embankments:
Located above populated areas.
Constructed where site conditions do not permit access for the desired scope of
exploration and testing, such as where steep slopes and/or dense ground cover limit
access to drilling and sampling equipment.
< PREVIOUS VIEW
13-8
Chapter 13
MAY 2009
Supported on a material that cannot be readily tested, such as setled fne refuse or
sof, sensitive clay.
Constructed over or adjacent to areas of past or active underground mining.
Where dynamic loading is a critical consideration during design.
More sophisticated instrumentation is ofen used to investigate existing facilities where limited data
relative to site conditions and construction practices (e.g., fll material characteristics) are available.
The value of installing and monitoring the instrumentation must be carefully judged against the abil-
ity to interpret the data (i.e., the ability to recognize adverse conditions and distress). When internal
movement instrumentation is required, the location and arrangement of instruments should be based
on the judgment of an expert, and the installation and monitoring should be conducted under the
direct supervision of an engineer or technician that is experienced with the equipment used.
Movements associated with dynamic loads (e.g., earthquake or nearby blasting) difer from those
associated with static loads in that measurements must be taken during the occurrence of the
dynamic condition. Generally, where earthquake considerations are important in the site selection
process, potential dynamic displacements are estimated based upon seismic engineering analyses
and no on-site measurements are involved. However, in the case of blasting, the amount of move-
ment realized is dependent upon geologic conditions, the explosive material and blasting technique,
and the location of the blast relative to the embankment. Ofen, blasting efects are minor and will not
afect embankment stability. Exceptions are when blasting is very close to an embankment or other
facility structure or rock abutment or the magnitude of the blast is unusually high. In these isolated
cases, surface monitoring of resultant movements may be justifed for verifcation that no signifcant
damage to the refuse disposal facility has occurred. Instrumentation for this purpose is discussed in
Section 13.2.4.2.
13.2.1.2 Movement Measurement Techniques
As listed in Table 13.4, Dunniclif (1993) identifes the following as general categories of deformation
measuring techniques:
Surveying methods
Surface extensometers
Tiltmeters
Probe extensometers
Fixed embankment extensometers
Fixed borehole extensometers
Inclinometers
Descriptions of instrumentation associated with each of these deformation measurement techniques
are presented in the following sections. Supplemental techniques for deformation measurement,
including transverse-deformation gages, liquid-level gages, time-domain refectometry, and fber-
optic gages are presented in Appendix 13A at the end of this chapter.
13.2.1.2.1 Surveying Methods
Surveying methods are generally used for monitoring the magnitude and rate of vertical and hori-
zontal movement of the ground surface, structures, and accessible parts of subsurface instruments
at construction sites (Dunniclif, 1993). For many applications, these methods are adequate for per-
formance monitoring, and instrumentation is used only if greater accuracy is needed or if subsurface
movements need to be determined. When instrumentation is employed, surveying methods are ofen
< PREVIOUS VIEW
13-9
Instrumentation and Performance Monitoring
MAY 2009
T
A
B
L
E
1
3
.
4
I
N
S
T
R
U
M
E
N
T
C
A
T
E
G
O
R
I
E
S
F
O
R
M
E
A
S
U
R
I
N
G
M
O
V
E
M
E
N
T
S
C
a
t
e
g
o
r
y
H
V
A
R
S
U
C
o
n
d
i
t
i
o
n
s
f
o
r
U
s
e
R
e
l
a
t
i
v
e
C
o
m
p
l
e
x
i
t
y
R
e
l
a
t
i
v
e
C
o
s
t
A
p
p
l
i
c
a
b
i
l
i
t
y
t
o
C
o
a
l
R
e
f
u
s
e
D
i
s
p
o
s
a
l
F
a
c
i
l
i
t
i
e
s
S
u
r
v
e
y
i
n
g
M
e
t
h
o
d
s
S
u
r
f
a
c
e
m
o
v
e
m
e
n
t
s
S
i
m
p
l
e
L
o
w
t
o
M
o
d
e
r
a
t
e
R
o
u
t
i
n
e
f
o
r
a
l
l
e
m
b
a
n
k
m
e
n
t
s
.
S
u
r
f
a
c
e
E
x
t
e
n
s
o
m
e
t
e
r
s
C
r
a
c
k
g
a
g
e
s
C
o
n
v
e
r
g
e
n
c
e
g
a
g
e
s
D
e
f
o
r
m
a
t
i
o
n
b
e
t
w
e
e
n
f
x
e
d
p
o
i
n
t
s
S
i
m
p
l
e
L
o
w
U
s
u
a
l
l
y
l
i
m
i
t
e
d
t
o
s
t
r
u
c
t
u
r
e
a
n
d
r
o
c
k
m
o
v
e
m
e
n
t
s
,
i
n
c
l
u
d
i
n
g
s
u
b
s
i
d
e
n
c
e
m
o
n
i
t
o
r
i
n
g
.
T
i
l
t
m
e
t
e
r
s
R
o
t
a
t
i
o
n
s
o
n
a
s
u
r
f
a
c
e
S
i
m
p
l
e
L
o
w
t
o
M
o
d
e
r
a
t
e
U
s
u
a
l
l
y
l
i
m
i
t
e
d
t
o
s
t
r
u
c
t
u
r
e
m
o
v
e
m
e
n
t
s
.
P
r
o
b
e
E
x
t
e
n
s
o
m
e
t
e
r
s
M
e
c
h
a
n
i
c
a
l
p
r
o
b
e
g
a
g
e
s
E
l
e
c
t
r
i
c
a
l
p
r
o
b
e
g
a
g
e
s
P
r
o
b
e
e
x
t
e
n
s
o
m
e
t
e
r
s
w
i
t
h
i
n
c
l
i
n
o
m
e
t
e
r
s
D
i
s
p
l
a
c
e
m
e
n
t
(
u
s
u
a
l
l
y
v
e
r
t
-
i
c
a
l
)
b
e
t
w
e
e
n
t
w
o
o
r
m
o
r
e
p
o
i
n
t
s
a
l
o
n
g
a
s
i
n
g
l
e
a
x
i
s
M
o
d
e
r
a
t
e
t
o
c
o
m
p
l
e
x
M
o
d
e
r
a
t
e
U
s
e
f
u
l
i
f
s
e
t
t
l
e
m
e
n
t
o
f
s
o
f
t
f
o
u
n
d
a
t
i
o
n
i
s
c
r
i
t
i
c
a
l
;
n
o
r
m
a
l
l
y
s
i
n
g
l
e
s
e
t
t
l
e
m
e
n
t
p
l
a
t
e
s
a
r
e
a
d
e
q
u
a
t
e
.
F
i
x
e
d
E
m
b
a
n
k
m
e
n
t
E
x
t
e
n
s
o
m
e
t
e
r
s
S
e
t
t
l
e
m
e
n
t
p
l
a
t
f
o
r
m
S
e
t
t
l
e
m
e
n
t
a
t
s
i
n
g
l
e
d
e
p
t
h
.
P
r
a
c
t
i
c
a
l
f
o
r
p
r
e
-
o
r
p
o
s
t
-
c
o
n
s
t
r
u
c
t
i
o
n
i
n
s
t
a
l
l
a
t
i
o
n
S
i
m
p
l
e
L
o
w
U
s
e
f
u
l
i
f
s
e
t
t
l
e
m
e
n
t
o
f
s
o
f
t
f
o
u
n
d
a
t
i
o
n
i
s
c
r
i
t
i
c
a
l
.
F
i
x
e
d
B
o
r
e
h
o
l
e
E
x
t
e
n
s
o
m
e
t
e
r
s
S
i
n
g
l
e
-
a
n
d
m
u
l
t
i
-
p
o
i
n
t
S
u
b
s
u
r
f
a
c
e
s
e
t
t
l
e
m
e
n
t
p
o
i
n
t
s
a
n
d
r
o
d
g
a
g
e
s
D
i
s
p
l
a
c
e
m
e
n
t
b
e
t
w
e
e
n
t
w
o
o
r
m
o
r
e
p
o
i
n
t
s
a
l
o
n
g
a
s
i
n
g
l
e
a
x
i
s
M
o
d
e
r
a
t
e
t
o
c
o
m
p
l
e
x
M
o
d
e
r
a
t
e
U
s
e
f
u
l
i
f
s
e
t
t
l
e
m
e
n
t
o
f
s
o
f
t
f
o
u
n
d
a
t
i
o
n
o
r
r
o
c
k
s
l
o
p
e
s
i
s
c
r
i
t
i
c
a
l
.
I
n
c
l
i
n
o
m
e
t
e
r
s
M
o
v
e
m
e
n
t
p
r
o
f
l
i
n
g
i
n
v
e
r
t
i
c
a
l
,
h
o
r
i
z
o
n
t
a
l
o
r
i
n
c
l
i
n
e
d
c
a
s
i
n
g
s
M
o
d
e
r
a
t
e
M
o
d
e
r
a
t
e
M
o
s
t
c
o
m
m
o
n
l
y
u
s
e
d
m
e
t
h
o
d
,
b
u
t
l
i
m
i
t
e
d
t
o
z
o
n
e
s
w
h
e
r
e
d
e
f
o
r
m
a
t
i
o
n
p
r
o
f
l
e
s
a
r
e
n
e
e
d
e
d
o
r
a
r
e
a
c
o
n
c
e
r
n
.
T
r
a
n
s
v
e
r
s
e
-
D
e
f
o
r
m
a
t
i
o
n
G
a
g
e
s
I
n
-
p
l
a
c
e
i
n
c
l
i
n
o
m
e
t
e
r
s
M
o
v
e
m
e
n
t
p
r
o
f
l
i
n
g
M
o
d
e
r
a
t
e
t
o
c
o
m
p
l
e
x
M
o
d
e
r
a
t
e
t
o
H
i
g
h
U
s
e
f
u
l
o
n
l
y
i
f
a
u
t
o
m
a
t
i
c
m
o
n
i
t
o
r
i
n
g
i
s
p
l
a
n
n
e
d
.
L
i
q
u
i
d
-
L
e
v
e
l
G
a
g
e
s
S
i
n
g
l
e
-
a
n
d
m
u
l
t
i
-
p
o
i
n
t
g
a
g
e
s
F
u
l
l
-
p
r
o
f
l
e
g
a
g
e
s
S
e
t
t
l
e
m
e
n
t
o
f
f
o
u
n
d
a
t
i
o
n
o
r
e
m
b
a
n
k
m
e
n
t
f
l
l
M
o
d
e
r
a
t
e
t
o
c
o
m
p
l
e
x
M
o
d
e
r
a
t
e
U
s
e
f
u
l
i
f
s
e
t
t
l
e
m
e
n
t
o
f
s
o
f
t
f
o
u
n
d
a
t
i
o
n
i
s
c
r
i
t
i
c
a
l
.
T
i
m
e
-
D
o
m
a
i
n
R
e
f
e
c
t
o
m
e
t
r
y
M
o
v
e
m
e
n
t
d
e
t
e
c
t
i
o
n
a
t
l
o
c
a
t
i
o
n
s
a
l
o
n
g
c
a
b
l
e
a
x
i
s
M
o
d
e
r
a
t
e
L
o
w
t
o
M
o
d
e
r
a
t
e
S
t
i
l
l
e
v
o
l
v
i
n
g
m
e
t
h
o
d
,
b
u
t
s
h
o
w
s
e
x
c
e
l
l
e
n
t
p
r
o
m
i
s
e
.
F
i
b
e
r
-
O
p
t
i
c
S
e
n
s
o
r
s
L
o
c
a
l
d
e
f
o
r
m
a
t
i
o
n
a
n
d
s
t
r
a
i
n
m
e
a
s
u
r
e
m
e
n
t
s
M
o
d
e
r
a
t
e
M
o
d
e
r
a
t
e
t
o
H
i
g
h
S
t
i
l
l
e
v
o
l
v
i
n
g
m
e
t
h
o
d
,
b
u
t
s
h
o
w
s
e
x
-
c
e
l
l
e
n
t
p
r
o
m
i
s
e
,
e
s
p
e
c
i
a
l
l
y
f
o
r
l
o
n
g
-
t
e
r
m
m
o
n
i
t
o
r
i
n
g
a
p
p
l
i
c
a
t
i
o
n
s
.
L
e
g
e
n
d
:
H
=
H
o
r
i
z
o
n
t
a
l
d
e
f
o
r
m
a
t
i
o
n
A
=
A
x
i
a
l
d
e
f
o
r
m
a
t
i
o
n
S
=
S
u
r
f
a
c
e
d
e
f
o
r
m
a
t
i
o
n
V
=
V
e
r
t
i
c
a
l
d
e
f
o
r
m
a
t
i
o
n
R
=
R
o
t
a
t
i
o
n
a
l
d
e
f
o
r
m
a
t
i
o
n
U
=
S
u
b
s
u
r
f
a
c
e
d
e
f
o
r
m
a
t
i
o
n
(
A
D
A
P
T
E
D
F
R
O
M
D
U
N
N
I
C
L
I
F
F
,
1
9
9
3
)
< PREVIOUS VIEW
13-10
Chapter 13
MAY 2009
used for locating the instruments relative to a reference datum. Table 13.5 summarizes the advan-
tages, limitations, approximate accuracy, relative complexity, cost, and applicability to coal refuse
disposal facilities of the various surveying methods identifed by Dunniclif (1993).
Optical Leveling
Most setlement surveys at construction sites are conducted using engineers levels at second- or
third-order accuracy. Second-order leveling surveys are generally confned to extending vertical con-
trol data over long distances. Second-order leveling involves limited sight distances ( 225 feet), bal-
ancing foresight and hindsight, careful plumbing of the level rod, and readings made on well-defned
marks and stable turning points. Third order leveling surveys are used to establish vertical control
and to maintain benchmarks for project control, construction survey control, topographic survey
control and major structure points. Third-order leveling permits greater sight distances ( 300 feet),
along with balancing foresight and hindsight, plumbing of the level rod, and readings made on well-
defned marks and stable turning points. For the magnitude of elevation change typically of interest
at coal refuse disposal facilities, a third-order survey is usually adequate.
An important requirement for measuring horizontal movements is the establishment of reference
monuments away from the embankment being monitored that are known to be fxed against hori-
zontal movement and are convenient for the method of measurement used. This requirement may
be difcult in many coal mining areas where natural slopes are steep with resulting downhill creep
of the surface soils and where surface strains from nearby mine subsidence may be present. Where
possible, reference monuments should be placed on fat natural ground or on bedrock by excavating
holes in the soil/rock and backflling them with concrete. For monuments in soil, the depth should
extend below the maximum depth of frost penetration. Where assurance of undisturbed location
is not possible for the primary reference monuments used for routine monitoring, they should be
closed into a larger survey traverse or triangulation network for occasional checking of their loca-
tions. The large traverse should preferably have at least two permanent monuments in an area mini-
mally susceptible to movement, such as adjacent to a highway in a broad valley botom that does not
overlie coal seams. Additional discussion of monuments and benchmarks is provided at the end of
this section.
Trigonometric Leveling
Trigonometric leveling uses electronic distance measuring (EDM) equipment to measure the slope
distance from the survey instrument to a prism on the surveyed location and to calculate the eleva-
tion and horizontal position of the surveyed location. The angle between the horizontal and sur-
veyed location can be measured using a semi-precise (6-arc-second) or a precise (1- or 2-arc-second)
theodolite.
Distance Measurement by Taping
Distance measurement by taping is probably the easiest measurement technique for horizontal move-
ments because it requires only the use of a steel tape. When a known fxed point can be located in the
direction of anticipated movement, this procedure consists of measuring the distance between the
fxed reference and the monuments of interest. Typical corrections for sag, temperature and slope
should be applied in order to obtain accurate measurements. The limitations of tape measurement
should not preclude its use, because it may be the only means for locating lost reference points cov-
ered by deep snow or inadvertently buried by refuse placement.
Electronic Distance Measurement
Except for distances less than about 200 feet, distance measurement by taping has been replaced by
EDM equipment. EDMs are used to directly measure the distance change or lateral position change
< PREVIOUS VIEW
13-11
Instrumentation and Performance Monitoring
MAY 2009
by triangulation. EDMs make use of electromagnetic radiation (or lasers) to measure the distance
between the instrument and a refector prism.
Theodolite and Scale
Ofsets from baseline using a theodolite and scale are measurements at a right angle to a baseline.
These surveys are typically used for grade control for embankments and roadways. Laser beam level-
ing and ofsets provide a faster and more accurate alternative to optical leveling.
Total Station Survey
Total station instruments combine electronic distance measurement, digital theodolites and micropro-
cessors to simultaneously measure slope length and angle, calculate horizontal and vertical distance,
and display the results in real time. Typically, a refector prism is manually positioned at locations of
interest, and the data are recorded and processed by the total station. This technology can provide
real-time monitoring of predetermined points through use of one or more automated total stations
combined with multiple-prism targets, a data acquisition system, and sofware interfaces. For this
application, prism targets are mounted on the surface of the features to be monitored and are pro-
grammed into the routine of the total station. The survey frequency can vary (typically every two to
fve minutes) depending on the number of targets programmed into the total stations routine. Each
automated total station can monitor up to about 100 prism targets.
Traverse Lines and Triangulation
Traverse lines and triangulation are conventional survey techniques that have been used for decades.
Their improved accuracy in recent years is due to the use of more precise equipment used to mea-
sure distance (EDMs) and angles (theodolites) between reference control points. A traverse is used
to determine change in lateral position through measurement of successive distances and angles. If
a traverse returns to its starting point, the sum of the interior angles of the enclosed polygon can be
calculated and adjusted for measurement errors. Triangulation can also be used to determine change
in lateral position. This is accomplished by accurately measuring a baseline ofset from the surveyed
locations and the angles between the ends of baseline and the surveyed locations, as illustrated in
Figure 13.1. Then by periodic re-sighting on the surveyed locations from the ends of the baseline,
changes in lateral position can be calculated, assuming the baseline is located on stable ground.
Airborne Mapping Systems
Airborne methods include photogrammetric and LIDAR mapping systems. Photogrammetry is a
remote-sensing technology in which geometric properties of objects are determined from photo-
graphic images. Photogrammetric methods can be used to record movement of hundreds of survey
points at one time and thus provide an overall patern of deformation, but the accuracy is afected by
weather conditions, baseline measurements and interpreter skill.
Light Detection and Ranging (LIDAR) is a remote sensing system used to collect topographic data
and develop topographic maps. LIDAR consists of a laser imaging device, an inertial navigation
system, a GPS receiver and a computer. The technology can be used to map and determine coordi-
nates of dense paterns of ground points, which can be used to develop an image of the ground sur-
face. The data can be used to produce digital elevation models (DEMs) and subsequently topographic
maps. When LIDAR is used, weather conditions must be monitored because the fights cannot be
fown during times of rain or fog, as the water vapor in the air could cause the laser beams to scater
and give false readings.
Global Positioning System (GPS)
The GPS consists of a constellation of 24 satellites. Each satellite orbits the Earth twice a day at an
altitude of about 12,500 miles and continuously transmits information on specifc radio frequencies
< PREVIOUS VIEW
13-12
Chapter 13
MAY 2009
T
A
B
L
E
1
3
.
5
S
U
R
V
E
Y
I
N
G
M
E
T
H
O
D
S
M
e
t
h
o
d
A
d
v
a
n
t
a
g
e
s
L
i
m
i
t
a
t
i
o
n
s
A
p
p
r
o
x
i
m
a
t
e
A
c
c
u
r
a
c
y
R
e
l
a
t
i
v
e
C
o
m
p
l
e
x
i
t
y
C
o
s
t
A
p
p
l
i
c
a
b
i
l
i
t
y
t
o
C
o
a
l
R
e
f
u
s
e
D
i
s
p
o
s
a
l
F
a
c
i
l
i
t
i
e
s
E
l
e
v
a
t
i
o
n
s
b
y
o
p
t
i
c
a
l
l
e
v
e
l
i
n
g
F
a
s
t
,
p
a
r
t
i
c
u
l
a
r
l
y
w
i
t
h
s
e
l
f
-
l
e
v
e
l
i
n
g
e
q
u
i
p
m
e
n
t
U
s
e
s
w
i
d
e
l
y
a
v
a
i
l
a
b
l
e
e
q
u
i
p
m
e
n
t
F
i
r
s
t
o
r
d
e
r
l
e
v
e
l
i
n
g
r
e
q
u
i
r
e
s
h
i
g
h
-
g
r
a
d
e
e
q
u
i
p
m
e
n
t
a
n
d
c
a
r
e
f
u
l
a
d
h
e
r
e
n
c
e
t
o
s
t
a
n
d
a
r
d
p
r
o
c
e
d
u
r
e
s
T
h
i
r
d
o
r
d
e
r
:
0
.
0
5
f
t
S
e
c
o
n
d
o
r
d
e
r
:
0
.
0
2
5
t
o
0
.
0
3
3
F
i
r
s
t
o
r
d
e
r
:
0
.
0
1
2
t
o
0
.
0
2
0
S
i
m
p
l
e
L
o
w
t
o
M
o
d
e
r
a
t
e
R
o
u
t
i
n
e
f
o
r
a
l
l
e
m
b
a
n
k
m
e
n
t
s
T
r
i
g
o
n
o
m
e
t
r
i
c
l
e
v
e
l
i
n
g
L
o
n
g
r
a
n
g
e
;
f
a
s
t
a
n
d
c
o
n
v
e
n
i
e
n
t
;
c
a
n
b
e
d
o
n
e
s
i
m
u
l
t
a
n
e
o
u
s
l
y
w
i
t
h
t
r
a
v
e
r
s
i
n
g
A
c
c
u
r
a
c
y
i
s
i
n
f
u
e
n
c
e
d
b
y
a
t
m
o
s
p
h
e
r
i
c
c
o
n
d
i
t
i
o
n
s
;
r
e
q
u
i
r
e
s
a
v
e
r
y
a
c
c
u
r
a
t
e
m
e
a
s
u
r
e
m
e
n
t
o
f
z
e
n
i
t
h
a
n
g
l
e
T
h
i
r
d
o
r
d
e
r
:
0
.
0
5
f
t
S
e
c
o
n
d
o
r
d
e
r
:
0
.
0
2
5
t
o
0
.
0
3
3
S
i
m
p
l
e
f
o
r
3
r
d
O
r
d
e
r
;
M
o
d
e
r
a
t
e
f
o
r
2
n
d
O
r
d
e
r
L
o
w
f
o
r
3
r
d
O
r
d
e
r
;
M
o
d
e
r
a
t
e
f
o
r
2
n
d
O
r
d
e
r
R
o
u
t
i
n
e
f
o
r
a
l
l
e
m
b
a
n
k
m
e
n
t
s
D
i
s
t
a
n
c
e
m
e
a
s
u
r
i
n
g
b
y
t
a
p
i
n
g
D
i
r
e
c
t
m
e
a
s
u
r
e
m
e
n
t
s
R
e
q
u
i
r
e
s
c
l
e
a
r
,
r
e
l
a
t
i
v
e
l
y
f
a
t
s
u
r
f
a
c
e
b
e
t
w
e
e
n
m
e
a
s
u
r
i
n
g
p
o
i
n
t
s
a
n
d
r
e
f
e
r
e
n
c
e
d
a
t
u
m
;
m
o
v
e
m
e
n
t
c
a
n
o
n
l
y
b
e
m
e
a
s
u
r
e
d
i
n
o
n
e
d
i
r
e
c
t
i
o
n
;
m
o
n
u
m
e
n
t
s
m
u
s
t
b
e
l
o
c
a
t
e
d
a
l
o
n
g
a
s
t
r
a
i
g
h
t
l
i
n
e
w
i
t
h
r
e
a
d
y
a
c
c
e
s
s
b
e
t
w
e
e
n
p
o
i
n
t
s
;
t
a
p
e
s
h
o
u
l
d
b
e
c
h
e
c
k
e
d
f
r
e
q
u
e
n
t
l
y
a
g
a
i
n
s
t
s
t
a
n
d
a
r
d
;
e
x
c
e
p
t
f
o
r
s
h
o
r
t
m
e
a
s
u
r
e
m
e
n
t
s
,
t
a
p
i
n
g
h
a
s
b
e
e
n
r
e
p
l
a
c
e
d
b
y
E
D
M
.
T
h
i
r
d
o
r
d
e
r
:
0
.
0
0
3
3
t
o
0
.
0
0
1
6
o
f
d
i
s
t
a
n
c
e
b
e
t
w
e
e
n
i
n
s
t
r
u
m
e
n
t
a
n
d
s
u
r
v
e
y
e
d
l
o
c
a
t
i
o
n
S
e
c
o
n
d
o
r
d
e
r
:
0
.
0
0
0
0
5
0
.
0
0
0
0
2
o
f
d
i
s
t
a
n
c
e
F
i
r
s
t
o
r
d
e
r
:
0
.
0
0
0
0
0
3
o
f
d
i
s
t
a
n
c
e
S
i
m
p
l
e
L
o
w
t
o
M
o
d
e
r
a
t
e
R
o
u
t
i
n
e
f
o
r
a
l
l
e
m
b
a
n
k
m
e
n
t
s
E
l
e
c
t
r
o
n
i
c
d
i
s
t
a
n
c
e
m
e
a
s
u
r
e
m
e
n
t
(
E
D
M
)
L
o
n
g
r
a
n
g
e
;
f
a
s
t
a
n
d
c
o
n
v
e
n
i
e
n
t
;
v
e
r
y
a
c
c
u
r
a
t
e
.
A
c
c
u
r
a
c
y
i
s
i
n
f
u
e
n
c
e
d
b
y
a
t
m
o
s
p
h
e
r
i
c
c
o
n
d
i
t
i
o
n
s
F
o
r
d
i
s
t
a
n
c
e
:
0
.
0
0
1
t
o
0
.
0
3
f
t
F
o
r
l
a
t
e
r
a
l
p
o
s
i
t
i
o
n
c
h
a
n
g
e
b
y
t
r
i
a
n
g
u
l
a
t
i
o
n
:
0
.
0
0
5
t
o
0
.
0
3
f
t
S
i
m
p
l
e
t
o
M
o
d
e
r
a
t
e
L
o
w
t
o
M
o
d
e
r
a
t
e
R
o
u
t
i
n
e
f
o
r
a
l
l
e
m
b
a
n
k
m
e
n
t
s
O
f
f
s
e
t
s
f
r
o
m
b
a
s
e
l
i
n
e
u
s
i
n
g
t
h
e
o
d
o
l
i
t
e
a
n
d
s
c
a
l
e
D
i
r
e
c
t
m
e
a
s
u
r
e
m
e
n
t
s
R
e
q
u
i
r
e
s
b
a
s
e
l
i
n
e
u
n
a
f
f
e
c
t
e
d
b
y
m
o
v
e
m
e
n
t
0
.
0
0
1
t
o
0
.
0
0
5
f
t
S
i
m
p
l
e
L
o
w
t
o
M
o
d
e
r
a
t
e
R
o
u
t
i
n
e
f
o
r
a
l
l
e
m
b
a
n
k
m
e
n
t
s
m
i
l
e
s
m
i
l
e
s
m
i
l
e
s
m
i
l
e
s
m
i
l
e
s
< PREVIOUS VIEW
13-13
Instrumentation and Performance Monitoring
MAY 2009
T
A
B
L
E
1
3
.
5
S
U
R
V
E
Y
I
N
G
M
E
T
H
O
D
S
(
C
o
n
t
i
n
u
e
d
)
M
e
t
h
o
d
A
d
v
a
n
t
a
g
e
s
L
i
m
i
t
a
t
i
o
n
s
A
p
p
r
o
x
i
m
a
t
e
A
c
c
u
r
a
c
y
R
e
l
a
t
i
v
e
C
o
m
p
l
e
x
i
t
y
C
o
s
t
A
p
p
l
i
c
a
b
i
l
i
t
y
t
o
C
o
a
l
R
e
f
u
s
e
D
i
s
p
o
s
a
l
F
a
c
i
l
i
t
i
e
s
L
a
s
e
r
b
e
a
m
l
e
v
e
l
i
n
g
a
n
d
o
f
f
s
e
t
s
F
a
s
t
e
r
t
h
a
n
c
o
n
v
e
n
t
i
o
n
a
l
o
p
t
i
c
a
l
m
e
t
h
o
d
s
;
r
e
a
d
i
n
g
s
c
a
n
b
e
m
a
d
e
b
y
o
n
e
p
e
r
s
o
n
S
e
r
i
o
u
s
l
y
a
f
f
e
c
t
e
d
b
y
a
i
r
t
u
r
b
u
l
e
n
c
e
,
h
u
m
i
d
i
t
y
,
a
n
d
t
e
m
p
e
r
a
t
u
r
e
d
i
f
f
e
r
e
n
t
i
a
l
;
r
e
q
u
i
r
e
s
c
u
r
v
a
t
u
r
e
a
n
d
r
e
f
r
a
c
t
i
o
n
c
o
r
r
e
c
t
i
o
n
s
b
e
y
o
n
d
a
b
o
u
t
2
0
0
f
t
0
.
0
1
t
o
0
.
0
3
f
t
S
i
m
p
l
e
L
o
w
R
o
u
t
i
n
e
f
o
r
s
i
g
n
i
f
c
a
n
t
e
m
b
a
n
k
m
e
n
t
s
T
r
a
v
e
r
s
e
l
i
n
e
s
U
s
e
a
b
l
e
w
h
e
r
e
d
i
r
e
c
t
m
e
a
s
u
r
e
m
e
n
t
s
a
r
e
n
o
t
p
o
s
s
i
b
l
e
A
c
c
u
r
a
c
y
d
e
c
r
e
a
s
e
s
a
s
n
u
m
b
e
r
o
f
l
e
g
s
i
n
t
r
a
v
e
r
s
e
i
n
c
r
e
a
s
e
s
;
t
r
a
v
e
r
s
e
s
h
o
u
l
d
b
e
c
l
o
s
e
d
i
f
p
o
s
s
i
b
l
e
0
.
0
0
0
0
3
t
o
0
.
0
0
0
0
0
7
o
f
d
i
s
t
a
n
c
e
b
e
t
w
e
e
n
i
n
s
t
r
u
m
e
n
t
a
n
d
s
u
r
v
e
y
e
d
l
o
c
a
t
i
o
n
S
i
m
p
l
e
L
o
w
t
o
M
o
d
e
r
a
t
e
R
o
u
t
i
n
e
f
o
r
s
i
g
n
i
f
c
a
n
t
e
m
b
a
n
k
m
e
n
t
s
T
r
i
a
n
g
u
l
a
t
i
o
n
U
s
e
a
b
l
e
w
h
e
r
e
d
i
r
e
c
t
m
e
a
s
u
r
e
m
e
n
t
s
a
r
e
n
o
t
p
o
s
s
i
b
l
e
;
b
a
s
e
l
i
n
e
c
a
n
b
e
l
o
c
a
t
e
d
t
o
a
v
o
i
d
a
c
t
i
v
e
c
o
n
s
t
r
u
c
t
i
o
n
a
r
e
a
s
R
e
q
u
i
r
e
s
a
c
c
u
r
a
t
e
m
e
a
s
u
r
e
-
m
e
n
t
o
f
a
n
g
l
e
s
a
n
d
b
a
s
e
l
i
n
e
l
e
n
g
t
h
0
.
0
0
0
0
3
t
o
0
.
0
0
0
0
0
1
o
f
d
i
s
t
a
n
c
e
b
e
t
w
e
e
n
i
n
s
t
r
u
m
e
n
t
a
n
d
s
u
r
v
e
y
e
d
l
o
c
a
t
i
o
n
M
o
d
e
r
a
t
e
L
o
w
t
o
M
o
d
e
r
a
t
e
R
o
u
t
i
n
e
f
o
r
s
i
g
n
i
f
c
a
n
t
e
m
b
a
n
k
m
e
n
t
s
P
h
o
t
o
g
r
a
m
m
e
t
r
i
c
m
e
t
h
o
d
s
C
a
n
r
e
c
o
r
d
m
o
v
e
m
e
n
t
o
f
h
u
n
d
r
e
d
s
o
f
p
o
t
e
n
t
i
a
l
p
o
i
n
t
s
a
t
o
n
e
t
i
m
e
f
o
r
d
e
t
e
r
m
i
n
a
t
i
o
n
o
f
o
v
e
r
a
l
l
d
e
f
o
r
m
a
t
i
o
n
p
a
t
t
e
r
n
W
e
a
t
h
e
r
c
o
n
d
i
t
i
o
n
s
c
a
n
l
i
m
i
t
u
s
e
;
i
n
t
e
r
p
r
e
t
a
t
i
o
n
r
e
q
u
i
r
e
s
s
p
e
c
i
a
l
i
s
t
;
f
o
r
g
o
o
d
a
c
c
u
r
a
c
y
t
h
e
b
a
s
e
l
i
n
e
s
h
o
u
l
d
n
o
t
b
e
l
e
s
s
t
h
a
n
o
n
e
-
f
f
t
h
o
f
t
h
e
s
i
g
h
t
d
i
s
t
a
n
c
e
0
.
0
0
0
2
t
o
0
.
0
0
0
0
1
f
t
C
o
m
p
l
e
x
E
x
p
e
n
s
i
v
e
N
o
n
-
r
o
u
t
i
n
e
f
o
r
s
i
g
n
i
f
c
a
n
t
e
m
b
a
n
k
m
e
n
t
s
G
l
o
b
a
l
p
o
s
i
t
i
o
n
i
n
g
s
y
s
t
e
m
(
G
P
S
)
O
p
e
r
a
t
e
s
w
i
t
h
l
i
t
t
l
e
a
t
t
e
n
t
i
o
n
f
r
o
m
p
e
r
s
o
n
n
e
l
;
c
a
n
b
e
s
e
t
t
o
t
r
i
g
g
e
r
a
w
a
r
n
i
n
g
d
e
v
i
c
e
;
v
e
r
y
a
c
c
u
r
a
t
e
;
d
o
e
s
n
o
t
r
e
q
u
i
r
e
l
i
n
e
o
f
s
i
g
h
t
;
m
e
a
s
u
r
e
-
m
e
n
t
s
c
a
n
b
e
m
a
d
e
i
n
a
l
m
o
s
t
a
n
y
w
e
a
t
h
e
r
c
o
n
d
i
t
i
o
n
R
e
q
u
i
r
e
s
o
p
e
n
s
k
y
l
i
n
e
o
f
s
i
g
h
t
S
t
a
t
i
c
a
n
d
R
e
l
a
t
i
v
e
P
o
s
i
t
i
o
n
i
n
g
:
(
0
.
0
1
6
f
t
+
1
p
p
m
)
R
e
q
u
i
r
e
s
1
0
t
o
3
0
m
i
n
u
t
e
s
o
b
s
e
r
v
a
t
i
o
n
p
e
r
p
o
i
n
t
d
e
p
e
n
d
i
n
g
o
n
m
e
t
h
o
d
o
f
s
u
r
v
e
y
a
n
d
w
h
e
t
h
e
r
s
i
n
g
l
e
o
r
d
u
a
l
f
r
e
q
u
e
n
c
y
r
e
c
e
i
v
e
r
i
s
u
s
e
d
.
C
o
m
p
l
e
x
M
o
d
e
r
a
t
e
t
o
E
x
p
e
n
s
i
v
e
N
o
n
-
r
o
u
t
i
n
g
e
f
o
r
s
i
g
n
i
f
c
a
n
t
e
m
b
a
n
k
m
e
n
t
s
(
A
D
A
P
T
E
D
F
R
O
M
D
U
N
N
I
C
L
I
F
F
,
1
9
9
3
)
N
o
t
e
:
=
s
q
u
a
r
e
r
o
o
t
o
f
d
i
s
t
a
n
c
e
i
n
m
i
l
e
s
m
i
l
e
s
< PREVIOUS VIEW
13-14
Chapter 13
MAY 2009
to ground-based receivers. Using sophisticated receivers and data-analysis techniques, receiver posi-
tions can be determined. The GPS satellites continuously transmit an estimate of their position, digi-
tal codes, and a precise time signal. A GPS receiver uses an internal clock and the codes to determine
the distances to at least 4 satellites. Distance is calculated by multiplying the time it takes the radio
signals to reach the receiver times the speed of light. Knowing where the satellites are located when
signals are transmited, the receiver calculates its position. GPS equipment is becoming more reliable,
cheaper, faster, and easier to use compared to conventional instruments. New hardware, feld proce-
dures and sofware have also been developed to assist users in data collection and processing. GPS
equipment is now used for a wide range of monitoring applications.
A stable reference datum is required for all survey measurements pertaining to deformation. A
benchmark is a reference datum for vertical movements and a horizontal control station or reference
monument is a reference datum for horizontal movements. As a minimum, the datum benchmark
should be placed in the ground, on a structure or on a rock outcrop in such a manner that the efects
of weather (frost heave) or equipment travel will not alter its position. It should also be located well
away from the embankment area so that increased earth loads or water pressures in the embank-
ment do not signifcantly change the ground elevation at the benchmark. Where rock is shallow, an
excavated hole to rock with concrete backfll and an embedded brass pin reference point is normally
acceptable. When rock is deep, the most reliable benchmark consists of a steel rod installed inside a
casing placed through overburden into bedrock. The rod should be isolated from the casing by nylon
spacers, and the casing should be provided with a protective cover, as shown in Figure 13.2. Because
the outer casing may be disturbed or compressed due to setlement of the surrounding material, the
FIGURE 13.5 LOCATION OF MONUMENTS USING TRIANGULATION
LEGEND
MONUMENT
TRIANGULATION POINT FOR
INSTRUMENT SETUP
BENCH
CREST OF EMBANKMENT
TOE OF
EMBANKMENT
KNOWN FIXED BASELINE
MEASURED ANGLE FOR
TRIANGULATION (TYP.)
FIGURE 13.1 LOCATION OF MONUMENTS USING TRIANGULATION
FIGURE 13.1 LOCATION OF MONUMENTS USING TRIANGULATION
< PREVIOUS VIEW
13-15
Instrumentation and Performance Monitoring
MAY 2009
FIGURE 13.2 EXAMPLE OF BENCHMARK IN ROCK
FIGURE 13.6 EXAMPLE OF ROCK BENCHMARK
13.2b ROCK BENCHMARK IN COMMON
HOLE WITH PIEZOMETER
PIEZOMETER
PROTECTIVE PIPE
STEEL ROD
HINGED
LID
THREADED CAP
ELASTIC MEMBRANE
TO KEEP HOLE CLEAN
PROTECTIVE
PIPE COVER
GREASED
SURFACE
STEEL ROD
OR PIPE
GROUT
CASING
CEMENT GROUT
OR CONCRETE
GROUND
SURFACE
PADLOCK
13.2a TYPICAL ARRANGEMENT
BEDROCK
FIGURE 13.2 EXAMPLE OF BENCHMARK IN ROCK
< PREVIOUS VIEW
13-16
Chapter 13
MAY 2009
benchmark should be placed on the top of the inner rod, which is isolated from the movement of the
casing. The protective cover will prevent damage due to material becoming wedged between the
inner rod and the casing or due to vandalism.
In areas with nearby mining activity, associated movements of the ground surface may occur on
adjacent hillsides (both surface heave and subsidence have been observed to occur for horizontal dis-
tances of more than several hundred feet from the nearest point of mining), so it may be necessary to
construct a benchmark in a deep hole so that mining in the coal seam does not afect the benchmark.
If the depth of mining is such that the cost of constructing a deep benchmark is prohibitive, a project
benchmark should be established and should be periodically checked against other benchmarks in
the region. In all cases, a minimum of three datum benchmarks should be established for periodic
checking. These benchmarks should be located on opposite sides of the facility and in locations where
they will not be afected by weather or subsidence conditions.
Reference points or monuments where surface setlement monitoring is desired can range from
stakes or rods driven into the ground to carefully installed concrete cylinders. Examples are shown
in Figure 13.3. In all cases, the point of measurement must be clearly identifed so that the same loca-
tion is used each time. Brass or steel rods or pipes set in concrete or rock or chiseled/painted crosses
are suitable.
Except for rapidly established and very short-duration monitoring programs, driven rods should be
avoided because this type of arrangement can be easily disturbed. If used, the minimum depth of
embedment should be at least two feet, and the maximum extension above grade should be one foot
or less to minimize tilting or bending of the rod. Also, because setlement points are ofen located in
areas of high equipment activity, they should be well marked with stakes and fags to minimize loss
and disturbance. Where loss of a monument would reduce the extent or accuracy of an important
survey, up to twice the number believed necessary may be installed. The monument shown on Figure
13.3c can be used where surface creep is signifcant and conditions at depth (e.g., below the frost line
or in rock or stable strata below the ground surface) beter represent slope performance.
13.2.1.2.2 Surface Extensometers
Surface extensometers are devices used to monitor changes in distance between two points at the
ground surface or on a structure (Dunniclif, 1993). Surface extensometers are usually referred to as
crack gages or convergence gages. Crack gages are used to monitor tension cracks behind slopes,
across a joint or fault in a rock mass, or in concrete structures. Convergence gages have been used
to measure the displacement between two surface monuments as part of the monitoring of ground
surface strain induced by longwall mining near an impoundment. Mechanical crack gages include
pins and tape, pins and steel ruler or caliper, pins and mechanical extensometer, and grid crack moni-
tors. For each of these devices, the pins or grid crack elements are anchored on opposite sides of the
discontinuity, and the distance between the anchored elements is measured periodically to determine
the change in separation. Each of these types of surface extensometer is inexpensive and has a pre-
cision ranging from 0.01 to 0.1 inches. The principal limitation of mechanical crack gages is the
relatively short span length between the pins. Additional details related to surface extensometers are
provided in Dunniclif (1993).
13.2.1.2.3 Tiltmeters
Tiltmeters are used for monitoring changes in inclination or rotation at the ground surface or on a
structure in either a horizontal or vertical plane (Dunniclif, 1993). A tiltmeter consists of a gravity-
sensing device that is sealed in a housing. Some tiltmeters are fxed in place, and others are portable
and can be used to monitor multiple locations by mounting on a fxture that is permanently atached
to the monitoring location. The most common types of tiltmeters have an accelerometer transducer,
< PREVIOUS VIEW
13-17
Instrumentation and Performance Monitoring
MAY 2009
FIGURE 13.3 TYPICAL MONUMENT INSTALLATIONS
FIGURE 13.7 TYPICAL MONUMENT INSTALLATIONS
B
E
L
O
W
F
R
O
S
T
L
I
N
E
O
R
P
L
A
C
E
D
I
N
R
O
C
K
CONCRETE
BRASS PLATE
MONUMENT OR
STEEL DOWEL
13.3d
C
L CONCRETE
GROUT
BOTTOM OF ROD
OR PIPE SHOULD
EXTEND BELOW
FROST LINE
STEEL ROD
OR PIPE
13.3a
C
L
ELASTIC
MEMBRANE
TO KEEP
HOLE CLEAN
THIN WALL
PIPE SLEEVE
THREADED
CAP
STEEL ROD
OR PIPE
~ 5'
~ 3'
13.3b
BRASS PLATE
MONUMENT
PUNCH MARK
THIN WALL
PIPE SLEEVE
CONCRETE
B
E
L
O
W
F
R
O
S
T
L
I
N
E
O
R
P
L
A
C
E
D
I
N
R
O
C
K
13.3c
FIGURE 13.3 TYPICAL MONUMENT INSTALLATIONS
< PREVIOUS VIEW
13-18
Chapter 13
MAY 2009
but devices with mechanical, vibrating wire and electrolytic transducers are also available. The typi-
cal range of accelerometer-type tiltmeters is 30 from horizontal or vertical; the precision is typically
50 arc-seconds; and the temperature sensitivity is typically in the range of 2 to 3 arc-seconds/F
(Dunniclif, 1993).
13.2.1.2.4 Probe Extensometers
Probe extensometers are used to monitor changes in distance between two or more locations along a
common axis by passing a probe through a pipe (Dunniclif, 1993). As the probe passes through the
pipe, it mechanically or electrically detects the measuring locations, and the distance between them
is determined by physical measurement of the probe positions. If an exact measurement of probe
location is required, one measuring point must be accessible so that its location relative to a reference
datum can be determined by surveying methods. Depending on project requirements, the pipe may
be vertical for measurement of setlement or heave, horizontal for measurement of lateral displace-
ment, or inclined. Dunniclif (1993) provides a complete summary of probe extensometer types, their
advantages, limitations and approximate accuracy.
The most commonly used probe extensometers are:
Gage with current-displacement induction coil
Magnetic reed switch gage
Combined probe extensometers and inclinometer casing
Induction-coil transducers are described in Appendix 13A.2.4.6. When used as a probe extensometer,
the device consists of a series of steel rings atached to and surrounding a telescoping or corrugated
plastic pipe and a reading device consisting of a primary coil housed in a probe that is atached to
a signal cable and current indicator. The pipe is installed in a borehole and backflled. When the
probe is inserted into the pipe, the steel rings can be located because the current indicator displays
a maximum value at each ring location. The measurements enable determination of the distance
between each ring, and a series of measurements over time permit determination of rate of deforma-
tion between the rings. Information related to installation, measurement precision and typical appli-
cations is provided in Dunniclif (1993).
The magnetic reed switch gage is described in Appendix 13A.2.4.5. When used as a probe extensom-
eter, the device consists of a series of circular magnetic anchors surrounding a rigid or telescoping
plastic pipe. The pipe is installed in a borehole and backflled. When the probe is inserted into the
pipe, the magnetic anchors are located and their position measured as the coupled oscillator in the
probe is activated at each ring location.
The Sondex system consists of a probe, a signal cable, a cable reel with a built-in voltmeter, and a
number of stainless steel sensing rings. A survey tape is typically connected to the probe. The sens-
ing rings are fxed to a continuous length of corrugated plastic pipe that is installed coaxially with
inclinometer casing or access pipe. The annulus between the corrugated pipe and the borehole wall
is grouted. The corrugated pipe and its sensing rings setle with the surrounding ground. To obtain
measurements, the operator draws the probe through the access pipe. A buzzer sounds when the
probe is near a ring, and the voltmeter indicates peaks when the probe is aligned with a ring. The
operator refers to the survey tape and records the depth of the ring. A sensitivity adjustment allows
operation adjacent to steel pipes, piles, or other metal objects. Setlement and heave can be calculated
by comparing the depth of each ring to its initial depth.
Several types of probe extensometers can be used in conjunction with inclinometer casing in a vertical
borehole to permit measurement of both horizontal and vertical deformations in one installation. Increx
< PREVIOUS VIEW
13-19
Instrumentation and Performance Monitoring
MAY 2009
and Sondex are examples of this type of probe extensometer. The Increx system consists of a number
of brass rings that are positioned at one-meter intervals along an inclinometer casing. The probe is
positioned to successively take readings between each pair of rings. Periodic surveys are compared
to the initial survey to determine changes in the distance between rings. If the distance has increased,
extension has occurred; if the distance has decreased, compression has occurred. Movements can be
referenced to the deepest ring, if it is located in stable ground, or the top of the casing can be optically
surveyed. The systems can measure changes as small as 0.001 mm with an accuracy of 0.01 mm/m.
A schematic illustration of a probe extensometer is provided in Figure 13.4.
13.2.1.2.5 Fixed Embankment Extensometers
Fixed embankment extensometers are devices placed in an embankment fll during construction to
monitor changes in distance between two or more points along a common axis without a movable
probe. They are used to measure setlement, horizontal deformation or strain (Dunniclif, 1993). The
most common type of fxed embankment extensometer is a setlement platform. A setlement plat-
form consists of a square plate or anchorage to which a riser is atached. If the fll height exceeds
about 25 feet, the riser should be isolated from the surrounding fll by an outer pipe within which
the riser can move freely. As the height of fll increases, the riser and outer pipe are raised by adding
additional pipe lengths. Movements of the setlement platform can be monitored by optical survey of
elevation of the top of the riser. A schematic illustration of a setlement platform is provided in Figure
13.5. Setlement platforms are easy to construct, but they are easily damaged during construction as
fll is placed and compacted near them. This problem can be avoided by using liquid- level gages, as
described in Appendix 13A.3.2.
FIGURE 13.8 PROBE EXTENSOMETER
READOUT UNIT
SURVEY TAPE
PROBE CONTAINING
REED SWITCH
BOREHOLE FILLED
WITH GROUT
FLUSH-COUPLED RIGID
PVC ACCESS PIPE
SPRING
CIRCULAR
MAGNET
MAGNET ANCHOR
(SPIDER MAGNET)
1345
1344590L5
(DUNNICLIFF, 1993)
FIGURE 13.4 PROBE EXTENSOMETER
FIGURE 13.4 PROBE EXTENSOMETER
< PREVIOUS VIEW
13-20
Chapter 13
MAY 2009
Tensioned-wire gages and fxed embankment extensometers with Linear Variable Diferential
Transformers (LVDTs) have been used to monitor horizontal displacements in an embankment fll
(Dunniclif, 1993). These devices have been largely replaced by probe or in-place inclinometers
(Section 13.2.1.2.7) installed in a vertical borehole.
13.2.1.2.6 Fixed Borehole Extensometers
Fixed borehole extensometers are devices without moveable probes installed in boreholes in soil or
rock and used to monitor the change in distance between two or more points along the borehole axis
(Dunniclif, 1993). If measurement of the exact deformation at a specifc location is required, one end
of the extensometer must be fxed in stable ground or it must be accessible for optical survey. Fixed
borehole extensometers can be used to monitor deformations behind rock slopes, foundation setle-
ments, the progression of subsidence above underground mines, and heave at the base of open-cut
excavations. A schematic illustration of a fxed borehole extensometer is provided in Figure 13.6.
The distance between the end of the rod and the face of the collar anchor can be measured using either
a mechanical device (e.g., dial gage) or electrical transducer (e.g., LVDT). The device illustrated in
Figure 13.6 is a single-point borehole extensometer (SPBX). A multiple-point borehole extensometer
(MPBX) consists of several downhole anchors within a single borehole along with a rod and measur-
FIGURE 13.9 SETTLEMENT PLATFORM
SURFACE OF
EMBANKMENT
COUPLED
RISER PIPE
EMBANKMENT
SQUARE PLATE
ORIGINAL
GROUND
SURFACE
SETTLEMENT OF PLATFORM
IS DETERMINED BY MEASURING
ELEVATION OF THE TOP OF THE
RISER PIPE, USING SURVEYING
METHODS
(DUNNICLIFF, 1993)
FIGURE 13.5 SETTLEMENT PLATFORM
FIGURE 13.5 SETTLEMENT PLATFORM
< PREVIOUS VIEW
13-21
Instrumentation and Performance Monitoring
MAY 2009
ing device for each. MPBXs allow the measurement of deformation or strain paterns over the length
of the borehole so that potential failure or large deformation zones can be identifed. Fixed borehole
extensometers can be manufactured to meet site-specifc requirements. The principal feature difer-
ences include choice of rod or tensioned wire atached to the downhole anchor, downhole anchor
type (i.e., mechanical, hydraulic or groutable), SPBX or MPBX, transducer type and extensometer
head. Dunniclif (1993) provides guidance for selecting among these options for particular applica-
tions. The accuracy of the device depends on the type of mechanical or electrical transducer used to
measure change in the anchor location.
13.2.1.2.7 Inclinometers
Inclinometers are devices used for monitoring displacements normal to the axis of a pipe by passing a
probe along the pipe (Dunniclif, 1993). They can be used to determine the depth and profle of lateral
displacement above and below a slide plane such as in a distressed slope. A gravity-sensing transducer
in the probe measures inclination with respect to vertical. Normally the probe is inserted into a vertical
or inclined borehole to measure lateral displacements in landslide zones or structures, but probes are
available for use in horizontal pipe for obtaining profles beneath embankments and structures.
Inclinometer systems have four primary components:
A probe casing with tracking grooves that is permanently installed in a borehole
(vertical or inclined application) or trench (horizontal application)
A probe with a pair of wheel carriages housed in steel to confne the gravity-sensing
transducer
A portable readout for power supply and measurement of probe inclination
A graduated electrical cable connecting the probe and readout unit
Figure 13.7 shows a schematic representation of the operation of an inclinometer for vertical and
near- vertical applications. The inclinometer casing is installed and grouted along its full length into
a borehole so that one pair of the four orthogonal grooves in the casing is normal to the displace-
FIGURE 13.10 FIXED BOREHOLE EXTENSOMETER
L
SLEEVED
STEEL ROD
BOREHOLE
DOWNHOLE
ANCHOR
COLLAR
ANCHOR
M
E
A
S
U
R
E
D
IS
T
A
N
C
E
F
R
O
M
F
A
C
E
O
F
C
O
L
L
A
R
A
N
C
H
O
R
T
O
E
N
D
O
F
R
O
D
:
C
H
A
N
G
E
IN
D
IS
T
A
N
C
E
=
L
L MEASURED WITH
DIAL GAUGE OR LVDT
(DUNNICLIFF, 1993)
FIGURE 13.6 FIXED BOREHOLE EXTENSOMETER
FIGURE 13.6 FIXED BOREHOLE EXTENSOMETER
< PREVIOUS VIEW
13-22
Chapter 13
MAY 2009
FIGURE 13.11 INCLINOMETER OPERATION
PROBE CONTAINING
GRAVITY-SENSING
TRANSDUCER
READOUT
UNIT
GRADUATED
ELECTRICAL
CABLE
BOREHOLE
COUPLING
BACKFILL
GROUT
1345
1344590L5
GUIDE
CASING
13.7a INSTRUMENT AND CASING ARRANGEMENT
ACTUAL ALIGNMENT
OF GUIDE CASING
(EXAGGERATED)
L sin
L sin
L
DISTANCE BETWEEN
SUCCESSIVE READINGS
TRUE
VERTICAL
and K
a
Correction Factors, Proceed-
ings, NCEER Workshop on Evaluation of Liquefaction Resistance of Soils, T.L. Youd and I.M. Idriss, eds.,
NCEER-97-0022, National Center for Earthquake Engineering Research, SUNY, Bufalo, pp. 167-190.
Harmsen, S., D. Perkins, and A. Frankel, 1999, Deaggregation of Probabilistic Ground Motions in the
Central and Eastern United States, Bulletin of the Seismological Society of America, Vol. 89, No. 1, pp. 1-13.
Harr, M.E., 1962, Groundwater and Seepage, McGraw-Hill, New York, NY.
Harteis, S.P. and D.R. Dolinar, 2006, Water and Slurry Bulkheads in Underground Coal Mines:
Design, Monitoring and Safety Concerns, Mining Engineering, Vol. 58, No. 12, pp. 41-47.
* Harteis, S.P., D.R. Dolinar and T.M. Taylor, 2008, Guidelines for Permiting Construction, and Monitor-
ing of Retention Bulkheads for Underground Coal Mines, Information Circular 9506, U.S. Department
of the Interior, National Institute for Occupational Health and Safety, Pitsburgh, PA.
Hartman, H.L., ed., 1992, Mining Engineering Handbook, 2nd Ed., Society for Mining, Metallurgy,
and Exploration (SME), Litleton, CO.
Haxo, H.E., Jr., Quality Assurance of Geomembranes Used as Linings for Hazardous Waste
Containment, Geotextiles and Geomembranes, International Geosynthetics Society, Vol. 3, No. 4,
pp. 225-247.
Head, K.H., 1986, Manual of Soil Laboratory Testing, Volume 3: Efective Stress Tests, Pentech Press,
London, UK.
Head, K.H., 1982, Manual of Soil Laboratory Testing, Volume 2: Permeability, Shear Strength and Com-
pressibility Tests, Pentech Press, London, UK.
Head, K.H., 1980, Manual of Soil Laboratory Testing, Volume 1: Soil Classifcation and Compaction Tests,
Pentech Press, London, UK.
* Hedin, R.S., G.R. Watzlaf and R.W. Naim, 1994, Passive Treatment of Acid Mine Drainage with
Limestone, Journal of Environmental Quality, Vol. 23, No. 6, pp. 1338-1345.
* Hedin, R.S., R.W. Naim and R.L.P. Kleinmann, 1993, Passive Treatment of Coal Mine Drainage,
Information Circular 9389, U.S. Department of the Interior, Bureau of Mines, Pitsburgh Research
Center, Pitsburgh, PA.
* Hegazy, Y.A., A.G. Cushing and C.J. Lewis, 2004, Physical, Mechanical, and Hydraulic Properties
of Coal Refuse for Slurry Impoundment Design, Proceedings, 2nd International Site Characterization
Conference, Porto, Portugal.
Henderson, F.M., 1966, Open Channel Flow, The Macmillan Company, New York, NY.
Henry, C.D., 2007, Fine Coal Recovery Plant Installation: Smith Branch Impoundment Pinnacle
Mining, Proceedings, Coal Prep Annual Coal Processing Exhibition and Conference, Coal Preparation
Society of America, Lexington, KY.
* Hershfeld, D.M., 1961, Rainfall Frequency Atlas of the United States for Durations from 30 Minutes to
24 Hours and Return Periods from 1 to 100 Years, Technical Paper No. 40, U.S. Department of Com-
merce, Weather Bureau, Washington, DC.
Herzog B.L., 1994, Slug Tests for Determining Hydraulic Conductivity of Natural Geologic
Deposits, Hydraulic Conductivity and Waste Contaminant Transport In Soil: ASTM STP 1142, D.E.
Daniel and S.J. Trautwein, eds., ASTM, Philadelphia, PA, pp. 95-110.
Himelick, E.B., 1981, Tree and Shrub Transplanting Manual, 2nd Ed., International Society of Arbori-
culture, Champaign, IL.
< PREVIOUS VIEW
R14
References
Hoek, E., 2000, Practical Rock Engineering, Rocscience, Toronto, Ontario, Canada.
Holtz, R.D. and W.D. Kovacs, 1981, Introduction to Geotechnical Engineering, Prentice-Hall, Engle-
wood Clis, NJ.
* Holtz, R.D., B.R. Christopher and R.R. Berg, 1998, Geosynthetic Design and Construction Guidelines, DTFH61-
93-C-00120, Federal Highway Administration (FHWA), National Highway Institute, Washington, DC.
Houlsby, A.C., 1990, Construction and Design of Cement Grouting A Guide To Grouting In Rock Foun-
dations, John Wiley & Sons, New York, NY.
Howard, A.K., 1977, Modulus of Soil Reaction for Buried Flexible Pipe, Journal of Geotechnical
Engineering, ASCE, Vol. 103, No. GT1, pp. 33-43.
Howard, A.K. and J.L. Hitch, eds., 1998, The Design and Application of Controlled Low-Strength
Materials (Flowable Fill), ASTM STP 1331, American Society for Testing and Materials. West Con-
shohocken, PA.
* Huang, Y.H., J. Li and G. Weeratunga, 1987, Strength and Consolidation Characteristic of Fine
Coal Refuse, University of Kentucky, Lexington, KY.
Hunt, R.E., 1986, Geotechnical Engineering Practices, McGraw-Hill, New York, NY.
Hustrulid, W.A., 1976, A Review of Coal Pillar Strength Formulas, Rock Mechanics, Vol. 8, No. 2,
pp. 115-145.
Hutchison, I.P.G. and R.D. Ellison, eds., 1992, Mine Waste Management: A Resource for Mining Indus-
try Professionals, Regulators and Consulting Engineers, Lewis Publishers, Chelsea, MI.
Hvorslev, J., 1948, Subsurface Exploration and Sampling of Soils for Civil Engineering Purposes, U.S.
Department of the Army, Corps of Engineers, Waterways Experiment Station, Vicksburg, MS.
* Hynes-Gri n, M.E. and A.G. Franklin, A.G., 1984, Rationalizing the Seismic Coe cient Method, Mis-
cellaneous Paper GL-84-13, U.S. Department of the Army, Corps of Engineers, Waterways Experi-
ment Station, Vicksburg, MS.
* Idriss, I.M. and R.W. Boulanger, 2007, Residual Shear Strength of Liquied Soil, Proceedings,
Modernization and Optimization of Existing Dams and Reservoirs, 27th Annual USSD Conference,
Denver, CO, pp. 621-634.
Indraratna, B. and S. Radampola, 2002, Analysis of Critical Hydraulic Gradient for Particle Movement
in Filtration, Journal of Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 128, No. 4, pp. 337-350.
Jaeger, J.C., N.G.W. Cook and R. Zimmerman, 2007, Fundamentals of Rock Mechanics, 4th Ed., John
Wiley & Sons (Blackwell Publishing), Hoboken, NJ.
Jky, J., 1944, The Coe cient of Earth Pressure at Rest, Journal for Society of Hungarian Architects
and Engineers, Budapest, Hungary, pp. 355-358.
Janbu, N., 1973, Slope Stability Computations, Embankment-Dam Engineering: Casagrande Volume,
R.C. Hirschfeld and S.J. Poulos, eds., John Wiley & Sons, Hoboken, NJ, pp. 47-86.
* Johnson, W.J., 2003, Applications of the Electrical Resistivity Method for Detection of Under-
ground Mine Workings, Proceedings, Workshop on Geophysical Technologies for Detecting Under-
ground Coal Mine Voids, Lexington, KY.
* Johnson, W.J. and J.C. Clark, 1992, High Resolution Shear Wave Surveying for Hydrogeological Inves-
tigations, Report No. DOE/CH-9211 (Contract No. 02112405 with Argonne National Laboratory),
U.S. Department of Energy, Washington, DC.
* Johnson, W.J., R.E. Snow and J.C. Clark, 2002, Surface Geophysical Methods for the Detection of
Underground Mine Workings, Proceedings, Symposium on Geotechnical Methods for Mine Mapping
Verications, Charleston, WV.
REV. AUG. 2010 < PREVIOUS VIEW
R15
Engineering and Design Manual Coal Refuse Disposal Facilities
* Jolley, T.R. and H.W. Russell, 1959, Control of Fires in Inactive Coal Deposits in Western U.S. Including
Alaska, 1948-1958, Report No. BM-IC-7932, U.S. Department of the Interior, Bureau of Mines, Wash-
ington, DC.
Jones, C.L., J.D. Higgins and R.D. Andrew, 2000, Colorado Rockfall Simulation Program, Version 4.0,
Colorado Department of Transportation, Colorado Geological Survey, Denver, CO.
* Kalamaras, G.S. and Z.T. Bieniawski, 1993, A Rock Mass Strength Concept for Coal Seams, Pro-
ceedings, 12th International Conference on Ground Control in Mining, West Virginia University, Mor-
gantown, WV, pp. 274-283.
Kalinski, M.E. and J.L. Phillips, 2008, Development of Methods to Predict Dynamic Behavior of
Fine Tailings: Preliminary Results from Two Sites in Appalachia, Geotechnical Special Publication
No. 181, Geotechnical Earthquake Engineering and Soil Dynamics IV, D. Zeng, M.T. Manzari and
D.R. Hiltunen, eds., American Society of Civil Engineers (ASCE), Reston, VA.
Kanamori, H., 1977, The Energy Release in Great Earthquakes, Journal of Geophysical Research,
Vol. 82, No. B20, pp. 2981-2987.
Karmis, M., 2003, Milos Statement Sustainable Development Conference Held in Greece,
Mining Engineering, Society of Mining, Metallurgy, and Exploration (SME), Litleton, CO, Vol. 55,
No. 8, pp. 31-32.
Karmis, M. and Z Agioutantis, 2004, A Risk Analysis Subsidence Approach for the Design of Coal
Refuse Impoundments Overlying Mine Workings, Proceedings, 8th International Symposium on
Environmental Issues and Waste Management in Energy and Mineral Production (SWEMP), Antalya,
Turkey, pp. 205-210.
Kashef, A.I., 1986, Groundwater Engineering, McGraw-Hill, New York, NY,
* Kaufman, W.W. and J.C. Ault, 1977, Design of Surface Mine Haulage Roads A Manual, BM-IC-8758,
U.S. Department of the Interior, Bureau of Mines, Washington, DC.
* Kendorski, F. S., 2006, Efect of Full-Extraction Underground Mining on Ground and Surface
Waters: A 25-Year Retrospective, Proceedings, 25th International Conference on Ground Control in
Mining, West Virginia University, Morgantown, WV, pp. 425-430.
* Kendorski, F. S., 1993, Efect of High-Extraction Coal Mining on Surface and Ground Water, Pro-
ceedings, 12th International Conference on Ground Control in Mining, West Virginia University, Mor-
gantown, WV, pp. 412-425.
* Kendorski, F.S., I. Khosla and M.M. Singh, 1979, Criteria for Determining When a Body of Surface
Water Constitutes a Hazard to Mining, Final Report, Contract J0285011, U.S. Department of the Inte-
rior, Bureau of Mines, Pitsburgh, PA.
* Kepler, D.A. and E.C. McCleary, 1994, Successive Alkalinity-Producing Systems (SAPS) for Treat-
ment of Acid Mine Drainage, Proceedings, International Land Reclamation and Mine Drainage Confer-
ence, SP06A-94, U.S. Department of the Interior, Bureau of Mines, Pitsburgh, PA, pp. 196-204.
Keys, W.S., 1997, A Practical Guide to Borehole Geophysics in Environmental Investigations, CRC Press,
Boca Raton, FL.
* Kim, A.G. and R.F. Chaiken, 1993, Fires in Abandoned Coal Mines and Waste Banks, Information Cir-
cular No. 9352, U.S. Department of the Interior, Bureau of Mines, Washington, DC.
Kim, A.G. and R.F. Chaiken, 1990, Relative Self-Heating Tendencies of Coal, Carbonaceous Shales
and Coal Refuse, Proceedings, 1990 Mining and Reclamation Conference and Exhibition, J. Skousen
and J. Sencindiver, eds., West Virginia University, Morgantown, WV.
* Kim, A.G. and A.M. Kociban, 1994, Cryogenic Slurry Method to Extinguish Waste Bank Fires,
Proceedings, International Land Reclamation and Mine Drainage Conference, U.S. Department of the
Interior, Bureau of Mines, Pitsburgh, PA, pp. 129-138.
< PREVIOUS VIEW
R16
References
Kirkwood, D.T. and K.K. Wu, 1995, Technical Considerations for the Design and Construction of
Mine Seals to Withstand Hydraulic Heads in Underground Mines, SME Preprint No. 95-100, Soci-
ety for Mining, Metallurgy, and Exploration (SME) Annual Meeting, Denver, CO.
* Kleinmann, R.L.P., ed., 2000, Prediction of Water Quality at Surface Coal Mines, National Mine Recla-
mation Center, West Virginia University, Morgantown, WV.
Koerner, R.M., 2006, Designing with Geosynthetics, 5th Ed., Prentice Hall, Upper Saddle River, NJ.
* Koerner, R.M. and G.R. Koerner, 2005, GSI White Paper #4, Reduction Factors Used in Geosyn-
thetic Design, GFR Magazine, Vol. 23, No. 2 (February 2005). Revised version (2007) available from
the Geosynthetic Institute, Folsom, PA.
* Kohli, K. and F. Block, 2007, Guidance Manual: Outcrop Barrier Design for Above Drainage Coal Mines,
U.S. Department of the Interior, Ofce of Surface Mining, Washington, DC.
Kravits, S.J. and J.J. Schwoebel, 1994, Directional Drilling Techniques for Exploration in Advance
of Mining, World Mining Equipment, Vol. 18, No. 1, pp. 24-28.
Krinitzsky, E.L., 2002, How to Obtain Earthquake Ground Motions for Engineering Design,
Engineering Geology, Vol. 65, No. 1, pp.1-16.
* Krinitzsky, E. L., 1995, State-of-the-Art for Assessing Earthquake Hazards in the United States: Selection
of Earthquake Ground Motions for Engineering, Miscellaneous Paper S-73-1, Report 29, U.S. Army
Corps of Engineers, Waterways Experiment Station, Vicksburg, MS.
Kulhawy, F.H. and P.W. Mayne, 1990, Manual on Estimating Soil Properties for Foundation Design,
Report EL-6800, Research Project No. 1493-6, Electric Power Research Institute (EPRI), Palo Alto, CA.
Kumar, S., V.K. Puri, B.M. Das, B.C. Devkota, 2001, Geotechnical Properties of Fly Ash and Lime-
Fly Ash Stabilized Coal Mine Refuse, Electronic Journal of Geotechnical Engineering, Vol. 6 (2001).
Ladd, C.C., 1991, Stability Evaluation During Staged Construction, Journal of Geotechnical Engi-
neering, ASCE, Vol. 117, No. 4, pp. 540-615.
Ladd, C.C. and R. Foot, 1994, A New Design Procedure for Stability of Sof Clay, Journal of Geo-
technical Engineering, ASCE, Vol. 100, No. 7, pp. 763-786.
Lambe, T.W., 1967, The Stress Path Method Journal of the Soil Mechanics and Foundation Division,
ASCE, Vol. 93, No. 6, pp. 309-331.
Lambe, T.W., 1951, Soil Testing for Engineers, John Wiley & Sons, New York, NY.
Lambe, T.W. and A.M. Marr, 1979, Stress Path Method: Second Edition, Journal of Geotechnical
Engineering, ASCE, Vol. 105, No. 6, pp. 727-738.
* LaVassar, J.M., 2000, Two Decades of Dam Safety Ofce Experience in Applying a Risk Based
Approach to Mine Impoundments, Proceedings, Tailings Dams 2000, Las Vegas, Association of
State Dam Safety Ofcials (ASDSO), Lexington, KY.
Legget, R.F., 1962, Geology and Engineering, McGraw-Hill, New York, NY.
Leonards, G.A., ed., 1962, Foundation Engineering, McGraw-Hill, New York, NY.
Leonards, G.A. and J. Narain, 1963, Flexibility of Clay and Cracking of Dams, Journal of the Soil
Mechanics and Foundations Division, ASCE, Vol. 89, No. SM2, pp. 47-98.
Leps, T. M., 1973, Flow Through Rockfll, Embankment-Dam Engineering: Casagrande Volume, R.C.
Hirschfeld and S.J. Poulos, eds., John Wiley & Sons, New York, NY, pp. 87-107.
Linsley, R.K. and J.B. Franzini, 1972, Water Resources Engineering, McGraw-Hill, New York, NY.
Linsley, R.K., M.A. Kohler and J.L. Paulhus, 1982, Hydrology for Engineers, 3rd Ed., McGraw-Hill,
New York, NY.
< PREVIOUS VIEW
R17
Engineering and Design Manual Coal Refuse Disposal Facilities
Lo, K-Y and Y.N. Lee, 1990, Hydraulic Fracture in Earth- and Rock-fll Dams, Canadian Geotechni-
cal Journal, Vol. 27, No. 4, pp. 496-506.
Londe, P., 1970, La Fissuration des Noyaux, Proceedings, 10th International Conference on Large
Dams, Vol. VI, pp. 384-387.
Love, E. R., W.H. Hammack, J. Sams, G. Veloski and T. Ackman, 2005, Case History Using Airborne
Thermal Infrared Imagery and Helicopter EM Conductivity to Locate Mine Pools and Discharges in
the Ketle Creek Watershed, North-Central Pennsylvania, Geophysics, Vol. 70, No. 6, pp. B73B81.
Lowe, J. and L. Karafath, 1960, Efect of Anisotropic Consolidation on the Undrained Strength of
Compacted Clays, Proceedings of the ASCE Research Conference on Shear Strength on Cohesive Soils,
New York, American Society of Civil Engineers, Reston, VA, pp. 837-858.
Luetich, S.M., J.P. Giroud and R.C. Bachus, 1992, Geotextile Filter Design Guide, Journal of Geo-
textiles and Geomembranes, Vol. 11, Nos. 4-6, pp. 355-370.
Lunne, T., P.K. Robertson and J.J.M. Powell, 1997, Cone Penetration Testing in Geotechnical Practice,
Spon Press, London, UK.
Luo, Y., S.S. Peng and H.J. Chen, 1996, Identifcation of Factors Afecting Horizontal Displace-
ment in Subsidence Process, Proceedings, 15th International Conference on Ground Control in Mining,
Colorado School of Mines, Golden, CO, pp. 155-165.
Mackey, R.E. and G.R. Koerner, 1999, Biological Clogging of Geotextile Filters A Five Year
Study, Proceedings, Geosynthetics 99, Boston, Geosynthetics Research Institute, Folsom, PA, Vol. 2,
pp. 783-798.
* Magnuson, M.O. and E.C. Baker, 1974, State-of-the-Art in Extinguishing Refuse Pile Fires, Pro-
ceedings, First Symposium on Mine and Preparation Plant Refuse Disposal, Louisville, KY, National
Coal Association, Washington, DC, pp. 165-182.
Mainali, G., 2006, Monitoring of Tailings Dams with Geophysical Methods, Licentiate Thesis, Lulea
University of Technology, Department of Civil and Environmental Engineering, Division of
Applied Geophysics, Lulea, Sweden.
Makdisi, F.I. and H.B. Seed, 1978, Simplifed Procedure for Estimating Dam and Embankment
Earthquake-Induced Deformations, Journal of the Geotechnical Engineering Division, ASCE, Vol. 104,
No. GT7, pp. 849-867.
* Makdisi, F.I. and H.B. Seed, 1977, A Simplifed Procedure for Computing Maximum Crest Accelera-
tion and Natural Period for Embankments, Report No. UCB/EERC-77/19, Earthquake Engineering
Research Center, University of California, Berkeley, CA.
Malvick, E.J., B.L. Kuter, R.W. Boulanger and R. Kulasingam, 2006, Shear Localization Due to
Liquefaction-Induced Void-Redistribution in a Layered Infnite Slope, Journal of Geotechnical and
Geoenvironmental Engineering, ASCE, Vol. 132, No. 10, pp. 1293-1303.
Maneval, D.R., 1969, Recent Advances in Extinguishment of Burning Coal Refuse Banks for Air
Pollution Reduction, Proceedings, Symposium on Pollution Control in Fuel Combustion, Processing and
Mining, Minneapolis, MN, pp. 27-41.
Marcuson, W.F., III, M.E. Hynes and A.G. Franklin, 1992, Seismic Stability and Permanent Defor-
mation Analyses: The Last Twenty-Five Years, Proceedings, Stability Performance of Slopes and
Embankments II, R.B. Seed and R.W. Boulanger, eds., University of California, Berkeley, American
Society of Civil Engineers (ASCE), Reston, VA, pp. 552-592.
* Mark, C., 2006, The Evolution of Intelligent Coal Pillar Design: 1981-2006, Proceedings, 25th
International Conference on Ground Control in Mining, University of West Virginia, Morgantown,
WV, pp. 325-334.
< PREVIOUS VIEW
R18
References
* Marks, B.D. and B.A. Tschantz, 2002, A Technical Manual on the Efects of Tree and Woody Vegeta-
tion Root Penetrations on the Safety of Earthen Dams, ASDSO/FEMA Workshop, compiled by Marks
Enterprises of NC, PLLC, Arden, NC.
* Marshall Miller & Associates (MM&A), 2006, An Analysis of In-Seam Seismic Refection Techniques to
Identify and Locate Abandoned Underground Mine Voids in Advance of Mining, Geophysical Void Detec-
tion Demonstration Project, prepared for MSHA by MM&A and Virginia Tech, Bluefeld, WV.
Martin, J.F., 1974, Quality of Efuents from Coal Refuse Piles, Proceedings, First Symposium on
Mine and Preparation Plant Refuse Disposal, Louisville, KY, National Coal Association, Washington,
DC, pp. 26-37.
* Martin, T.E. and M. P. Davies, 2000, Development and Review of Surveillance Programs for Tail-
ings Dams, Proceedings, Tailings Dams 2000, Las Vegas, Association of State Dam Safety Ofcials
(ASDSO), Lexington, KY, pp. 367-380.
Mayne, P.W. and F.H. Kulhawy, 1982, K
o
-OCR Relationships in Soil, Journal of Geotechnical Engi-
neering, ASCE, Vol. 108, No. GT6, pp. 851-872.
* Mayne, P.W., B.R. Christopher and J. DeJong, 2002, Subsurface Investigations (Geotechnical Site Char-
acterization), FHWA-NHI-01-031, U.S. Department of Transportation, Federal Highway Adminis-
tration (FHWA), National Highway Institute, Washington, DC.
Maynord, S.T., 1978, Practical Riprap Design, Miscellaneous Paper H-78-7, U.S. Department of the
Army, Corps of Engineers, Waterways Experiment Station, Vicksburg, MS.
* McCook, D.K., 2006, Pipes in Drain Systems Are They a Necessary Evil? 2006 Annual Confer-
ence Proceedings, Boston, Association of State Dam Safety Ofcials (ASDSO), Lexington, KY.
* McCook, D.K., 2002, Things Your Mother Never Told You about Earth Dam Design, 2002 Annual
Conference Proceedings, Tampa, Association of State Dam Safety Ofcials (ASDSO), Lexington, KY.
* McCook, D.K., and J.R. Talbot, 2008, A Comprehensive Sand and Gravel Filter Design Approach
with Examples, Flow Charts, and Step by Step Process, 2008 Annual Conference Proceedings, Indian
Wells, Association of State Dam Safety Ofcials (ASDSO), Lexington, KY.
McCutcheon, H.P., 1983, Liquefaction and Cyclic Mobility of Coal Refuse Material, M.S. Thesis, Carn-
egie Mellon University, Pitsburgh, PA.
McGrath, T.J. and R.J. Hoopes, 1998, Bedding Factors and E Values for Buried Pipe Installations
Backflled with Air-Modifed CLSM, The Design and Application of Controlled Low-Strength Materials
(Flowable Fill), ASTM STP 1331-EB, A.K. Howard and J.L. Hitch, eds., American Society for Testing
and Materials (ASTM), West Conshohocken, PA.
* McGuire, R.K., 1978, FRISK: Computer Program for Seismic Risk Analysis Using Faults as Earthquake
Sources, Open File Report 78-1007, U.S. Department of the Interior, U.S. Geological Survey, Denver, CO.
* McGuire, R.K., 1976, FORTRAN Computer Program for Seismic Risk Analysis, Open-File Report 76-
67, U.S. Department of the Interior, U.S. Geological Survey, Denver, CO.
McGuire, R.K., 1974, Seismic Structural Response Risk Analysis Incorporating Peak Response Regressions
on Earthquake Magnitude and Distance, Research Report R74-51, Massachusets Institute of Tech-
nology, Department of Civil Engineering, Cambridge, MA.
McKenna, G., 2006, Rules of Thumb for Geotechnical Instrumentation Costs, Geotechnical News,
BiTech Publishers, Richmond, BC, Canada, Vol. 24, No. 2, pp. 46-47.
McLaren, R.J. and A.M. DiGioia, Jr., 1987, The Typical Engineering Properties of Fly Ash, Pro-
ceedings, Geotechnical Practice for Waste Disposal 87, R.D. Woods, ed., New York, American Society
of Civil Engineers (ASCE), Reston, VA, pp. 683697.
< PREVIOUS VIEW
R19
Engineering and Design Manual Coal Refuse Disposal Facilities
* McNay, L.M., 1971, Coal Refuse Fires, an Environmental Hazard, Report No. BM-IC-8515, U.S.
Department of the Interior, Bureau of Mines, Washington, DC.
McNeill, J.D., 1990, Use of Electromagnetic Methods for Groundwater Studies, Investigations in
Geophysics No. 5: Geotechnical and Environmental Geophysics, S.H. Ward, ed., Society of Exploration
Geophysicists, Tulsa, OK, Vol. I, pp. 191218.
* Meek, F.A., 1994, Evaluation of Acid Prevention Techniques Used in Surface Mining, Proceed-
ings, International Land Reclamation and Mine Drainage Conference, SP 06B-94, U.S. Department of the
Interior, Bureau of Mines, Pitsburgh, PA, pp. 41-48.
Mesri, G. and M.E.M. Abdel-Ghafar, 1993, Cohesion Intercept in Efective Stress Stability Analy-
sis, Journal of Geotechnical Engineering, Vol. 119, No. 8, pp. 1229-1249.
Michael, P. and M. Superfesky, 2007, Excess Spoil Minimization and Fill Stability, Proceedings,
2007 Annual Meeting, Society for Mining, Metallurgy and Exploration (SME), Denver, CO.
*Michael, P., R. Murguia and L. Kosareo, 2005, The Flowability of Impounded Coal Refuse: A Review of
Recent Work and Current Ideas in the Engineering Profession, OSM Final Report, U.S. Department of
the Interior, Ofce of Surface Mining, Washington, DC.
Michalek, S.J., G.H. Gardner and K.K. Wu, 1996, Accidental Releases of Slurry and Water from Coal
Impoundments through Abandoned Underground Mines, U.S. Department of Labor, MSHA, Safety and
Health Technology Center, Pitsburgh, PA.
Mihok, E.A. and C.E. Chamberlain, 1968, Factors in Neutralizing Acid Mine Water with Lime-
stone, Proceedings, 2nd Symposium on Coal Mine Drainage Research, Mellon Institute, Pitsburgh, PA,
pp. 165-173.
Mikkelsen, P.E., 2002, Cement-Bentonite Grout Backfll for Borehole Instruments, Geotechnical
News, Vol. 20, No. 4, pp. 38-42.
* Mine Safety and Health Administration (MSHA), 2008, Final Reports on Geophysical Void Detection
Demonstration Projects, U.S. Department of Labor, MSHA, Arlington, VA.
* Mine Safety and Health Administration (MSHA), 2007, MSHA Coal Mine Impoundment Inspection and
Plan Review Handbook, Handbook No. PH07-01, U.S. Department of Labor, MSHA, Arlington, VA.
* Mine Safety and Health Administration (MSHA), 2003, Report to Congress: Responses to Recom-
mendations in the National Research Councils Report, Coal Waste Impoundments: Risks, Responses and
Alternatives, U.S. Department of Labor, MSHA, Arlington, VA.
* Mine Safety and Health Administration (MSHA), 1999, Haul Road Inspection Handbook, PH99-I-4,
U.S. Department of Labor, MSHA, Arlington, VA.
Mine Safety and Health Administration (MSHA), 1993, Coal Impoundment Inspection Procedures
Handbook, PH89-V-4, U.S. Department of Labor, MSHA, Arlington, VA.
Mine Safety and Health Administration (MSHA), 1983, Design Guidelines for Coal Refuse Piles and
Water, Sediment, or Slurry Impoundments, and Impounding Structures, Amendment to Informational
Report 1109, Technical Support Center, Denver, CO and Pitsburgh, PA.
* Mining Enforcement and Safety Administration (MESA), 1975, Engineering and Design Manual,
Coal Refuse Disposal Facilities, U.S. Department of the Interior, MESA, Washington, DC.
Mishra, G.C. and A.K. Singh, 2005, Seepage through a Levee, International Journal of Geomechan-
ics, ASCE, Vol. 5, No. 1, pp. 74-79.
Mitchell, R.J., 1983, Earth Structures Engineering, Allen and Unwin, Boston MA.
Molinda, G.M., C. Mark and D. Debasis, 2001, Using the Coal Mine Roof Rating (CMRR) to Assess
Roof Stability in U.S. Coal Mines, Journal of Mines, Metals & Fuels, Vol. 49, No. 8-9, pp. 314-321.
< PREVIOUS VIEW
R20
References
Morgenstern, N.R. and V.E. Price, 1965, The Analysis of the Stability of General Slip Surfaces,
Gotechnique, Vol. 15, No. 1, pp. 79-93.
Morris, P.H. and D.J. Williams, 2000, A Revision of Blights Model of Field Vane Testing, Canadian
Geotechnical Journal, Vol. 37, No. 5, 1089-1098.
Moser, A.P., 2001, Buried Pipe Design, 2nd Ed., McGraw-Hill, New York, NY.
* Myers, J.W., J.J. Pfeifer, E.M. Murphy and F.E. Grifth, 1966, Ignition and Control of Burning of Coal
Mine Refuse, Report No. BM-RI-6758, U.S. Department of the Interior, Bureau of Mines, Washington, DC.
Naesgaard, E., P.M. Byrne, M. Seid-Karbasi and S.S. Park, 2005, Modeling Flow Liquefaction, Its
Mitigation, and Comparison with Centrifuge Tests, Proceedings, Geotechnical Earthquake Engineer-
ing Satellite Conference, Osaka, Japan.
National Academy of Sciences, 1975, Underground Disposal of Coal Mine Wastes, Washington, DC.
National Coal Board (NCB), 1975, Subsidence Engineers Handbook, Mining Department of the NCB,
London, UK.
National Coal Board (NCB), 1972, Review of Research on Properties of Spoil Tip Materials, NCB,
London, UK.
National Coal Board (NCB), 1970, Spoil Heaps and Lagoons, Technical Handbook, NCB, London, UK.
National Research Council (NRC), 2002, Coal Waste Impoundments: Risks, Responses, and Alternatives,
Commitee on Coal Waste Impoundments, Commitee on Earth Resources, Board on Earth Sciences
and Resources, Division on Earth and Life Studies, National Academy Press, Washington, DC.
National Stone, Sand & Gravel Association (NSSGA formerly National Stone Association), 1989,
Quarried Stone for Erosion and Sediment Control, Alexandria, VA.
National Weather Service (NWS), 2006a, NOAA Atlas 14, Precipitation Atlas of the United States,
Volume 1, The Semi-Arid Southwest, U.S. Department of Commerce, National Oceanic and Atmo-
spheric Administration, NWS (available on-line only).
National Weather Service (NWS), 2006b, NOAA Atlas 14, Precipitation Atlas of the United States,
Volume 2, The Ohio River Basin and Surrounding States, U.S. Department of Commerce, National
Oceanic and Atmospheric Administration, NWS (available on-line only).
* National Weather Service (NWS), 1999, Hydrometeorolgical Report No. 59, Probable Maximum Precipi-
tation for California, U.S. Department of Commerce, National Oceanic and Atmospheric Administra-
tion, NWS, Washington, DC.
* National Weather Service (NWS), 1998, Hydrometeorolgical Report No. 58, Probable Maximum Precipi-
tation for California Calculation Procedures, U.S. Department of Commerce, National Oceanic and
Atmospheric Administration, NWS, Washington, DC.
* National Weather Service (NWS), 1994, Hydrometeorolgical Report No. 57, Probable Maximum Precipi-
tation Pacifc Northwest States, U.S. Department of Commerce, National Oceanic and Atmospheric
Administration, NWS, Washington, DC
* National Weather Service (NWS), 1988, Hydrometeorolgical Report No. 55A, Probable Maximum Pre-
cipitation Estimates, United States between the Continental Divide and the 105th Meridian, U.S. Depart-
ment of Commerce, National Oceanic and Atmospheric Administration, NWS, Washington, DC.
* National Weather Service (NWS), 1986, Hydrometeorolgical Report No. 56, Probable Maximum Pre-
cipitation Estimates with Areal Distribution for Tennessee River Drainages Less than 3000 Square Miles
in Area, U.S. Department of Commerce, National Oceanic and Atmospheric Administration, NWS,
Washington, DC.
* National Weather Service (NWS), 1982, Hydrometeorolgical Report No. 52, Application of Probable
Maximum Precipitation Estimates, United States East of the 105th Meridian, U.S. Department of Com-
merce, National Oceanic and Atmospheric Administration, NWS, Washington, DC.
< PREVIOUS VIEW
R21
Engineering and Design Manual Coal Refuse Disposal Facilities
* National Weather Service (NWS), 1980, Hydrometerological Report No. 53, Seasonal Variation of 10-
Square-Mile Probable Maximum Precipitation Estimates, United States East of the 105th Meridian, U.S.
Department of Commerce, National Oceanic and Atmospheric Administration, NWS, Washington, DC.
* National Weather Service (NWS), 1978, Hydrometeorolgical Report No. 51, Probable Maximum Precipi-
tation Estimates, United States East of the 105th Meridian, U.S. Department of Commerce, National
Oceanic and Atmospheric Administration, NWS, Washington, DC.
* National Weather Service (NWS), 1977, Hydrometeorolgical Report No. 49, Probable Maximum Pre-
cipitation Estimates, Colorado River and Great Basin Drainages, U.S. Department of Commerce,
National Oceanic and Atmospheric Administration, NWS, Washington, DC.
National Weather Service (NWS), 1973, NOAA Atlas 2, Precipitation Frequency Atlas of the Western
United States, U.S. Department of Commerce, National Oceanic and Atmospheric Administration,
NWS, Washington, DC.
* National Weather Service (NWS), 1965, Technical Paper No. 52, Two- to 10-Day Precipitation for
Return Periods from 1 to 100 Years in Alaska, U.S. Department of Commerce, National Oceanic and
Atmospheric Administration, NWS, Washington, DC.
* National Weather Service (NWS), 1964, Technical Paper No. 49, Two- to 10-Day Precipitation for
Return Periods from 2 to 100 Years for the Contiguous United States, U.S. Department of Commerce,
National Oceanic and Atmospheric Administration, NWS, Washington, DC.
* National Weather Service (NWS), 1963, Technical Paper No. 47, Probable Maximum Precipitation and
Rainfall Frequency Data for Alaska for Areas to 400 Square Miles, Durations to 24 Hours, and Return Peri-
ods from 1 to 100 Years, U.S. Department of Commerce, National Oceanic and Atmospheric Admin-
istration, NWS, Washington, DC.
* Natural Resources Conservation Service (NRCS), 2007a, National Engineering Handbook, Chapter 7:
Hydrologic Soil Groups, 210-VI-NEH, U.S. Department of Agriculture, NRCS, Washington, DC.
* Natural Resources Conservation Service (NRCS), 2007b, National Engineering Handbook, Chapter 45:
Filter Diaphragms, 210-VI-NEH, U.S. Department of Agriculture, NRCS, Washington, DC.
* Natural Resources Conservation Service (NRCS), 2007c, NRCS Fillable Form Template, U.S. Depart-
ment of Agriculture, NRCS, Washington, DC.
* Natural Resources Conservation Service, 2005a, National Engineering Handbook, Chapter 52: Structural
Design of Flexible Conduits, 210-VI-NEH, U.S. Department of Agriculture, NRCS, Washington, DC.
* Natural Resources Conservation Service (NRCS), 2005b, Earth Dams and Reservoirs, TR-60, 210-VI-
TR60, U.S. Department of Agriculture, NRCS, Washington, DC.
* Natural Resources Conservation Service (NRCS), 2004a, National Engineering Handbook, Chapter 9:
Hydrologic Soil Cover Complexes, 210-VI-NEH, U.S. Department of Agriculture, NRCS, Washington, DC.
* Natural Resources Conservation Service (NRCS), 2004b, National Engineering Handbook, Chapter 10:
Estimation of Direct Runof from Storm Rainfall, 210-VI-NEH, U.S. Department of Agriculture, NRCS,
Washington, DC.
* Natural Resources Conservation Service (NRCS), 2004c, National Engineering Handbook, Chapter 11:
Snowmelt, 210-VI-NEH, U.S. Department of Agriculture, NRCS, Washington, DC.
* Natural Resources Conservation Service (NRCS), 2002, Pond, Conservation Practice Standard No.
378, U.S. Department of Agriculture, NRCS, Washington, DC.
* Natural Resources Conservation Service (NRCS), 1994, National Engineering Handbook, Chapter
26: Gradation Design of Sand and Gravel Filters, U.S. Department of Agriculture, NRCS, Washing-
ton, DC.
< PREVIOUS VIEW
R22
References
* Natural Resources Conservation Service (NRCS), 1986, Urban Hydrology for Small Watersheds, TR-
55, 210-VI-TR55, U.S. Department of Agriculture, NRCS, Washington, DC.
* Newman, D.A., 2003, Rock Mechanics and the Analysis of Underground Mine Stability Adjacent
to Coal Refuse Impoundments, Proceedings, 22nd International Conference on Ground Control in
Mining, West Virginia University, Morgantown, WV.
Newmark, N.M., 1965, Efects of Earthquakes on Dams and Embankments, Gotechnique, Vol. 15,
No. 2, pp. 139-160.
Nicholas, A.K. and R.J. Hutnik, 1971, Ectomycorrhizal Establishment and Seedling Response on Vari-
ously Treated Deep-Mine Coal Refuse, Special Report SR-89, School of Forest Resources, Pennsylvania
State University, University Park, PA.
* Nicholls, H.R., C.F. Johnson and W.I. Duvall, 1971, Blasting Vibrations and Their Efects on Structures,
Bulletin 656, U.S. Department of the Interior, Bureau of Mines, Washington, DC.
*Nieto, A.S., 1979, Evaluation of Damage Potential to Earth Dams by Subsurface Coal Mining at
Rend Lake, Illinois, Proceedings, Geotechnics of Mining: 10th Ohio River Valley Seminar, Lexington,
KY, pp. 9-18.
Nilson, A.H., D. Darwin and C.W. Dolan, 2002, Design of Concrete Structures, 13th Ed., McGraw-
Hill, New York, NY.
Nutli, O.W., 1973, Seismic Wave Atenuation and Magnitude Relations for Eastern North Amer-
ica, Journal of Geophysical Research, Vol. 78, No. 5, pp. 876-885.
OBrien, J.J., J.A. Havers and F.W. Stubbs, eds., 1996, Standard Handbook of Heavy Construction, 3rd
Ed., McGraw-Hill, New York, NY.
OConnor, K.M. and C.H. Dowding, 1999, GeoMeasurements by Pulsing TDR Cables and Probes, CRC
Press, Boca Raton, FL.
Oglesby, C.H. and R.G. Hicks, 1982, Highway Engineering, 4th Ed., John Wiley & Sons, New York, NY.
* Ohio Department of Transportation (ODOT), 2006, Geotechnical Bulletin GB3: Rock Cut Slope and Catch-
ment Design, Division of Production Management, Ofce of Geotechnical Design, Columbus, OH.
* Olin, D.A. and J.B. Atkins, 1988, Estimating Flood Hydrographs and Volumes for Alabama Streams,
Water Resources Investigation Report 88-4041, U.S. Department of the Interior, U.S. Geological
Survey, Water Resources Division, Tuscaloosa, AL.
Olsen, R.S. and J.K. Mitchell, 1995, CPT Stress Normalization and Prediction of Soil Classifca-
tion, Proceedings of the International Symposium on Cone Penetration Testing, CPT 95, Linkping,
Sweden, Swedish Geotechnical Society, Vol. 2, pp. 257-262.
Olson, S.M. and T.D. Stark, 2003, Yield Strength Ratio and Liquefaction Analysis of Slopes and
Embankments, Journal of Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 129, No. 8,
pp. 727-737.
Olson, S.M. and T.D. Stark, 2002, Liquefed Strength Ratio from Liquefaction Flow Failure Case
Histories, Canadian Geotechnical Journal, National Research Council of Canada, Vol. 39, pp. 629-647.
Owens, H.L., 2008, personal communication.
Owens, H.L., 1987, MSHA Report of Investigation, Impoundment Principal Spillway Pipe Failure, Mont-
coal Preparation Plant (ID No. 46-03430), Peabody Coal Company, Raleigh County, WV.
Owens, H.L., 1977, MESA Memorandum on Failure of the Fine Coal Refuse Impoundment (ID No.
1211WV40009-02), Pond Fork Preparation Plant, Island Creek Coal Company, Boone County, WV.
Packer, P.E. and E.F. Aldon, 1978, Revegetation Techniques for Dry Regions, Reclamation of Dras-
tically Disturbed Land, F.W. Schaller and P. Suton, eds., American Society of Agronomy, Madison,
WI, pp. 425-450.
< PREVIOUS VIEW
R23
Engineering and Design Manual Coal Refuse Disposal Facilities
Page, A.L., R.H. Miller and D.R. Keeney, eds., 1982, Methods of Soil Analysis, Part 2: Chemical and
Microbiological Properties, Agronomy Monograph No. 9, 2nd Ed., American Society of Agronomy
and Soil Science of America, Madison, WI.
Pariseau, W.G., 2006, Design Analysis in Rock Mechanics, CRC Press, Boca Raton, FL.
Pariseau, W.G., 1975, Limit Design of Mine Pillars under Uncertainty, Proceedings, 16th U.S. Sym-
posium on Rock Mechanics, University of Minnesota, Minneapolis, MN, pp. 287-301.
Park, R. and T. Paulay, 1975, Reinforced Concrete Structures, John Wiley & Sons, New York, NY.
Paterson, R.J. and K.D. Ferguson, 1994, The Gibraltar North Project Assessing Acid Rock Drain-
age, Proceedings, 3rd International Conference on the Abatement of Acidic Drainage, Pitsburgh, PA,
U.S. Bureau of Mines Special Publication SP 06B-94, Vol. 2, pp. 12-21.
*Pearson, M.L., T. Preber and P.J. Conroy, 1981, Outcrop Barrier Design Guidelines for Appalachian Coal
Mines, Open File Report 134-81, U.S. Department of the Interior, Bureau of Mines, Washington, DC.
Peck, R.B., 1972, Observations and Instrumentation: Some Elementary Considerations, Highway
Focus, U.S. Department of Transportation, Federal Highway Administration, Vol. 4, No. 2, pp. 1-5.
Reprinted in Judgment in Geotechnical Engineering: The Professional Legacy of Ralph B. Peck, 1984, J.
Dunniclif and D.U. Deere, eds., John Wiley & Sons, New York, NY, pp. 128-130.
Pelton, F., 2000, Guidelines for Instrumentation and Measurement for Monitoring Dam Performance, 0-
7844-0531-X, ASCE Task Commitee on Instrumentation and Monitoring Dam Performance, Amer-
ican Society of Civil Engineers (ASCE), Reston, VA.
Peng, S.S. and H.S. Chiang, 1984, Longwall Mining, John Wiley & Sons, New York, NY.
Peng, S.S. and Y. Luo, 1993, A New Method for Designing Support Area to Protect Surface Struc-
tures over Underground Coal Mining Areas, Transactions, AIME-SME, Vol. 294, pp. 1927-1932.
* Pennsylvania Department of Environmental Protection (PADEP), 2000, Erosion and Sediment Pol-
lution Control Program Manual, Document No. 363-2134-008, PADEP, Ofce of Water Management,
Bureau of Watershed Management, Division of Waterways, Wetlands and Erosion Control, Har-
risburg, PA.
Pennsylvania Department of Transportation (PennDOT), 2007, Specifcations Publication 408/2007,
Commonwealth of Pennsylvania, PennDOT, Harrisburg, PA.
* Pennsylvania State University (PSU), 2006, An In-Seam Seismic (ISS) Method Based Mine Void Detection
Technique, Final Report for Phase I, prepared for MSHA by Maochen Ge, PSU, University Park, PA.
Performance Pipe, 2003, The Performance Pipe Engineering Manual, Performance Pipe Division of
Chevron Phillips Chemical Company, LP, Plano, TX.
* Perin, R.J., 2000, Foundation Treatment of Stress Relief Fractures Hillseams in a Tailings
Embankment Cutof Trench, Proceedings, Tailings Dams 2000, Las Vegas, ASDSO/USCOLD, pp.
197-209.
Perlea, V.G., M.E. Rothford and D.C. Serafni, 2009, Use of Seismic Deformation Analysis in
Remediation Design, Proceedings, 29th Annual USSD Conference, Nashville, United States Society
on Dams, Denver, CO.
Phelps, L.B., W.B. Wells and L.W. Saperstein, 1981, An Analysis of Surface Coal Mine Spoil Bulk
Density, Society of Mining Engineering Annual Meeting, SME Preprint No. 81-53, Chicago, IL.
Poulos, S.J., 1988, Chapter 9 Liquefaction and Related Phenomena, Advanced Dam Engineering
for Design, Construction, and Rehabilitation, R.B. Jansen, ed., Van Nostrand Reinhold, New York, NY.
Poulos, S.J., 1981, The Steady State of Deformation, Journal of the Geotechnical Engineering Divi-
sion, ASCE, Vol. 107, No. GT5, pp. 553-562.
< PREVIOUS VIEW
R24
References
Poulos, H.G. and E.H. Davis, 1974, Elastic Solutions for Soil and Rock Mechanics, John Wiley & Sons,
New York, NY.
Poulos, S.J., G. Castro and J.W. France, 1985, Liquefaction Evaluation Procedure, Journal of the
Geotechnical Engineering Division, ASCE, Vol. 111, No. GT6, pp. 772-792.
Prat, H.K., R.C. McMordie and R.M. Dundas, 1972, Foundations and Abutment - Bennet and
Mica Dams, Journal of the Soil Mechanics and Foundations Division, ASCE, Vol. 98, SM10, pp. 1053-1072.
Price, W.A., K.A. Morin and N.M. Hut, 1997, Guidelines for the Prediction of Acid Rock Drainage
and Metal Leaching for Mines in British Columbia: Part II, Recommended Procedures for Static and
Kinetic Tests, Proceedings, 4th International Conference on Acid Rock Drainage, Vancouver, BC, Canada,
American Society of Surface Mining and Reclamation, Vol. I, pp. 15-30.
Qiu, Y.J., and D.C. Sego, 2001, Laboratory Properties of Mine Tailings, Canadian Geotechnical Jour-
nal, Vol. 38, No. 1, pp. 183-190.
Quaranta, J.D., L.D. Banta, and J.A. Altobello, 2008, Remote Monitoring of a High Hazard Coal
Waste Impoundment in Mountainous Terrain Case Study, Proceedings, Tailings and Mine Waste 08,
Vail, CO, pp. 125-136.
Reddi, L.N., 2003, Seepage in Soils Principles and Applications, John Wiley & Sons, Hoboken, NJ.
Rellensmann, O. and E. Wagner, 1957, The Efect on Railways of the Ground Movements Due to
Mining, Proceedings, First European Congress on Ground Movement, University of Leeds, Leeds, UK,
pp. 74-82.
Rich, D.H. and K.R. Hutchinson, 1990, Neutralization and Stabilization of Combined Refuse
Using Lime Kiln Dust at High Power Mountain, Proceedings, 1990 Mining and Reclamation Confer-
ence and Exhibition, Charleston, WV, Vol. I, pp. 55-60.
Richter, C., 1958, Elementary Seismology, W.H. Freeman and Co., San Francisco, CA.
Rico, M., G. Benito and A. Diez-Herrero, 2008, Floods from Tailings Dam Failures, Journal of Haz-
ardous Materials, Vol. 152, No. 2, pp. 846-852.
* Rixner, J.J., S.R. Kraemer and A.D. Smith, 1986, Prefabricated Vertical Drains, Vol. I, Engineering
Guidelines, Report No. FHWA/RD-86/168, Ofce of Engineering and Highway Operations Research
and Development, Federal Highway Administration, McLean, VA.
* Robbins, C.H., 1986, Techniques for Simulating Flood Hydrographs and Estimating Flood Volumes for
Ungaged Basins in Tennessee, Water Resources Investigation Report 86-4192, U.S. Department of the
Interior, U.S. Geological Survey, Water Resources Division, Nashville, TN.
Robertson, P.K., 2008, discussion of paper, Liquefaction Potential of Silts from CPTu, Canadian
Geotechnical Journal, Vol. 45, pp. 140-141.
Robertson, P.K., 2004, Evaluating Soil Liquefaction and Post-Earthquake Deformations Using the
CPT, Proceedings of ISC-2 on Geotechnical and Geophysical Site Characterization, Porto, Portugal, V.
Fonseca and P.W. Mayne, eds., Millpress, Roterdam, Vol. 1, pp. 233-252.
Robertson, P.K., 1994, Suggested Terminology for Liquefaction, Proceedings of the 47th Canadian
Geotechnical Conference, Halifax, NS, Canadian Geotechnical Society, Alliston, ON, Canada, pp. 277-286.
Robertson, P.K. and R.G. Campanella, 1983, Interpretation of Cone Penetration Tests: Part I -
Sands; Part II Clays, Canadian Geotechnical Journal, Vol. 20, No. 4, pp. 719-745.
Robertson, P.K. and C.E. Wride, 1998, Evaluating Cyclic Liquefaction Potential Using the Cone
Penetration Test, Canadian Geotechnical Journal, Vol. 35, No. 3, pp. 442-459.
Rupert, G.B. and G.B. Clark, 1977, Criteria for the Proximity of Surface Blasting to Underground
Coal Mines, Proceedings, 18th U.S. Symposium on Rock Mechanics, Colorado School of Mines, Key-
stone, CO, pp. 3C3-1 to 3C3-10.
< PREVIOUS VIEW
R25
Engineering and Design Manual Coal Refuse Disposal Facilities
* Rusnak, J. and C. Mark, 2000, Using the Point Load Test to Determine the Uniaxial Compres-
sive Strength of Coal Measure Rock, Proceedings, 19th International Conference on Ground Control in
Mining, S.S. Peng and C. Mark, eds., West Virginia University, Morgantown, WV, pp. 362-371.
Rutledge, P.C. and J.P. Gould, 1973, Movements of Articulated Conduits under Earth Dams on
Compressible Foundations, Embankment-Dam Engineering: Casagrande Volume, R.C. Hirschfeld and
S.J. Poulos, eds., John Wiley & Sons, New York, NY, pp. 209-237.
* Sabatini, P.J., R.C. Bachus, P.W. Mayne, J.A. Schneider and T.E. Zetler, 2002, Geotechnical Engineer-
ing Circular No. 5: Evaluation of Soil and Rock Properties, DTFH61-94-C-00099, U.S. Department of
Transportation, Federal Highway Administration (FHWA), Ofce of Bridge Technology, Washing-
ton, DC.
Saikia, C.K. and P.G. Somerville, 1997, Simulated Hard-Rock Motions in Saint Louis, Missouri,
from Large New Madrid Earthquakes (M
w
6.5), Bulletin of the Seismological Society of America,
Vol. 87, No. 1, pp. 123-139.
* Sames, G.P. and N.N. Moebs, 1992, Roof Control of Stress-Relief Jointing near Outcrops in Central
Appalachian Drif Mines, Information Circular 9313, U.S. Department of the Interior, Bureau of
Mines, Pitsburgh, PA.
* Sames, G.P. and N.N. Moebs, 1989, Hillseam Geology and Roof Instability Near Outcrop in East-
ern Kentucky Drif Mines, Report of Investigations 9267, U.S. Department of the Interior, Bureau
of Mines, Pitsburgh, PA.
Sarma, S.K., 1973, Stability Analysis of Embankments and Slopes, Gotechnique, Vol. 23, No. 3,
pp. 423-433.
* Savy, J.B., W. Foxall, N.A. Abrahamson and D.L. Bernreuter, 2002, Guidance for Performing Proba-
bilistic Seismic Hazard Analysis for a Nuclear Plant Site: Example Application to the Southeastern United
States, NUREG/CR-6607 (UCRL-ID-133494), Ofce of Nuclear Regulatory Research, U.S. Nuclear
Regulatory Commission (USNRC), Washington, DC.
Saxena, S.K., D.E. Lourie and J.S. Rao, 1984, Compaction Criteria for Eastern Coal Waste Embank-
ments, Journal of the Geotechnical Engineering Division, ASCE, Vol. 110, No. GT2, pp. 262-284.
* Schaefer, M.G., 1992, Dam Safety Guidelines - Technical Note 1: Dam Break Inundation Analysis and
Downstream Hazard Classifcation, Publication 92-55E, Washington State Department of Ecology,
Olympia, WA.
* Schnabel, P.B., J. Lysmer and H.B. Seed, 1972, SHAKE: A Computer Program for Earthquake Response
Analysis of Horizontally Layered Sites, Report No. EERC 72-12, University of California, Berkeley, CA.
* Schubert, C.E., W.J. Johnson and B.W. Hassinger, 1982, Improved Resource Characterization for Coal
Mining(Geophysics and Drilling Technology), Contract No. USDOE AC22-80PC30114, U.S. Depart-
ment of Energy, Ofce of Coal Mining, Washington, DC.
Schwartz, D.P. and K.J. Coppersmith, 1986, Seismic Hazards: New Trends in Analysis Using Geologic
Data, Studies in Geophysics-Active Tectonics, National Academy Press, Washington, DC, pp. 215-230.
Scovazzo, V.A., 2008, personal communication regarding experience with angle of draw.
* Scovazzo, V.A., 1995, Comparison of Empirical, Analytical, and Numerical Methods of Pillar
Design and the Development of a Site Specifc Pillar Formula, Proceedings, 14th International Con-
ference on Ground Control in Mining, West Virginia University, Morgantown, WV, pp. 116-123.
Seed, H.B., 1987, Design Problems in Soil Liquefaction, Journal of Geotechnical Engineering, ASCE,
Vol. 113, No. 8, pp. 827-845.
Seed, H.B., 1979, Soil Liquefaction and Cyclic Mobility Evaluation for Level Ground During
Earthquakes, Journal of the Geotechnical Engineering Division, ASCE, Vol. 105, No. GT2, pp. 201-255.
< PREVIOUS VIEW
R26
References
Seed, R.B. and L.F. Harder, Jr., 1990, SPT-Based Analysis of Cyclic Pore Pressure Generation and
Undrained Residual Strength, Proceedings of the H. Bolton Seed Memorial Symposium, Vol. 2, Univer-
sity of California, Berkeley, CA, BiTech Publishers, Richmond, BC, Canada, pp. 351-376.
Seed, H.B. and I.M. Idriss, 1982, Ground Motions and Soil Liquefaction During Earthquakes,
Earthquake Engineering Research Institute, Berkeley, CA.
Seed, H.B. and I.M. Idriss, 1971, Simplifed Procedure for Evaluating Soil Liquefaction Potential,
Journal of the Geotechnical Engineering Division, ASCE, Vol. 97, No. 9, pp. 1249-1273.
* Seed, R.B., K.O. Cetin, R.E.S. Moss, A.M. Kammerer, J. Wu, J.M. Pestina, M.F. Riemer, R.B. Sancio,
J.D. Bray, R.E. Kayen and A. Faris, 2003, Recent Advances in Soil Liquefaction Engineering: A Unifed
and Consistent Framework, Report No. EERC 2003-06, 26th Annual ASCE Los Angeles Geotechnical
Spring Seminar, Long Beach CA.
Serafm, J.L. and J.P. Pereira, 1983, Considerations on the Geomechanical Classifcation of Bien-
iawski, Proceedings, International Symposium on Engineering Geology and Underground Construction,
Lisbon, Portugal, Theme II, Vol. 1, pp. II.33-II.42.
Shackelford, R., 2000, Underground Surveying A Primer in Coal Mine Surveying and Map-
ping, Point of Beginning Electronic Magazine, Sept. 29, 2000 issue.
Sherard, J.L., 1973, Embankment Dam Cracking, Embankment-Dam Engineering: Casagrande
Volume, R.C. Hirschfeld and S.J. Poulos, eds., John Wiley & Sons, New York, NY, pp. 272-353.
Sherard, J.L. and L.P. Dunnigan, 1985, Filters and Leakage Control in Embankment Dams, Pro-
ceedings, Symposium on Seepage and Leakage from Dams and Impoundments, Denver, American Society
of Civil Engineers (ASCE), Reston, VA, pp. 1-31.
Sherard, J.L., L.P. Dunnigan, R.S. Decker and E.F. Steele, 1976, Pinhole Test for Identifying Disper-
sive Clays, Journal of the Geotechnical Engineering Division, ASCE, Vol. 102, No. GT1, pp. 69-85.
Sherard, J.L., R.J. Woodward, S.F. Gizienski, and W.A. Clevenger, 1963, Earth and Earth-Rock Dams:
Engineering Problems of Design and Construction, John Wiley & Sons New York, NY.
Shoup, D., 1992, Protecting Geotechnical Sensors and Cable from Lightning Damage, Sensors in
the Real World, Slope Indicator Company.
Singh, M.M., 1992, Mine Subsidence, SME Mining Engineering Handbook, 2nd Ed., H. Hartman,
ed., Society for Mining, Metallurgy, & Exploration (SME), Litleton, CO, pp. 938-971.
Singh, M.M., 1978, Experience with Subsidence Due to Mining, Proceedings, International Con-
ference on Evaluation and Prediction of Subsidence, New York, American Society of Civil Engineers
(ASCE), Reston, VA, pp. 92-112.
* Singh, M.M. and F.S. Kendorski, 1981, Strata Disturbance Prediction for Mining Beneath Surface
Water and Waste Impoundments, Proceedings, 1st Conference on Ground Control in Mining, West
Virginia University, Morgantown, WV, pp. 76-89.
* Sirles, P.C., 2006, Use of Geophysics for Transportation Projects: A Synthesis of Highway Practice,
NCHRP Synthesis 357, Transportation Research Board, National Cooperative Highway Research
Program, Washington, DC.
*Siskind, D.E., M.S. Stagg, J.W. Kopp and C.H. Dowding, 1980, Structure Response and Damage Pro-
duced by Ground Vibration from Surface Mine Blasting, Report of Investigations 8507, U.S. Depart-
ment of the Interior, Bureau of Mines, Pitsburgh, PA.
* Skelly and Loy, 1977, Guidelines for Mining near Surface Waters, Open File Report 29-77, U.S.
Department of the Interior, Bureau of Mines, Washington, DC.
* Skousen, J.G., A.W. Rose, G. Geidel, R. Foreman, R. Evans, W.W. Hellier and members of the
Avoidance and Remediation Working Group of the Acid Drainage Technology Initiative, 1998,
< PREVIOUS VIEW
R27
Engineering and Design Manual Coal Refuse Disposal Facilities
Handbook of Technologies for Avoidance and Remediation of Acid Mine Drainage, National Mine Land
Reclamation Center, West Virginia University, Morgantown, WV.
Skousen, J.G., J.C. Sencindiver and R.M. Smith, 1987, A Review of Procedures for Surface Mining and
Reclamation in Areas with Acid-Producing Materials, West Virginia University Energy and Water
Research Center Publication EWRC 871, Morgantown, WV.
Slemmons, D.B. and C.M. Depolo, 1986, Evaluation of Active Faulting and Associated Hazards,
Studies in Geophysics-Active Tectonics, National Academy Press, Washington, DC, pp. 45-62.
Smith, R.M., A.A. Sobek, T. Arkle, Jr., J.C. Sencindiver and J.R. Freeman, 1976, Extensive Overburden
Potenials for Soil and Water Quality, EPA 600/1-76-184, U.S. Department of the Environment, U.S.
Environmental Protection Agency, Washington, DC.
Smith, R.M., W.E. Grube, T.Arkle, Jr. and A. Sobek, 1974, Mine Spoil Potentials for Soil and Water
Quality, EPA 670/2-74-070, U.S. Department of the Environment, U.S. Environmental Protection
Agency, Washington, DC.
* Smith, S.A. and D.M. Hosler, 2006, Current Research in Dam Drain Clogging and Its Prevention,
Proceedings, 2006 ASDSO Annual Conference, Boston, Association of State Dam Safety Ofcials, Lex-
ington, KY.
Snow, R.E., J.L. Olson and W. Schultz, 2000, Slurry Impoundment Internal Drain Design and Con-
struction, Proceedings of the 7th Conference on Tailings and Mine Waste 00, Fort Collins, CO, pp. 65-73.
Sobel, P., 1994, Revised Livermore Seismic Hazard Estimates for Sixty-Nine Nuclear Power Plant Sites
East of the Rocky Mountains, NUREG 1488, Division of Engineering, Ofce of Nuclear Reactor Reg-
ulation, U.S. Nuclear Regulatory Commission (USNRC), Washington, DC.
Somerville, P.G., N.F. Collins, N.A. Abrahamson, R.W. Graves and C.K. Saikia, 2001, Ground Motion
Atenuation Relations for the Central and Eastern United States, Final Report to U.S. Geological Survey,
URS Group, Inc., Pasadena, CA.
Song, S-H., Y. Song and B-D. Kwon, 2005, Application of Hydrogeological and Geophysical Meth-
ods to Delineate Leakage Pathways in an Earth Fill Dam, Exploration Geophysics, Vol. 36, No. 1,
pp. 92-96.
Sowers, G.F. and G.B. Sowers, 1970, Introductory Soil Mechanics and Foundations, 4th Ed., Macmillan
Publishing Company, New York, NY.
Spencer, E., 1967, A Method of Analysis of the Stability of Embankments Assuming Parallel Inter-
slice Forces, Gotechnique, Vol. 17, No. 1, pp. 11-26.
Stewart, B.M. and G. Robinson, 1994, MSHA Accident Investigation Report, Non-Injury Inundation
Report, Martin County Coal Corporation, 1-C Mine (ID No. 15-03752), Martin County, KY.
* Stewart, B.M. and L.A. Atkins, 1983, Physical Property Data on Coarse Anthracite Waste, Report of
Investigations 8782, U.S. Department of the Interior, Bureau of Mines, Pitsburgh, PA.
* Stewart, B.M. and L.A. Atkins, 1982, Engineering Properties of Combined Coarse and Fine Coal Wastes,
Report of Investigations 8623, U.S. Department of the Interior, Bureau of Mines, Pitsburgh, PA.
* Stewart, B.R., W.L. Daniels, L.W. Zelazny and M.L. Jackson, 2001, Evaluation of Leachates from
Coal Refuse Blended with Fly Ash at Diferent Rates, Journal of Environmental Quality, Vol. 30,
No. 5, pp. 1382-1391.
* Stewart, B.R., W.L. Daniels, and M.L. Jackson, 1997, Evaluation of Leachate Quality from the
Co-Disposal of Coal Fly Ash and Coal Refuse, Journal of Environmental Quality, Vol. 26, No. 5,
pp. 1412-1424.
Szymanski, M.B., 1999, Evaluation of Safety of Tailings Dams, BiTech Publishers, Richmond, BC, Canada.
< PREVIOUS VIEW
R28
References
* Talbot, J.R., S.J. Poulos, and R.C. Hirschfeld, 2000, Remediation of Seepage Problems through
Cutof or Reduction of Flow and through Collection and Control of Seepage Including the Use of
Geosynthetics, The National Dam Safety Program, Research Needs Workshop: Seepage through Embank-
ment Dams, White Paper No. 5, Atachment 8, Denver, U.S. Department of Homeland Security,
Federal Emergency Management Agency (FEMA), Washington, DC.
Talwani, P. and W.T. Schaefer, 2001, Recurrence Rates of Large Earthquakes in the South Carolina
Coastal Plain Based on Paleoliquefaction Data, Journal of Geophysical Research, Vol. 106, No. B4,
pp. 6621-6642.
Terzaghi, K. and R.B. Peck, 1967, Soil Mechanics in Engineering Practice, 2nd Ed., John Wiley & Sons,
New York, NY.
Terzaghi, K., R.B. Peck and G. Mesri, 1996, Soil Mechanics in Engineering Practice, 3rd Ed., Wiley-
Interscience, New York, NY.
* Thacker, B.K., 2000, Perimeter Embankments to Increase Capacity, Mitigate Subsidence Impacts,
and Abandon Coal Refuse Disposal Impoundments, Proceedings, Tailing Dams 2000, Las Vegas,
ASDSO/USCOLD, pp. 143-154.
Thacker, B.K., 1997, Liquefaction Mitigation Procedures for Coal Refuse Dams Built by the Modi-
fed Upstream Method, Proceedings, XXVIII Ohio River Valley Soils Seminar, Lexington, KY.
Thacker, B.K., C.R. Ulrich, J.C. Athanaskes and G. Smith, 1988, Evaluation of Coal Refuse
Impoundment Built by the Upstream Method, Geotechnical Special Publicaton No. 21, Hydraulic Fill
Structures, American Society of Civil Engineers (ASCE), New York, NY, pp. 730-749.
Thiers, G.R. and H.B. Seed, 1968, Cyclic Stress-Strain Characteristics of Clays, Journal of the Soil
Mechanics and Foundations Division, ASCE, Vol. 94, No. SM2, pp. 555-569.
Timoshenko, S. and S. Woinowsky-Krieger, 1959, Theory of Plates and Shells, 2nd Ed., McGraw-Hill,
New York, NY.
Toro, G.R. and R.K. McGuire, 1987, An Investigation into Earthquake Ground Motion Charac-
teristics in Eastern North America, Bulletin of the Seismological Society of America, Vol. 77, No. 2,
pp. 468-489.
Toro, G.R., N.A. Abrahamson and J.F. Schneider, 1997, Model of Strong Ground Motions from
Earthquakes in Central and Eastern North America: Best Estimates and Uncertainties, Seismologi-
cal Research Leters, Vol. 68, No. 1, pp. 41-57.
Tutle, M.P, E.S. Schweig, J.D. Sims, R.H. Laferty, L.W. Wolf and M.L. Haynes, 2002, The Earth-
quake Potential of the New Madrid Seismic Zone, Bulletin of the Seismological Society of America,
Vol. 92, No. 6, pp. 2080-2089.
Ullrich, C.R., B.R. Thacker and N.R. Roberts, 1991, Dynamic Properties of Fine-Grained Coal
Refuse, Proceedings, Second International Conference on Recent Advances in Geotechnical Earthquake
Engineering and Soil Dynamics, St. Louis, MO, pp. 393-397.
University of British Columbia-Geophysical Inversion Facility (UBC-GIF), 2005, Three-Dimen-
sional Modeling of Self-Potentials Caused by Seepage Flows, Investigation of Geophysical Methods
for Assessing Seepage and Internal Erosion in Embankment Dams, CEA Technologies, Inc. Project No.
T992700-0205, Montreal, QC, Canada.
University of California, 2007a, Pacifc Earthquake Engineering Research Center: NGA Database,
Coyote Lake Gilroy #6, 230 (CDMG Station 57383), August 6, 1979, M = 5.7. (htp://peer.berkeley.
edu/index.html)
University of California, 2007b, Pacifc Earthquake Engineering Research Center: NGA Database,
Saguenay W. Chicoutime Nord (Site 16 T), November 25, 1988, M = 5.9. (htp://peer.berkeley.edu/
index.html)
< PREVIOUS VIEW
R29
Engineering and Design Manual Coal Refuse Disposal Facilities
University of California, 2007c, Pacifc Earthquake Engineering Research Center: NGA Database,
Whitier Narrows Station: 24399, Mt. Wilson CIT Seismic Station October 1, 1987, M = 5.99.
(htp://peer.berkeley.edu/index.html).
* U.S. Army Corps of Engineers (USACE), 2008, Hydrologic Modeling System, HEC-HMS, Users
Manual, Version 3.3, CDP-74A, U.S. Department of the Army, USACE, Washington, DC.
* U.S. Army Corps of Engineers (USACE), 2004, Cathodic Protection for Civil Works Structures, EM
1110-2-2704, U.S. Department of the Army, USACE, Washington, DC.
* U.S. Army Corps of Engineers (USACE), 2003, Engineering and Design Slope Stability, EM 1110-2-
1902, U.S. Department of the Army, USACE, Washington, DC.
U.S. Army Corps of Engineers (USACE), 2002, Coastal Engineering Manual Part II, EM 1110-2-
1100, U.S. Department of the Army, USACE, Washington, DC.
* U.S. Army Corps of Engineers (USACE), 2000, Hydrologic Modeling System HEC-HMS: Technical
Reference Manual, Report No. CPD-74B, U.S. Department of the Army, USACE, Hydrologic Engi-
neering Center, Davis, CA.
* U.S. Army Corps of Engineers (USACE), 1999, Engineering and Design Response Spectra and Seis-
mic Analysis for Concrete Hydraulic Structures, Engineer Manual 1110-2-6050, U.S. Department of the
Army, USACE, Washington, DC.
* U.S. Army Corps of Engineers (USACE), 1998a, Engineering and Design Conduits, Culverts and
Pipes, EM 1110-2-2902, U.S. Department of the Army, USACE, Washington, DC.
* U.S. Army Corps of Engineers (USACE), 1998b, HEC-1 Flood Hydrograph Package: Users Manual,
U.S. Department of the Army, USACE, Hydrologic Engineering Center, Davis, CA.
* U.S. Army Corps of Engineers (USACE), 1995a, Geophysical Exploration for Engineering and Environ-
mental Investigations, EM 1110-1-1802, U.S. Department of the Army, USACE, Washington, DC.
* U.S. Army Corps of Engineers (USACE), 1995b, Construction Control for Earth and Rock-Fill Dams,
EM 1110-2-1911, U.S. Department of the Army, USACE, Washington, DC.
* U.S. Army Corps of Engineers (USACE), 1995c, Instrumentation of Embankment Dams and Levees,
EM 1110-2-1908, U.S. Department of the Army, USACE, Washington, DC.
* U.S. Army Corps of Engineers (USACE), 1994, Hydraulic Design of Flood Control Channels, EM 1110-
2-1601, U.S. Department of the Army, USACE, Washington, DC.
* U.S. Army Corps of Engineers (USACE), 1993, Engineering and Design Seepage Analysis and Con-
trol for Dams, EM 1110-2-1901, U.S. Department of the Army, USACE, Washington, DC.
* U.S. Army Corps of Engineers (USACE), 1990a, Construction with Large Stone, EM 1110-2-2302, U.S.
Department of the Army, USACE, Washington, DC.
* U.S. Army Corps of Engineers (USACE), 1990b, (Revised April 1992), Hydraulic Design of Spillways,
EM 1110-2-1603, U.S. Department of the Army, USACE, Washington, DC.
* U.S. Army Corps of Engineers (USACE), 1986, Laboratory Soils Testing, EM-1110-2-1906, U.S.
Department of the Army, USACE, Washington, DC.
U.S. Army Corps of Engineers (USACE), 1985, Electrical Protection Cathodic Protection, EM-1110-2-
1922, U.S. Department of the Army, USACE, Washington, DC.
* U.S. Army Corps of Engineers (USACE), 1984a, Grouting Technology, EM 1110-2-3506, U.S. Depart-
ment of the Army, USACE, Washington, DC.
* U.S. Army Corps of Engineers (USACE), 1984b, Users Manual, HMR52, Probable Maximum Storm
(Eastern United States), U.S. Department of the Army, USACE, Hydrologic Engineering Center,
Davis, CA.
< PREVIOUS VIEW
R30
References
* U.S. Army Corps of Engineers (USACE), 1980, Hydraulic Design of Reservoir Outlets, EM-1110-2-
1602, U.S. Department of the Army, USACE, Washington, DC.
* U.S. Bureau of Reclamation (USBR), 2007a, Design Standards No. 13 for Embankment Dams: Chapter
5, Protective Filters, U.S. Department of the Interior, USBR, Technical Services Center, Denver, CO.
* U.S. Bureau of Reclamation (USBR), 2007b, Uplif and Crack Flow Resulting from High Velocity
Discharges over Open Ofset Joints: Laboratory Studies, Report DSO-07-07, USBR, Technical Service
Center, Denver, CO.
* U.S. Bureau of Reclamation (USBR), 2001, Water Measurement Manual, 3rd Ed., Revised Reprint,
U.S. Government Printing Ofce, Washington, DC.
* U.S. Bureau of Reclamation (USBR), 1998, Earth Manual, 3rd Ed., U.S. Department of the Interior,
USBR, Earth Sciences and Research Laboratory, Technical Services Center, Denver, CO.
* U.S. Bureau of Reclamation (USBR), 1992a, Design Standards No. 13 Embankment Dams, Chapter 2,
Embankment Design, U.S. Department of the Interior, USBR, Technical Services Center, Denver, CO.
U.S. Bureau of Reclamation (USBR), 1992b, Concrete Manual, Part 2: A Manual for the Control of Con-
crete Construction, 9th Ed., Water Resources Technical Publication, U.S. Department of the Interior,
Technical Services Center, USBR, Denver, CO.
* U.S. Bureau of Reclamation (USBR), 1991, Design Standards No. 13 Embankment Dams, Chapter 10,
Embankment Construction, U.S. Department of the Interior, USBR, Technical Services Center, Denver, CO.
* U.S. Bureau of Reclamation (USBR), 1990, Cavitation in Chutes and Spillways, Engineering Monograph
No. 42, U.S. Department of the Interior, USBR, Denver, CO.
* U.S. Bureau of Reclamation (USBR), 1988, Downstream Hazard Classifcation Guidelines, ACER Tech-
nical Memorandum No. 11, U.S. Department of the Interior, USBR, Denver, CO.
* U.S. Bureau of Reclamation (USBR), 1987a, Design of Small Dams, 3rd Ed., U.S. Department of the
Interior, USBR, Washington, DC.
* U.S. Bureau of Reclamation (USBR), 1989, Design Standards No. 13 - Embankment Dams, Chapter
13, Seismic Design and Analysis, U.S. Department of the Interior, USBR, Technical Services Center,
Denver, CO.
* U.S. Bureau of Reclamation (USBR), 1984, Design Standards No. 13 Embankment Dams, Chapter 3,
Foundation Treatment, U.S. Department of the Interior, USBR, Technical Services Center, Denver, CO.
U.S. Bureau of Reclamation (USBR), 1973, Design of Small Dams, 2nd Ed., U.S. Department of the
Interior, USBR, Washington, DC.
U.S. Commitee on Large Dams (USCOLD), 1999, Updated Guidelines for Selecting Seismic Parameters
for Dam Projects, USCOLD Commitee on Earthquakes, USCOLD, Denver, CO.
U.S. Department of Agriculture (USDA), 1985, Dimensioning of Filter-Drainage Diaphragms for Con-
duits According to TR-60, Technical Note, Series No. 709, USDA, Soil Conservation Service, South
National Technical Center, Forth Worth, TX (see NRCS 2007b for current criteria).
* U.S. Department of Agriculture (USDA), 1978, National Engineering Handbook, Section 8, Engineering
Geology, NEH-8 (now Part 631-Geology), USDA, Soil Conservation Service, Washington, DC.
* U.S. Department of Agriculture (USDA), 1973a, Flow Net Construction and Use, Soil Mechanics
Note No. 5, 210-VI-SMN-5, USDA, Soil Conservation Service, Washington, DC.
* U.S. Department of Agriculture (USDA), 1973b, A Method for Estimating Volume and Rate of Runof
in Small Watersheds, SCS-TP-149, USDA, Soil Conservation Service, Washington, DC.
U.S. Department of Agriculture (USDA), 1972, National Engineering Handbook, Section 4, Hydrology,
NEH-4 (now Part 630-Hydrology), USDA, Soil Conservation Service, Washington, DC.
< PREVIOUS VIEW
R31
Engineering and Design Manual Coal Refuse Disposal Facilities
* U.S. Department of Agriculture (USDA), 1969, Computation of Joint Extensibility Requirements, Tech-
nical Release No. 18, 210-VI-TR-18, USDA, Soil Conservation Service, Washington, DC.
U.S. Department of Agriculture (USDA), 1968, Hydraulics of Broad-Crested Spillways, Technical
Release 39, USDA, Soil Conservation Service, Washington, DC.
* U.S. Department of Agriculture (USDA), 1958, Structural Design of Underground Conduits, Technical
Release No. 5, 210-VI-TR-5, USDA, Soil Conservation Service, Washington, DC.
U.S. Department of Agriculture (USDA), 1956, National Engineering Handbook - Chapter 5: Hydrau-
lics, 210-VI-NEH, USDA, Soil Conservation Service, Washington, DC.
U.S. Department of Energy (USDOE), 1996, Guidelines for the Use of Probabilistic Seismic Hazard
Curves at Department of Energy Sites for Department of Energy Facilities, DOE-STD-1024-92, originally
published December 1992, incorporating Change Notice #1, January 1996, Washington, DC.
* U.S. Environmental Protection Agency (USEPA), 2000, Coal Remining Best Management Practices
Guidance Manual, EPA 821-R-00-007, U.S. Department of the Environment, USEPA, Ofce of Sci-
ence and Technology, Washington, DC.
* U.S. Environmental Protection Agency (USEPA), 1993, Use of Airborne, Surface, and Borehole Geo-
physical Techniques at Contaminated Sites A Reference Guide, EPA/625/R-92/007, U.S. Department of
the Environment, USEPA, Ofce of Research and Development, Washington, DC.
U.S. Environmental Protection Agency (USEPA), 1983, Neutralization of Acid Mine Drainage: Design
Manual, USEPA-600/2-83-001, U.S. Department of the Environment, USEPA, Cincinnati, OH.
* U.S. Geological Survey (USGS), 2005, Comparison of Peak Discharge and Runof Characteristic Esti-
mates from the Rational Method to Field Observations for Small Basins in Central Virginia, Scientifc
Investigation Report 2005-5254, U.S. Department of the Interior, Geological Survey, Denver, CO.
* U.S. Geological Survey (USGS), 1989, Guide for Selecting Mannings Roughness Coefcients for
Natural Channels and Flood Plains, Water Supply Paper 2339, USGS, Denver, CO.
U.S. Nuclear Regulatory Commission (USNRC), 1997, Recommendations for Probabilistic Seismic
Hazard Analysis: Guidance on Uncertainty and Use of Experts, NUREG/CR-6372, Panel on Seismic
Hazard Evaluation, National Research Council, Washington, DC.
U.S. Nuclear Regulatory Commission (USNRC), 1993, Revised Livermore Seismic Hazard Estimates for
69 Nuclear Power Plant Sites East of the Rocky Mountains, NUREG-1488, USNRC, Washington, DC.
* U.S. Society on Dams (USSD), 2007, Strength of Materials for Embankment Dams, USSD Com-
mitee on Materials for Embankment Dams, Denver, CO.
Virginia Department of Mine Land Reclamation (VA DMLR), 2006, Slurry Injection Permit Require-
ments, presented at Underground Injection Workshop sponsored by the Ofce of Surface Mining
(OSM), National Mine Land Reclamation Center, University of West Virginia, Morgantown, WV.
Van Aller, H.W., 2004, Piping, Seepage Conduits and Dam Failure, presented at ASDSO South-
eastern Regional Technical Seminar, Jackson, MS.
* Van Aller, H.W., 1998, Filter Diaphragm Design Considerations: A Collection of Useful Refer-
ences and Design Guidelines, originally presented at 1990 ASDSO Mid-Atlantic Regional Techni-
cal Seminar, Baltimore, MD.
van der Veen, R., 2002, Automatic Data Acquisition Systems and Databases, Geotechnical News,
BiTech Publishers, Richmond, BC, Canada, Vol. 20, No. 1, pp. 24-28.
van Zyl, D. and A.M. Robertson, 1980, Subsurface Drainage of Tailings Impoundments: Some
Design, Construction and Management Considerations, Proceedings, Symposium on Uranium Mill
Tailings Management, Fort Collins, CO, pp. 153-177.
< PREVIOUS VIEW
R32
References
Veneziano, D. and J. Van Dyck, 1985, Analysis of Earthquake Catalogues for Incompleteness
and Recurrence Rates: Seismic Hazard Methodology for Nuclear Facilities in the Eastern United
States, Project No. P101-29, Vol. 2, Appendix A-6, Electric Power Research Institute (EPRI), Palo
Alto, CA.
* Vick, S.G., 2000, Tailings Dam Safety - Implications for the Dam Safety Community, Proceedings,
Tailings Dams 2000, Las Vegas, ASDSO/USCOLD, pp. 1-19.
* Vogel, W.G., 1987, A Manual for Training Reclamation Inspectors in the Fundamentals of Soils and
Revegetation, U.S. Department of the Interior, Ofce of Surface Mining and Enforcement, Washing-
ton, DC.
Von Thun, J.L., 1996, Risk Assessment of Nambe Falls Dam, Geotechnical Special Publication
No. 58, C.D. Shackelford, P.P. Nelson and M.J.S. Roth, eds., American Society of Civil Engineers
(ASCE), Reston, VA, pp. 604-635.
* Wahl, T.L., 1998, Prediction of Embankment Dam Breach Parameters: A Literature Review and Needs
Assessment, Dam Safety Research Report DSO-98-004, U.S. Department of the Interior, Bureau of
Reclamation, Dam Safety Ofce, Denver, CO.
Wald, D.J., V. Quitoriano, T.H. Heaton and H. Kanamori, 1999, Relationships between Peak
Ground Acceleration, Peak Ground Velocity and Modifed Mercalli Intensity in California, Earth-
quake Spectra, Vol. 15, No. 3, pp. 557-564.
* Walker, J.S., 1993, State-of-the-Art Techniques for Backflling Abandoned Mine Voids, Information Circu-
lar No. 9359, U.S. Department of the Interior, Bureau of Mines, Pitsburgh, PA.
Wang, Z., E.W. Woolery, B. Shi and J.D. Kiefer, 2003, Communicating with Uncertainty: A Critical
Issue with Probabilistic Seismic Hazard Analysis, EOS Transactions, American Geophysical Union
(AGU), Vol. 84, No. 46, pp. 501, 506, 508.
Ward, S.H., ed., 1990, Investigations in Geophysics, No. 5: Geotechnical and Environmental Geophysics,
Vols. I-III, Society of Exploration Geophysicists, Tulsa, OK.
Wartman, J., J.D. Bray and R.B. Seed, 2004, Inclined Plane Studies of the Newmark Sliding Block Pro-
cedure, Journal of Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 129, No. 8, pp. 673-684.
Watkins, R.K. and L.R. Anderson, 1999, Structural Mechanics of Buried Pipes, CRC Press, New
York, NY.
Weaver, K.D. and D.A. Bruce, 2007, Dam Foundation Grouting, ASCE Press, Reston, VA.
Wells, D.L. and K.J. Coppersmith, 1994, New Empirical Relationships among Magnitude, Rup-
ture Length, Rupture Width, Rupture Area, and Surface Displacement, Bulletin of the Seismological
Society of America, Vol. 84, No. 4, pp. 974-1002.
West Virginia Department of Transportation (WVDOT), 1999, Materials Procedure, Nuclear Density
Test by the Roller Pass Method, MP 700.00.24.
West Virginia Surface Mine Drainage Task Force, 1979, Suggested Guidelines for Methods of
Operation, Surface Mining of Areas with Potentially Acid-Producing Materials, Proceedings,
West Virginia Surface Mine Drainage Task Force Symposium, West Virginia Department of Natural
Resources, Charleston, WV.
West Virginia Water Research Institute, 2005, Guidance Document for Coal Waste Impoundment Facili-
ties & Coal Waste Impoundment Inspection Form, West Virginia University, Morgantown, WV.
Wetzel, K.L. and S.A. Hofman, 1989, Distribution of Water-Quality Characteristics that May Indicate
the Presence of Acid Mine Drainage in the Eastern Coal Province of the United States, Hydrologic Inves-
tigations Atlas HA-705, U.S. Department of the Interior, Geological Survey, Denver, CO.
< PREVIOUS VIEW
R33
Engineering and Design Manual Coal Refuse Disposal Facilities
Whitaker, B.N. and D.J. Reddish, 1989, Subsidence: Occurrence, Prediction and Control, Elsevier,
Amsterdam, Netherlands.
* Wightman, W.E., F. Jalinoos, P. Sirles and K. Hanna, 2003, Application of Geophysical Methods to
Highway Related Problems, Contract No. DTFH68-02-P-00083, Federal Highway Administration,
Central Federal Lands Highway Division, Lakewood, CO.
Wildeman, T.R., J.J. Gusek, and G.A. Brodie, 1993, Wetland Design for Mining Operations, Bitech
Publishers, Richmond, BC, Canada.
Wilkins, J.K., 1956, Flow of Water through Rockfll and Its Application to the Design of Dams,
Proceedings, 2nd Australia-New Zealand Conference on Soil Mechanics and Foundation Engineering,
International Society for Soil Mechanics and Foundation Engineering, University of Canterbury,
Christchurch, NZ.
Wilson, A.H. and D.P. Ashwin, 1972, Research into the Determination of Pillar Size: An Hypoth-
esis Concerning Pillar Stability, The Mining Engineer, Vol. 131, No. 141, pp. 409-427.
Wilson, S.D. and L.R. Squier, 1969, Earth and Rockfll Dams, Proceedings, Seventh International
Conference on Soil Mechanics and Foundation Engineering, Mexico, pp. 137-223.
Wizniak, L. and R.J. Mitchell, 1987, Crown Pillar Subsidence Prediction from Centrifuge Models,
Proceedings, International Symposium on Prediction and Performance in Geotechnical Engineering, Cal-
gary, Alberta, Canada, R.C. Joshi and F.J. Grifths, eds., pp. 387-395.
Workhorse Technologies, 2006, personal communication.
Workman, J.L. and P.C. Satchwell, 1987, Blasting as an AML Reclamation Method, 2 Vols., U.S. Depart-
ment of the Interior, Ofce of Surface Mining, Washington, DC.
Wride, C.E., E.C. McRoberts and P.K. Robertson, 1999, Reconsideration of Case Histories for
Estimating Undrained Shear Strength in Sandy Soils, Canadian Geotechnical Journal, Vol. 36, No. 5,
pp. 907-933.
Wu, K.K., H.L. Owens and J.W. Fredland, 2003, MSHAs Review of Impoundment Plans, 2003
SME Annual Meeting, Preprint 03-099, Cincinnati, Society for Mining Metallurgy and Exploration
(SME), Litleton, CO.
Wyllie, D.C. and C.W. Mah, 2004, Rock Slope Engineering, Spon Press, New York, NY.
* Wyrick, G.G. and J.W. Borchers, 1981, Hydrologic Efects of Stress-Relief Fracturing in an Appalachian
Valley, Water Supply Paper 2177, U.S. Department of the Interior, U.S. Geological Survey, Reston, VA.
Yearly, D.C., 2003, Sustainable Development for the Global Mining and Metals Industry, Mining
Engineering, Vol. 55, No. 8, pp. 45-48.
Yoshimine, M., P.K. Robertson, and C.E. Wride, 1999, Undrained Shear Strength of Clean Sands to
Trigger Flow Liquefaction, Canadian Geotechnical Journal, Vol. 36, No. 5, pp. 891-906.
Youd, T. L., I.M. Idriss, R.D. Andrus, I. Arango, G. Castro, J.T. Christian, R. Dobry, W.D.L. Finn,
L.F. Harder, M.E. Hynes, K. Ishihara, J.P. Koester, S.S.C. Liao, W.F. Marcuson, G.R. Martin, J.K.
Mitchell, Y. Moriwaki, M.S. Power, P.K. Robertson, R.B. Seed and K.H. Stokoe, 2001, Liquefaction
Resistance of Soils: Summary Report from the 1996 NCEER and 1998 NCEER/NSF Workshops on
Evaluation of Liquefaction Resistance of Soils, Journal of Geotechnical and Geoenvironmental Engi-
neering, ASCE, Vol. 127, No. 10, pp. 817-833.
Young, W.C. and R.G. Budynas, 2001, Roarks Formulas for Stress and Strain, 7th Ed., McGraw-Hill,
New York, NY.
Zeng, X. and J.A. Goble, 2008, Dynamic Properties of Coal Waste Refuse in a Tailings Dam, Geotech-
nical Special Publication No. 181, Geotechnical Earthquake Engineering and Soil Dynamics IV, D. Zeng,
M.T. Manzari, and D.R. Hiltunen, eds., American Society of Civil Engineers (ASCE), Reston, VA.
< PREVIOUS VIEW
R34
References
Zeng, Y., J.G. Anderson, and G. Yu, 1994, A Composite Source Model for Computing Realistic
Synthetic Strong Ground Motions, Geophysical Research Leters, Vol. 21, No. 8, pp. 725-728.
Zhang, J., W. Wagner, A. Prakash, H. Mehl and S. Voight, 2004, Detecting Coal Fires Using
Remote Sensing Techniques, International Journal of Remote Sensing, Vol. 25, No. 16, pp. 3193-3220.
< PREVIOUS VIEW