0% found this document useful (0 votes)
171 views

DrainageDesignGuide PDF

Uploaded by

Eric Chan
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
171 views

DrainageDesignGuide PDF

Uploaded by

Eric Chan
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 795

STATE OF FLORIDA DEPARTMENT OF TRANSPORTATION

DRAINAGE DESIGN GUIDE

OFFICE OF DESIGN, DRAINAGE SECTION JANUARY 2019


TALLAHASSEE, FLORIDA
January 2019

This page left blank intentionally.


January 2019

Drainage Design Guide


Table of Contents

TABLE OF CONTENTS

Chapter 1 Introduction
Chapter 2 Hydrology
Chapter 3 Open Channel
Chapter 4 Culvert
Chapter 5 Bridge Hydraulics
Chapter 6 Storm Drains
Chapter 7 Exfiltration Systems
Chapter 8 Optional Pipe Material
Chapter 9 Stormwater Management Facility
Chapter 10 Temporary Drainage Design

APPENDICES

Appendix A Data Collection/Published Data


Appendix B Hydrology Design Aids
Appendix C Open Channel Flow Design Aids
Appendix D Gutter Flow Using HEC-RAS
Appendix E Partial Depth Pipe Flow Graphs
Appendix F Application Guidelines for Pipe End Treatments
Appendix G Risk Evaluations
Appendix H Shoulder Gutter Transition Slope at Bridges
Appendix I Inlet Efficiencies
Appendix J Exfiltration Design - Hydro-Geotechnical Aids - K for Saturated and
Compacted Lab Soil Specimens
Appendix K Drainage Well Design - Reasonable Assurance Report
Appendix L Exfiltration Design - Sample Plans
Appendix M Pipe Corrosion Tables and Charts
Appendix N A Rational for Stormwater Rule Standards

Table of Contents i
January 2019

Drainage Design Guide


Chapter 1: Introduction

CHAPTER 1: INTRODUCTION

1. INTRODUCTION ................................................................................................... 1-1

1.1 Background ...................................................................................................... 1-1

1.2 Purpose ............................................................................................................. 1-1

1.3 Revisions .......................................................................................................... 1-2

1.4 Definitions of Terms and Acronyms ............................................................... 1-3

Chapter 1: Introduction 1-i


January 2019

Drainage Design Guide


Chapter 1: Introduction

1. INTRODUCTION

1.1 BACKGROUND
In 1987, Florida Department of Transportation (FDOT) published the Drainage Manual as
a three volume set: Volume 1 - Policy; Volumes 2A and 2B - Procedures; Volume 3 -
Theory.
In October 1992, the FDOT revised Volume 1 - Policy to Volume 1- Standards and
designated Volumes 2A, 2B, and 3 as general reference documents.
In January 1997, the FDOT renamed Volume 1 - Standards to “Drainage Manual”. In the
years that followed, the FDOT developed numerous handbooks to replace Volumes 2A,
2B, and 3 of the original 1987 Drainage Manual. With this, the Drainage Manual was
maintained as a “standards” document while the handbooks provided guidance
addressing drainage design practice, analysis and computational methods, along with
design aids and reference material.
In 2016, the Department consolidated the handbooks into the Drainage Design Guide.
Chapters 2 through 10 of the Drainage Design Guide each represent a handbook in
previous form. The appendices of the handbooks, with a few exceptions, were
incorporated as appendices in the Drainage Design Guide. Whereas, the remaining
handbook appendices were inserted into the appropriate chapter of the Drainage Design
Guide.

1.2 PURPOSE
The Drainage Design Guide is a reference for designers, which provides guidelines for
common drainage and stormwater aspects of FDOT projects. The guidelines do not
replace the need for professional engineering judgment or preclude the use of other
information. These guidelines are suggested or preferred approaches, not requirements.
The Drainage Manual provides minimum standards and governs over the Drainage
Design Guide, when discrepancies are noted between both documents.
The technical information in these guidelines is written by Central Office Drainage and is
then reviewed and commented upon by the district drainage engineers. The district
drainage engineer has the final project specific decisions concerning the application of
these guidelines, especially given the subjective judgment required to do good drainage
design. If you have project specific questions on this material, please collaborate with
your district drainage engineer.

Chapter 1: Introduction 1-1


January 2019

Drainage Design Guide


Chapter 1: Introduction

1.3 REVISIONS
Any comments or suggestions concerning this handbook may be made by e-mailing the
State Drainage Engineer. The FDOT will routinely make revisions to keep the Drainage
Design Guide consistent with other FDOT documents and to reflect changes and trends
in drainage design.

Chapter 1: Introduction 1-2


January 2019

Drainage Design Guide


Chapter 1: Introduction

1.4 DEFINITIONS OF TERMS AND ACRONYMS


AASHTO American Association of State Highway and Transportation
Officials
Abstraction Hydrologic processes that remove water from precipitation
before it becomes surface runoff; types include evaporation,
infiltration, transpiration, interception, depression storage, and
detention storage.
Abutment The portion of a bridge containing the embankment at each
end of the bridge. Abutments may be sloped or vertical.
Accretion The build-up of land or bottom elevation.
ADA Americans with Disabilities Act
Aggradation The build-up of a stream bed over time along the entire stream
reach due to deposition of sediments eroded from the channel
or banks farther upstream in the watershed.
Annulus The area between the outside of a pipe and the precast
opening in which the pipe is placed.
ASTM American Society for Testing and Materials
Attenuation In flood control: to temporarily hold back or store stormwater
to control the rate of discharge. Also, see Detention.
Backwater Backwater is defined as the increase of water surface
elevation induced upstream from a bridge, culvert, dike, dam,
another stream at a higher stage, or other similar structures;
or as conditions that obstruct or constrict a channel relative to
the elevation occurring under natural channel and floodplain
conditions.
Bay In coastal hydrology: a recess in the shore or an inlet of a sea
between two capes or headlands; a bay is not as large as a
gulf, but larger than a cove.
Berm An embankment typically used for containment or separation
of water.
BMP Best Management Practice. Refers to standard practices used
to improve stormwater quality prior to discharge.
CFS Cubic Feet per Second

Chapter 1: Introduction 1-3


January 2019

Drainage Design Guide


Chapter 1: Introduction

Channel section The cross section of a channel taken at an angle perpendicular


to the direction of water flow in the channel.
Conveyance A measure of the carrying capacity of a channel or pipe
section. Often denoted as “K”. K = Q/(slope)0.5.

Coefficient of A measure of the rate of flow of water through a medium (soil,


permeability membrane, fabric, etc.) under a given hydraulic gradient in
units of length/time (i.e., ft/day; cm/sec).
Critical depth (Dc) The depth associated with the minimum total energy for a
particular flow rate in a particular cross section. The flow depth
can drop through critical depth at the outlet of a pipe section if
the water surface downstream is low enough.
Critical duration As defined by Rule 14-86 F.A.C.: “Critical Duration” means the
length of time of a specific storm frequency that creates the
largest volume or highest rate of net stormwater runoff (post-
improvement runoff less pre-improvement runoff) for typical
durations up through and including the 10-day duration for
closed basins and up through the 3-day duration for basins
with positive outlets. The critical duration for a given storm
frequency is determined by calculating the peak rate and
volume of stormwater runoff for various storm durations and
then comparing the pre-improvement and post-improvement
conditions for each of the storm durations. The duration
resulting in the highest peak rate or largest net total stormwater
volume is the “critical duration” storm (volume is not applicable
for basins with positive outlets). See Chapter 9 for additional
discussion.
Cross drain A structure supporting a public roadway that crosses
transversely over a watercourse.
Curve number A dimensionless site-specific runoff parameter developed by
the (former) Soil Conservation Service (now Natural
Resources Conservation Service) to empirically estimate
rainfall excess; it accounts for infiltration losses and initial
abstractions.

Chapter 1: Introduction 1-4


January 2019

Drainage Design Guide


Chapter 1: Introduction

Darcy’s Law Darcy’s Law characterizes the flow through porous media,
assuming that the viscosity, temperature, and density of the
fluids are constant. The flow rate is a function of the
proportionality constant (coefficient of permeability), the
hydraulic gradient, and the flow area; Q = k i A.
Degradation The lowering of land or bottom elevation. In stream stability
assessment, the lowering occurs through natural erosion of
sediment without sufficient incoming sediment to replenish.
DEM Digital Elevation Model
Department Florida Department of Transportation
Depth of flow The vertical distance between the lowest point of a channel
section and the free surface.
Detention To temporarily hold back or store stormwater to control the rate
of discharge. Normally, the term “Wet Detention” is associated
with water quality treatment. Sometimes the term is used for
flood control attenuation.
Drainage Manual Refers to the current release of the Florida Department of
Transportation Drainage Manual

Diurnal tide The diurnal tide is represented by one high tide and one low
tide per day.
Diversion structure For stormwater treatment, a diversion structure may be used
to divert the “first flush” of stormwater to a facility for treatment.
Drainage basin A subdivision of a watershed.
Duration The time from beginning to end of a rain storm event used to
perform runoff calculations.
Ebb phase The period when the water level of the tide is falling.
ECB Erosion Control Blanket. A temporary degradable mat
composed of natural or polymer fibers used to reduce erosive
impact in low-velocity ditches during short periods of
construction. (See Ch. 3.)
Environmental Conceptual approval granted via an individual or general
Resource Permit permit for a surface water management system issued
(ERP) pursuant to Part IV, Chapter 373, Florida Statutes.

Chapter 1: Introduction 1-5


January 2019

Drainage Design Guide


Chapter 1: Introduction

Estuary A body of water affected by tidal influence as well as


freshwater inflows from a riverine system.
Exfiltration The loss of water from a drainage system as a result of
percolation or absorption into the surrounding soil.
Exfiltration trench A subsurface system consisting of a conduit, such as a
perforated pipe, surrounded by natural or artificial aggregate
that temporarily stores and filters stormwater runoff. Also
known as a French Drain.
FDEP Florida Department of Environmental Protection
FDM FDOT Design Manual
FDOT Florida Department of Transportation
FHWA Federal Highway Administration
Flood Inundation of land by water to depths greater than typically
occur during a normal wet season. See Chapter 4 for
definitions of: Design, Base, Greatest, and Overtopping
Floods.
Flood hydrograph A continuous plot of the surface runoff flow rate versus time.
The volume is equal to the volume of water contained in the
rainfall excess hyetograph.
Flood phase The period when the water level of the tide is rising.

FM Florida Method of Testing Materials. This is the standard


FDOT method of testing materials.
Frequency In hydrology, frequency is the inverse value of the anticipated
recurrence interval. A 4-percent chance of recurrence (of
rainfall or flood event) in any year is referred to as a 25-year
frequency.

Chapter 1: Introduction 1-6


January 2019

Drainage Design Guide


Chapter 1: Introduction

Froude Number (Fr) The Fr value is the dimensionless ratio of inertial forces to
gravity forces. If Fr values are less than 1, gravity forces
dominate and the open channel is said to be operating in the
sub-critical range of flow. If Fr values are greater than 1, inertial
forces dominate and the open channel is said to be operating
in the super-critical range of flow.
v
Fr = 1
2
(gL)
Full flow friction loss For pipes flowing full, the full flow friction loss is the full flow
friction slope times the pipe length.
Full flow friction The slope obtained from Manning’s Equation using an area
slope equal to the full cross sectional area of the pipe and a flow rate
equal to the design flow rate.
S = [Qn /(1.49AR2/3)] 2
Where:
Q = design flow rate
A & R = based on full cross section area of pipe
Gabions Wire mesh forms filled with stones. These include mattresses
and baskets.
Grout-filled Woven fabric forms that are filled with concrete grout. These
mattresses include Filter Point Linings and Articulating Block Mats.
Gutter drain A pipe, used along steep slopes, to convey stormwater from
shoulder gutter inlets on elevated roadways to drainage
conveyance systems below at a much lower elevation.
HEC Hydraulic Engineering Circular. Produced by the FHWA.
HGL Hydraulic grade line. In open channel flow, it is the water
surface along the channel reach. In pressure flow, it is a
theoretical line connecting hydraulic gradient points (points to
which the water would rise in a tube or inlet connecting the flow
pipe to atmospheric pressure) along the flow path.
HG Hydraulic gradient. The difference in water surface divided by
the flow distance (dimensionless value often expressed in
percent).

Chapter 1: Introduction 1-7


January 2019

Drainage Design Guide


Chapter 1: Introduction

Hindcast To retrospectively employ measured data to develop a model


wind or wave field for a specific historical event.
Hydraulic The ratio of discharge perpendicular through a unit area per
conductivity unit of head (i.e., cfs/ft2 - ft).
Hydraulic depth The ratio of the water flow cross section area to top width.
A
D=
T
Hydraulic head The difference in water surface (i.e., potential energy)
available to drive flow (between an inlet and an outlet;
upstream to downstream; through a filter, etc.)
Hydraulic radius The ratio of the water flow cross sectional area to its wetted
perimeter.
A
R=
P
Hydrology The science dealing with the disposition of water on the Earth.
Hyetograph A graphical representation of the distribution of rainfall over
time.
Infiltration Abstraction process in which water flows or is absorbed into
the ground.
Infiltration rate The maximum rate at which water can enter the soil from the
surface under specified conditions. The units are length per
time.
Inlet In coastal hydrology: a short, narrow waterway connecting a
bay, lagoon, or similar body of water with a large parent body
of water.
Intensity The rate of precipitation, usually in inches/hour.
Karst A geological term to describe a landform underlain by highly
porous limestone rock with solution channels. Springs,
disappearing streams, and sinkholes are typical of Karst
topography. “Closed basins” are associated with Karst
topography.

Chapter 1: Introduction 1-8


January 2019

Drainage Design Guide


Chapter 1: Introduction

LiDAR Light Detection And Ranging. This is a remote-sensing


method that uses light in the form of a pulsed laser to measure
ranges (variable distances) to the Earth. These light pulses—
combined with other data recorded by the airborne system—
generate precise, three-dimensional information about the
shape of the Earth and its surface characteristics.
Manning’s Equation A formula used to estimate the average velocity of a liquid
flowing in a conduit that does not completely enclose the liquid,
i.e., open channel flow.
MHW Mean High Water. The average height of tidal high waters over
a 19-year period. For shorter periods of observations,
corrections are applied to eliminate known variations and
reduce the results to the equivalent of a mean 19-year value.
All high water heights are included in the average where the
type of tide is semi-diurnal or mixed. Only the higher high water
heights are included in the average where the type of tide is
diurnal. So determined, mean high water in the latter case is
the same as mean higher high water.
MHHW Mean Higher High Water. The average of the higher high water
height of each tidal day observed over the National Tidal
Datum Epoch. For stations with shorter series, comparison of
simultaneous observations with a control tide station is made
to derive the equivalent datum of the National Tidal Datum
Epoch. For locations with diurnal tides—one high tide and one
low tide per day—this datum will be unavailable. At most
locations, there are semi-diurnal tides—the tide cycles through
a high and low water level twice each day, with one of the two
high tides being higher than the other and one of the two low
tides being lower than the other.

Chapter 1: Introduction 1-9


January 2019

Drainage Design Guide


Chapter 1: Introduction

MLW Mean Low Water. The average height of the low waters over a
19-year period. For shorter periods of observations,
corrections are applied to eliminate known variations and
reduce the results to the equivalent of a mean 19-year value.
All low water heights are included in the average where the
type of tide is either semi-diurnal or mixed. Only lower low
water heights are included in the average where the type of
tide is diurnal. So determined, mean low water in the latter
case is the same as mean lower low water.
MLLW Mean Lower Low Water. The average of the lower low water
height of each tidal day observed over the National Tidal
Datum Epoch. For stations with shorter series, comparison of
simultaneous observations with a control tide station is made
to derive the equivalent datum of the National Tidal Datum
Epoch. For locations with diurnal tides—one high tide and one
low tide per day—this datum will be unavailable. At most
locations, there are semi-diurnal tides—the tide cycles through
a high and low water level twice each day, with one of the two
high tides being higher than the other and one of the two low
tides being lower than the other.
MSL Mean Sea Level. The arithmetic mean of hourly heights
observed over the National Tidal Datum Epoch. Shorter series
are specified in the name; i.e., monthly mean sea level and
yearly mean sea level.

MTL Mean Tide Level. The arithmetic mean of mean high water and
mean low water.
Minor losses All losses that are not due to friction. Generally, these are
energy losses due to changes or disturbances in the flow path.
Minor losses include entrance, exit, bend, and junction losses.
Neap tide Tide of decreased range occurring semi-monthly as the result
of the moon being in quadrature.
NFIP National Flood Insurance Program. Administered by the
Federal Emergency Management Agency (FEMA) pursuant to
44 Code of Federal Regulations (CFR) parts 59 through 80.
Part 65 pertains to mapping of Special Hazard Areas.

Chapter 1: Introduction 1-10


January 2019

Drainage Design Guide


Chapter 1: Introduction

NHW Normal High Water. For bridge hydraulics, the water stage
associated with a flow that has a 43-percent chance of
recurrence (2.33-year frequency) in a given year. In some
cases, stain lines may be used to estimate NHW.
Non-uniform flow A flow condition where the depth of flow changes with respect
to distance along a channel or conduit. Non-uniform flow may
be classified as either rapidly varied or gradually varied.
Rapidly varied flow also is known as a local phenomenon,
examples of which include the hydraulic jump and hydraulic
drop. The primary example of gradually varied flow occurs
when sub-critical flow is restricted by a culvert or storage
reservoir. The water surface profile caused by such a
restriction generally is referred to as a backwater curve.
Normal depth The depth of flow in a channel determined by the channel
properties and physical slope using Manning’s Equation. The
solution is not direct because the channel depth is unknown
and, therefore, requires an iterative process using trial and
error to solve implicitly for depth.
NRCS Natural Resources Conservation Service (formerly Soil
Conservation Service)
NTDE National Tidal Datum Epoch. The specific 19-year period
adopted by the National Ocean Service as the official time
segment over which tide observations are taken and reduced
to obtain mean values (e.g., mean lower low water, etc.) for
tidal datums. It is necessary for standardization because of
periodic and apparent secular trends in sea level. The present
NTDE is 1983 through 2001 and is actively considered for
revision every 20 years to 25 years. Tidal datums in certain
regions with anomalous sea level changes (Alaska, Gulf of
Mexico) are calculated on a Modified 5-Year Epoch.
NWFWMD Northwest Florida Water Management District
Open channel flow Fluid flow in which the liquid surface is subject to atmospheric
pressure (i.e., has an open or free water surface). Open
channel conditions are the basis for most hydraulic
calculations.

Chapter 1: Introduction 1-11


January 2019

Drainage Design Guide


Chapter 1: Introduction

Overland flow Water that travels over the ground surface to the stream
channel, usually limited to a maximum length of 100 feet.
Physical velocity The velocity in a pipe that is flowing full, but not under
pressure. This condition is sometimes called gravity full flow
and the velocity is determined from Manning's Equation.
Actual velocity may be greater than or less than physical
velocity depending on actual flow conditions.
Positive outlet As defined by Rule 14-86 F.A.C.: A point of stormwater
discharge into surface waters that, under normal conditions,
would drain by gravity through surface waters ultimately to the
Gulf of Mexico, the Atlantic Ocean, or into sinks or closed lakes
provided the receiving water body has been identified by the
appropriate Water Management District as functioning as if it
recovered from runoff by means other than transpiration,
evaporation, percolation, or infiltration.
Prismatic channel An artificial channel with non-varying cross section and
constant bottom slope.
Recovery time For stormwater facilities; the time it takes to recover the
volume of water stored above the facility control elevation.
Regression equation A statistical method that correlates peak discharge with
physical features such as watershed area and stream slope.
Retention To retain stormwater and prevent any surface water discharge.
The retained stormwater is either infiltrated into the ground or
evaporated.
Riverine flow For bridge hydraulics, those crossings with no tidal influence
during the design storm, such as (a) inland rivers, or (b)
controlled canals with a salinity structure oceanward
intercepting the design hurricane surge.
Runoff Precipitation remaining after appropriate hydrologic
abstractions have been accounted for.
Runoff coefficient Empirical parameter used to calculate rainfall excess as a fixed
percentage of precipitation; it accounts for interception,
surface storage, and infiltration.

Chapter 1: Introduction 1-12


January 2019

Drainage Design Guide


Chapter 1: Introduction

Scour Erosion of streambed material, typically at hydraulic


conveyance. See Chapters 5, 4, and 3.
Scupper A drain used on a bridge deck that has a free discharge (as
opposed to drainage collected in a pipe system or down-drain).
Semi-diurnal tide Two high tides and two low tides per day.
SFWMD South Florida Water Management District
SHWT Seasonal High Water Table. Elevation to which the ground and
surface water can be expected to rise due to a normal wet
season.
Side drain A side drain conveys non-public access roads across roadside
swales or ditches.
Significant wave The average height of the one-third highest waves of a given
height wave group. Note that the composition of the highest waves
depends upon the extent to which the lower waves are
considered.
SJRWMD St. Johns River Water Management District
Skimmer A continuous baffle around a discharge structure or weir that
skims floatable debris and oil upstream while allowing flow
under the lower edge toward the discharge structure.
Spit A small point of land or a narrow shoal projecting into a body
of water from the shore.
Spread The horizontal distance of the stormwater flowing down a
pavement and gutter section from the face of the gutter to the
water’s edge.
Spring tide A tide that occurs at or near the time of the new or full moon
and which rises highest and falls lowest from the mean sea
level.
SRWMD Suwannee River Water Management District
Stage The elevation or vertical distance of the free surface above a
given point.
Standard Plans Standard Plans for Road and Bridge Construction
State water quality Water quality standards adopted by the state pursuant to
standards Chapter 403, Florida Statutes.

Chapter 1: Introduction 1-13


January 2019

Drainage Design Guide


Chapter 1: Introduction

Steady flow A flow condition where the discharge or rate of flow at any
location along a channel or conduit remains constant with
respect to time. The maintenance of steady flow in any channel
reach requires that the rates of inflow and outflow be constant
and equal.
Storm surge A long wave generated offshore that may propagate into
coastal bays and estuaries. The five components of storm
surge are: wind setup, atmospheric pressure setup, Coriolis
effect, wave setup, and the rainfall effect.

Stormwater injection Wells used for stormwater runoff disposal into pervious
wells underground soils or the water table.
Swell Wind-generated waves that have traveled out of their
generating area. Swell characteristically exhibits a more
regular and longer period and has flatter crests than waves
within their fetch.
SWFWMD Southwest Florida Water Management District
tc Time of concentration. The time required for runoff to travel
from the hydraulically most distant point of a watershed to the
design point.
Tailwater The water surface elevation at the downstream end of a
hydraulic conveyance.
Thalweg In hydraulics, the line joining the deepest points along a flow
path.
Tidally dominated For bridge hydraulics, crossings where the tidal influences are
flow dominated by the design hurricane surge. Large bays, ocean
inlets, and open sections of the Intracoastal Waterway typically
are tidally dominated so much so that even extreme rainfall
events have little influence on the design flows in these
systems.
Tidally influenced Flows in tidal creeks and rivers opening to tidally dominated
flow waterways are affected by both river flow and tidal fluctuations.
Tidally affected river crossings do not always experience flow
reversal; however, backwater effects from the downstream
tidal fluctuation can induce water surface elevation fluctuations
up through the bridge reach.

Chapter 1: Introduction 1-14


January 2019

Drainage Design Guide


Chapter 1: Introduction

TN Total nitrogen. Various species of nitrogen, both particulate


and dissolved.
Top width The width of the channel section at the free surface.
TP Total phosphorus. Various species of phosphorus, both
particulate and dissolved.
Treatment Generally referring to stormwater management practices to
improve the quality of stormwater discharged.
Treatment volume The volume of runoff usually associated with the first flush of
pollutants, which must be retained, detained, or filtered to
remove pollutants and improve discharge water quality.
TRM Turf Reinforcement Mat. A long-term, non-biodegradable mat
composed of synthetic fibers used to increase erosion
resistance in ditches during long periods of construction. (See
Ch. 3.)
Turbulent flow A flow condition where the viscous forces are weak relative to
the inertial forces. In turbulent flow, the water particles move in
irregular paths that are neither smooth nor fixed, and the result
is a random mixing motion. Turbulent flow is the most common
type occurring in roadway drainage facilities.
Underdrain system For stormwater management facilities; a system of perforated
pipes below a pond that are designed to lower the groundwater
table to facilitate pond volume recovery, and/or to filter
stormwater runoff prior to discharge.
Uniform flow A flow condition where the mean velocity and depth of flow are
constant with respect to distance along a channel or conduit of
constant cross section, slope, and roughness. When the
requirements for uniform flow are met, the depth of flow for a
given discharge is defined as the normal depth of flow.
Unsteady flow A flow condition where the discharge at any location in the
channel changes with respect to time. During periods of
stormwater runoff, the inflow hydrograph to an open channel
is usually unsteady. However, in practice, open channel flow
is generally assumed to be steady at the discharge rate for
which the channel is being designed (i.e., peak discharge of
the inflow hydrograph).

Chapter 1: Introduction 1-15


January 2019

Drainage Design Guide


Chapter 1: Introduction

USDW Underground source drinking water. An aquifer that contains a


total of dissolved solids concentration of less than 10,000
milligrams per liter or parts per million (ppm).
Velocity head The velocity head represents the kinetic energy of the fluid per
unit volume and is computed by:
Where is the kinetic correction factor for non-
uniform velocity distribution.
Or, ignoring the effect of a non-uniform velocity distribution,
velocity head is v2/2g
Watershed An area bounded peripherally by a drainage divide that
concentrates runoff to a particular watercourse or body; the
catchment’s area or drainage basin from which the waters of a
stream are drawn.
Watershed lag time Time from the center of mass of the rainfall excess to the runoff
hydrograph peak.
Wave height The vertical distance between a wave’s crest and the
preceding trough.
Wave radiation Excess flow of momentum in the horizontal plane due to
stress waves.
Wave runup The vertical distance above the still water level where breaking
waves propel water up a sloping surface.
Wave setup Vertical increase in the water surface above the still water level
near shore due to onshore mass transport of water due to
wave radiation stresses.
Wave shoaling Transformation of wave profile due to inshore propagation.
Weir A flow restriction with a fixed flowline, width, and height; used
to control discharge from a stormwater management facility.
Well casing A well casing serves as a lining to limit discharge to the aquifer.
It also provides structural support against caving materials
outside the well. Materials commonly used are wrought iron
and steel.
Wetted perimeter The length of the line of intersection of the channel wetted
surface with a cross-sectional plane perpendicular to the
direction of flow.

Chapter 1: Introduction 1-16


January 2019

Drainage Design Guide


Chapter 1: Introduction

Wind set-down The vertical drop below the still water level on the windward
side of a water body due to wind stresses on the surface of the
water.
Wind setup The vertical rise above the still water level on the leeward side
of a water body due to wind stresses on the surface of the
water.
Wind wave Waves being formed and built up by the wind.
WMD Water Management District

Chapter 1: Introduction 1-17


January 2019

Drainage Design Guide


Chapter 2: Hydrology

CHAPTER 2: HYDROLOGY

2. HYDROLOGY ....................................................................................................... 2-1

2.1 Drainage Data ...................................................................................................... 2-1

2.2 Procedure Selection ........................................................................................... 2-1


2.2.1 Rainfall Data .................................................................................................. 2-2
2.2.2 Time of Concentration.................................................................................... 2-3
2.2.2.1 Velocity Method ...................................................................................... 2-3
2.2.2.2 Kirpich (1940) Equation .......................................................................... 2-6
2.2.3 Peak Runoff Rates—Ungaged Sites .............................................................. 2-7
2.2.3.1 Rational Equation ................................................................................... 2-7
2.2.3.2 Regression Equations ........................................................................... 2-10
2.2.4 Flood Hydrographs ...................................................................................... 2-17
2.2.4.1 Modified Rational Method ..................................................................... 2-18
2.2.4.2 NRCS Hydrograph ................................................................................ 2-18

Chapter 2: Hydrology 2-i


January 2019

Drainage Design Guide


Chapter 2: Hydrology

2. HYDROLOGY

2.1 DRAINAGE DATA


Identifying drainage data needs should be a part of the early design phase of a project,
best accomplished at the same time that you select appropriate procedures for performing
hydrologic and hydraulic calculations. Several categories of data may be relevant to a
particular project:

• Published data on precipitation, soils, land use, topography, streamflow, and flood
history
• Field investigations and surveys to:
o determine drainage areas
o identify pertinent features
o obtain high water information
o survey lateral ditch alignments
o survey bridge and culvert crossings

Information on types of data available and the sources of that data are presented in
Appendix A of this document.

2.2 PROCEDURE SELECTION


Occasionally, you will be able to find streamflow measurements for determining peak
runoff rates for pre-project conditions. Where measurements are available, the Florida
Department of Transportation (Department) usually relies upon agencies such as the
United States Geological Survey (USGS) to perform the statistical analysis of streamflow
data; however, guidelines for determining flood flow frequencies from observed
streamflow data may be obtained from Bulletin 17B of the U.S. Water Resources Council
(revised 1981, 1982).

Where streamflow measurements are not available, it is accepted practice to estimate


peak runoff using the rational method or one of the regression equations developed for
Florida. In general, you should use the method that best reflects project conditions, while
also documenting the reasons for using that method.

Considering peak runoff rates for design conditions generally is adequate for conveyance
systems such as storm drains or open channels. However, if the design must include
flood routing (e.g., storage basins or complex conveyance networks), you typically must
create a flood hydrograph. Computer programs are available to help develop a runoff
hydrograph.

Chapter 2: Hydrology 2-1


January 2019

Drainage Design Guide


Chapter 2: Hydrology

In general, you can apply procedures using streamflow analysis and unit hydrograph
theory to all watershed categories.

Table 2.2-1 shows guidelines for selecting peak runoff rate and flood hydrograph
procedures.

TABLE 2.2-1
GUIDELINES FOR SELECTING PEAK RUNOFF RATE AND FLOOD HYDROGRAPHS

Peak Runoff Rates Flood Hydrographs


a
Natural Developed Modified
Flow Developed Developed Leon Rational Method
Watershed Streamflow Rational USGS USGS Tampa County or NRCS
Application Category Analysis Method Equations Equations Equations Equations Unit Hydrograph

Storm Drains 0 to 600 acres X X

Cross Drains 0 to 600 acres X X X X X


Side Drains
600+ acres X X X

Stormwater None X X
Management

a
The modified rational method is not recommended for drainage basins with tc greater than 15 minutes.

2.2.1 Rainfall Data


The Department has developed Intensity-Duration-Frequency (IDF) curves for 11 zones
in Florida using depth-duration-frequency data from HYDRO-35 and TP-40. The curves
are available on the Drainage Internet Site. The IDF curves developed by the Department
provide a reasonable basis for design, and—in areas where intensities would vary—they
reflect values near locations of higher development.

Depth-frequency data for durations of 1, 2, 4, 7, and 10 days, which depict maps of Florida
with contours of precipitation depth for return period frequencies of 2, 5, 10, 25, 50, and
100 years also are available on the Internet.

Frequency can be defined either in terms of an exceedance probability or a return period.


Exceedance probability is the probability that an event having a specified volume and
duration will be exceeded in a specified time period (usually one year). Return period is
the average length of time between events having the same volume and duration. The
problem with using return period is that it can be misinterpreted. If a 50-year flood occurs
one year, some people believe that it will be 50 years before another flood of that
magnitude occurs. Instead, because floods occur randomly, there is a finite probability
that the 50-year flood could occur in two consecutive years.
The exceedance probability (p) and return period (T) are related as follows:

Chapter 2: Hydrology 2-2


January 2019

Drainage Design Guide


Chapter 2: Hydrology

1
p= (2.2-1)
T
A 25-year storm has a 0.04 or 4-percent exceedance probability (probability of occurrence
in any given year), a 50-year storm has a 0.02 or 2-percent exceedance probability, etc.

2.2.2 Time of Concentration


The time of concentration is defined as the time it takes runoff to travel from the most
remote point in the watershed to the point of interest. You can use either of the following
methods for calculating the time of concentration:

2.2.2.1 Velocity Method


The Velocity Method is a segmental approach, which you can use to account for
overland flow, shallow channel flow (rills or gutters), and main channel flow. By
considering the average velocity in each segment being evaluated, you can
calculate a travel time using the equation:

Li
ti = (2.2-2)
60 vi
where:
ti = Travel time for velocity in segment i, in minutes
Li = Length of the flow path for segment i, in feet
vi = Average velocity for segment i, in feet/second

The time of concentration is calculated as:

t c = t1 + t 2 + t 3 + ... t i (2.2-3)
where:
tc = Time of concentration, in minutes
t 1, t 2, t 3, t i = Travel time in minutes for segments 1, 2, 3, i, respectively

The segments should have uniform characteristics and velocities. Determining


travel time for overland flow, shallow channel flow, and main channel flow are
discussed below.

Chapter 2: Hydrology 2-3


January 2019

Drainage Design Guide


Chapter 2: Hydrology

(A) Overland Flow (t1)


If you know the average slope and the land use, you can determine the time of
concentration for overland flow using Figure B-2 in Appendix B (Hydrology Design
Aids). This chart gives reasonable values and is used by district drainage staff
around the state.

The Federal Highway Administration (FHWA) prefers the Kinematic Wave


Equation developed by Ragan (1971) for calculating the travel time for overland
conditions. Figure B-1 in Appendix B (Hydrology Design Aids) presents a
nomograph that you can use to solve this equation, as follows:

0.93 L0.6 n0.6


t1 = 0.4 0.3 (2.2-4)
i s

where:
t1 = Overland flow travel time, in minutes
L= Overland flow length, in feet (maximum 100 feet recommended by
Natural Resources Conservation Service [NRCS])
n= Manning roughness coefficient for overland flow (See Table B-1 in
Appendix B, Hydrology Design Aids)
i= Rainfall intensity, in inches/hour
S= Average slope of overland flow path, in feet/feet

Manning's n values reported in Table B-1 in Appendix B were determined


specifically for overland flow conditions and are not appropriate for conventional
open channel flow calculations. Equation 2.2-4 generally involves a trial-and-error
process using the following steps:

1. Assume a trial value of rainfall intensity (i).


2. Find the overland travel time (t1) using Figure B-1 (Appendix B).
3. Find the actual rainfall intensity for a storm duration of t1, using the appropriate
IDF curve.
4. Compare the trial and actual rainfall intensities. If they are not similar, select a
new trial rainfall intensity and repeat the process.

Chapter 2: Hydrology 2-4


January 2019

Drainage Design Guide


Chapter 2: Hydrology

(B) Shallow Channel Flow (t2)


Knowing the slope of the flow segment, average velocities for shallow channel flow
(shallow concentrated flow) are obtained from Figure B-3 in Appendix B (Hydrology
Design Aids).

Calculate the velocity using this equation:

V = kS 0.5 (2.2-5)
where:
V= Velocity (feet per second)
S= Longitudinal slope in feet / feet
k= Constant for different flow types. (Refer to table below Figure B-3 in
Appendix B)

You also can calculate gutter flow velocities using the following equation:

1.12 0.5 0.67 0.67


V = S SX T (2.2-6)
n
where:
S= Longitudinal slope
n= Manning’s n for street and pavement gutters (Appendix B, Table B-
2)

SX and T are as shown on (1) below.

For a triangular gutter, Sx and T are as shown in (2) above.

S X 1S X 2
SX = (2.2-7)
SX1 + SX 2

Use the conventional form of Manning’s Equation to evaluate shallow channel


flow.

Chapter 2: Hydrology 2-5


January 2019

Drainage Design Guide


Chapter 2: Hydrology

(C) Main Channel Flow (t3)


Flow in rills, gullies, and/or gutters empties into channels or pipes. Assume that
open channels begin where either a blue line stream shows on USGS quad maps
or where the channel is visible on aerial photos.

Evaluate average velocities for main channel flow using Manning’s Equation.

1.486 0.67 0.5


V = R S (2.2-8)
n

where:
V= Velocity in feet per second
n= Manning’s n value from Table B-3 (Appendix B)
R= Hydraulic Radius (A/P)
S= Longitudinal Slope in feet/feet

2.2.2.2 Kirpich (1940) Equation


You can use the Kirpich Equation for rural areas to estimate the watershed tc directly. The
Kirpich Equation is based on data reported by Ramser (1927) for six small agricultural
watersheds near Jackson, Tennessee. The slope of these watersheds was steep, the
soils well drained, the timber cover ranged from zero percent to 56 percent, and
watershed areas ranged from 1.2 acres to 112 acres. Although these data appear to be
limited and site-specific, the Kirpich Equation has given good results in Florida
applications. The Kirpich Equation is expressed as:

L0.77
t c = 0.0078 Fs (2.2-9)
S 0.385

where:
tc = Time of concentration, in minutes
L= Length of travel, in feet
S= Slope, in feet/feet
Fs = 1.0 for natural basins with well-defined channels, overland flow on bare
earth, and mowed grass roadside channels
= 2.0 for overland flow on grassed surfaces
= 0.4 for overland flow on concrete or asphaltic surfaces
= 0.2 for concrete channels

Chapter 2: Hydrology 2-6


January 2019

Drainage Design Guide


Chapter 2: Hydrology

Separate the flow path into different reaches if there are breaks in the slope and changes
in the topography. Add together the times of travel in each reach to obtain the time of
concentration (see Equation 2.2-3).

2.2.3 Peak Runoff Rates—Ungaged Sites


Synthetic procedures recommended for developing peak flow rates include the rational
equation and USGS regression equations.

2.2.3.1 Rational Equation


The rational equation is an easy method for calculating peak flow rates. The equation is
expressed as:
Q=C i A (2.2-10)

where:
Q= Peak flow rate (cubic feet per second)
C= Runoff coefficient
i= Rainfall intensity (inches per hour)
A= Area (acres)

(A) Runoff Coefficent


The runoff coefficient is a dimensionless number that represents the percent of
rainfall that runs off a site. Table B-4 in Appendix B (Hydrology Design Aids)
presents runoff coefficient ranges for various land uses, soil types, and watershed
slopes. Perform a site review and use your best engineering judgment to select
the coefficient within these ranges. Table B-5 in Appendix B presents adjustment
factors for pervious area runoff coefficients for design storm frequencies greater
than 10 years. (Note: The adjusted runoff coefficient should not be greater than 1.
See Example 2.2-1.) For sites with several land uses, the weighted average of the
runoff coefficient is expressed as:

Σ C i Ai
Weighted C = (2.2-11)
ATotal

(B) Rainfall Intensity


The rainfall intensity is determined from the appropriate IDF curve based on the
time of concentration and the storm frequency (recurrence interval). IDF curves
are available from the Department’s Internet site at:
http://www.fdot.gov/roadway/Drainage/ManualsandHandbooks.shtm

Chapter 2: Hydrology 2-7


January 2019

Drainage Design Guide


Chapter 2: Hydrology

(C) Assumptions and Limitations


1. Rainfall is constant for the duration of the time of concentration.
2. Peak flow occurs when the entire watershed is contributing.
3. Drainage area is limited to those given in the Drainage Manual.

Example 2.2-1: Use of the Rational Method

A flooding problem exists along a farm road near Zolfo Springs in Hardee County (sandy
soil, zone 8). A low water crossing is to be replaced by a culvert to improve the road safety
during rainstorms. The drainage area is shown above and has an area of 108.1 acres.
Determine the maximum flow the culvert must pass for a 25-year storm.

1. Determine the weighted "C," assuming sandy soil. From the sketch and Tables B-
4 and B-5 in Appendix B, develop a summary of "C" values, adjusted for design
storm frequency.

Chapter 2: Hydrology 2-8


January 2019

Drainage Design Guide


Chapter 2: Hydrology

Description "C" Value Adjustment Adjusted C Area Ci Ai


Park 0.20 1.1 0.22 53.9 11.9
Commercial
0.95 N/A 0.95 3.7 3.5
Development
Single Family 0.40 1.1 0.44 50.5 22.2
TOTALS 108.1 37.6

Σ C i Ai 37.60
Weighted C = = = 0.35
A 108.1

2. Determine intensity. To determine the intensity, the time of concentration (tc) must
first be determined.

a. Overland flow (1,100 ft) – "Residential" at 2-percent slope.

From Figure B-2 (Appendix B)


Velocity = 57 ft/min

Distance1 = 1100 ft = 19.3 min .


t1 = Σ
Velocity1 57 ft/ min

b. Channelized flow (2,150 ft) – "Grassed Waterway" at 1-percent slope.

From Figure B-3 (Appendix B)


Velocity = 1.6 ft/sec

c. Time of concentration is estimated as:

tc = t1 + t2 = 19.3 + 22.4 = 41.7 min.

d. Intensity is obtained from the Department’s IDF curves using a duration equal
to the time of concentration (tc). IDF curves are available on the Department’s
Internet Site.

i25 = 4.8 in/hr

Chapter 2: Hydrology 2-9


January 2019

Drainage Design Guide


Chapter 2: Hydrology

3. Calculate the peak flow.

Q25 = C x i25 x A = 0.35 x 4.8 x 108.1 = 182 cubic feet per second

2.2.3.2 Regression Equations


(A) Urban Conditions
You can use regression equations developed by the USGS (Verdi, 2006) to estimate peak
runoff for natural flow conditions.

The USGS equations in “Magnitude and Frequency of Floods for Rural Streams in
Florida, 2006” by Verdi (2006) supersede the information presented by Bridges (1982)
and in the USGS Water Supply Paper (WSP) No. 1674 by Pride (1958). Although not
recommended as a design procedure, you can use the method presented in WSP No.
1674 as an independent check for evaluating natural flow estimates for watershed areas
between 100 and 10,000 square miles.

The Statistical Analysis System (SAS) was used to perform multiple regression analyses
of flood peak data from 275 gagging stations in Florida and 30 in the adjacent states of
Georgia and Alabama. Tables B-10 through B-13 in Appendix B (Hydrology Design Aids)
show the USGS Regression Equations for each designated region in the State of Florida.

The natural flow regression equations for Regions 1 through 4 take the following general
form:

QT = C Aa ( ST + 1.0) b (2.2-12)

where:
QT = Peak runoff rate for return period T, in ft3/sec.
C= Regression constant (See Appendix B, B-10 through B-13)
A= Drainage area in square miles
ST = Basin storage, the percentage of the drainage basin occupied by lakes,
reservoirs, swamps, and wetland. In-channel storage of a temporary nature,
resulting from detention ponds or roadway embankments, is not included in
the computation of ST.
a, b = Regression exponents (See Appendix B, B-10 through B-13).

The standard error of prediction, in percent, is reported for each natural flow regression
equation for each of the Regions 1 through 4, Tables B-10 through B-13 (Appendix B).
The standard error of prediction is a measure of how well the regression equation
estimates flood flows when applied to ungaged basins.

Chapter 2: Hydrology 2-10


January 2019

Drainage Design Guide


Chapter 2: Hydrology

The square of the multiple regression coefficient (R2), unit less, and the standard error, in
percent, are reported for each regression equation for the urban and Tampa Bay area
and Leon County, Tables B-14 through B-16 (Appendix B). The R2 value provides a
measure of the equation’s ability to account for variation in the dependent variable. The
standard error is the standard deviation of the distribution of residuals about the
regression line.

The standard error of model, in percent, is reported for each West-Central Florida
regression equation, Table B-17 (Appendix B). The standard error of model is a measure
of how well the regression equation model estimates flood flows.

When applying the regression equations, you should consider the following limitations:

1. The relationship of the regression equations for areas with basin characteristics
outside the ranges given above. Do not use the equations for watershed conditions
outside the range of applicability shown in Tables B-10 through B-13 in Appendix
B (Hydrology Design Aids).
2. In areas of karst topography for the Tampa Area and Leon County regression
equations, some basins may contain closed depressions and sinkholes, which do
not contribute to direct runoff. When you determine the drainage area from 7.5-
minute topographic maps, subtract any area containing sinkholes or depressions
(non-contributing areas) from the total drainage area.
3. Regression equations are not applicable where manmade changes have a
significant effect on the runoff. These changes may include construction of dams,
reservoirs, levees and diversion canals, strip mines, and areas with significant
urban development.

To apply the USGS regression equations, you should take the following steps:

1. Locate the appropriate region on Figure B-4 (Appendix B).


2. Select the appropriate equations (from Appendix B, Tables B-10 through B-13) for
the region in which your site is located.
3. Determine the input parameters for your selected regression equation.
4. Calculate peak runoff rates for the desired return periods.

(B) Urban Conditions


You can use regression equations developed by the USGS as part of a nationwide project
to estimate peak runoff for urban watershed conditions. Regionalized regression
equations for the Tampa Bay area, Leon County, and West-Central Florida also are
available.

Chapter 2: Hydrology 2-11


January 2019

Drainage Design Guide


Chapter 2: Hydrology

(1) Nationwide Equations

Sauer, et al. (1983), provide two seven-parameter equations and a third set based on
three parameters. The seven-parameter equations based on lake and reservoir
(presented in Appendix B, Table B-14) are recommended. The equations account for
regional runoff variations through the use of the equivalent rural peak runoff rate (RQ).
The equations adjust RQ to an urban condition using the basin development factor (BDF),
the percentage of impervious area (IA), and other variables. These equations have the
following general form:

UQ T = C AB1 SL B 2 (i 2 + 3)B 3 (ST + 8 )B 4 (13 - BDF )B 5 IAB 6 ( RQT )B7 (2.2-13)


where:
UQT = Peak discharge, in cubic feet per second, for the urban watershed for
recurrence interval T
C= Regression constant (See Appendix B, Table B-14)
A= Contributing drainage area in square miles
SL = Channel slope (feet/mile) between points 10 percent and 85 percent of the
distance from the design point to the watershed boundary
i2 = Rainfall intensity, in inches, for the two-hour, two-year occurrence
ST = Basin storage, the percentage of the drainage basin occupied by lakes,
reservoirs, swamps, and wetland. In-channel storage of a temporary nature,
resulting from detention ponds or roadway embankments, is not included in
the computation of ST.
BDF = Basin development factor is an index of the prevalence of (1) channel
improvements, (2) impervious channel linings, (3) storm drains, and (4) curb
and gutter streets and ranges from 0 to 12. More discussion and an example
follow these definitions.
IA = Impervious area is the percentage of the drainage basin occupied by
impervious surfaces, such as buildings, parking lots, and streets.
RQT = Peak discharge, in cubic feet per second, for an equivalent rural drainage
basin in the same hydrologic area as the urban basin for recurrence interval
T. This value is developed using the USGS regression equations for natural
flow conditions for the appropriate region.
B1 to B7 = Regression exponents (See Appendix B, Table B-14)

Basin Development Factor—Determine the BDF from drainage maps and by field
inspection of the watershed. First, divide the basin into three sections so that each sub-
area contains approximately one-third of the drainage area. Mark distances along main
streams and tributaries so that, within each third, the travel distances of two or more
streams are about equal. Generally, you can draw the lines on the drainage map by visual
estimate without the need for measurements. Complex basin shapes and drainage
patterns require more judgment when subdividing.

Chapter 2: Hydrology 2-12


January 2019

Drainage Design Guide


Chapter 2: Hydrology

You will examine four drainage aspects for each subsection, assigning a code of zero or
one to each aspect for each subsection. The BDF, therefore, can range from zero for an
undeveloped watershed to 12 for a completely urbanized watershed. A code of zero does
not mean that the watershed is completely unaffected by urbanization. A basin could have
some impervious area, some improved channels and some curb and gutter streets and
still have a BDF of zero. The four drainage aspects are:

1. Channel Improvements—If 50 percent or more of the main channels and


principal tributaries (those that drain directly into the main channel) have
been improved from natural conditions, assign a code of one; otherwise,
assign a code of zero. Improvements include straightening, enlarging,
deepening, and clearing.
2. Channel Linings—Assign a code of one if more than 50 percent of the length
of the main channels and principal tributaries have impervious linings, such
as concrete; otherwise, assign a code of zero. Lined channels are an
indication of a more developed drainage system in which channels probably
have been improved.
3. Storm Drains—Storm drains are enclosed drainage structures (usually pipes)
frequently used on the secondary tributaries (those that drain into principal
tributaries) that receive drainage directly from streets or parking lots. Many
of these drains empty into open channels; in some basins, however, they
empty into channels enclosed as box or pipe culverts. When more than 50
percent of the secondary tributaries within a sub-basin consist of storm
drains, assign a code of one to this aspect; otherwise, assign a code of zero.
Note that if 50 percent or more of the main drainage channels and principal
tributaries are enclosed, you also would assign the aspects of channel
improvements and channel linings a code of one.
4. Curb and Gutter Streets—If more than 50 percent of a sub-basin is urbanized
(covered by residential, commercial, or industrial development), and if more
than 50 percent of the streets and highways in the sub-basin are constructed
with curbs and gutters, then assign a code of one to this aspect; otherwise,
assign a code of zero. Drainage from curb and gutter streets frequently
empties into storm drains.

These guidelines are not intended to be precise measurements. A certain amount of


subjectivity will be involved, and you should perform field checks to obtain the best
estimate.

Example 2.2-2: Estimating the BDF

A watershed is divided into three sub-areas based on homogeneity of hydrologic


conditions. Information for the watershed is collected from topographic maps and field
reviews and is tabulated below:

Chapter 2: Hydrology 2-13


January 2019

Drainage Design Guide


Chapter 2: Hydrology

Main Length of Length of Length of Length of Length of


Road
channel secondary channel channel storm curb &
Subarea length
length tributaries improved lined drains gutter
(ft)
(ft) (ft) (ft) (ft) (ft) (ft)

Upper 2500 5180 2850 460 0 1345 690

Middle 3800 3940 4700 2020 1770 2330 3020

Lower 3000 2160 5610 1720 1570 1510 3180

The BDF is determined as follows:

Channel Improvements
Upper third: 460 ft have been straightened and deepened
460/2,500 < 50% Code = 0
Middle third: 2,020 ft have been straightened and deepened
2,020/3,800 > 50% Code = 1
Lower third: 1,720 ft have been straightened and deepened
1,720/3,000 > 50% Code = 1
Channel Linings
Upper third: 0 ft have been lined
0/2,500 < 50% Code = 0
Middle third: 1,770 ft have been lined
1,770/3,800 < 50% Code = 0
Lower third: 1,570 ft have been lined
1,570/3,000 > 50% Code = 1
Storm Drains on Secondary Tributaries
Upper third: 1,345 ft have been converted to storm drains
1,345/5,180 < 50% Code = 0
Middle third: 2,330 ft have been converted to storm drains
2,330/3,940 > 50% Code = 1
Lower third: 1,510 ft have been converted to storm drains
1,510/2,160 > 50% Code = 1
Curb and Gutter Streets
Upper third: 690 ft of curb and gutter street
690/2,850 < 50% Code = 0
Middle third: 3,020 ft of curb and gutter street
3,020/4,700 > 50% Code = 1
Lower third: 3,180 ft of curb and gutter street
3,180/5,610 > 50% Code = 1
Total BDF = 7

Chapter 2: Hydrology 2-14


January 2019

Drainage Design Guide


Chapter 2: Hydrology

(2) Tampa Bay Area, Leon County, West-Central Florida:

You can use regression equations developed as part of a nationwide project by the USGS
(Sauer et al., 1983) to estimate peak runoff for urban watershed conditions. Regionalized
regression equations for urban watersheds in the Tampa Bay area and for Leon County
are presented by Lopez and Woodham (1983), Franklin and Losey (1984), and Hammett
and DelCharco (2001) respectively. Tables B-15, B-16, and B-17 in Appendix B show the
USGS Regionalized Regression Equations for the Tampa Bay area, Leon County, and
West-Central Florida respectively.

(a) Tampa Bay Area

For urban drainage areas of less than 10 square miles in the Tampa Bay area, the general
form of the regression equations are:

For 2-, 5-, and 10-year frequencies:


QT = C AB1 BDF B2 SLB3 (DTENA + 0.01 )B4 (2.2-14)

For 25-, 50-, and 100-year frequencies:


QT = C AB1( 13 − BDF ) B2 SLB 3 (2.2-15)

where:
QT = Peak runoff rate for return period T, in cubic feet per second
C= Regression constant (See Appendix B, Table B-15)
A= Drainage area in square miles
BDF = Basin development factor (dimensionless)
SL = Channel slope (feet/mile) between points 10 percent and 85 percent of the
distance from the design point to the watershed boundary.
DTENA = Surface area of lakes, ponds, and detention and retention basins expressed
as a percent of drainage area.
B1, B2, etc. = Regression exponents (See Appendix B, Table B-15)

The equations are not to be used for watershed conditions outside the range of
applicability shown in Table B-15 (Appendix B). To apply the Tampa Bay regression
equations:

1. Determine input parameters, including drainage area, basin development


factor (see Example 2.2-2), channel slope, and the surface area of lakes,
ponds, etc.

Chapter 2: Hydrology 2-15


January 2019

Drainage Design Guide


Chapter 2: Hydrology

2. Calculate peak runoff rates for the desired return periods.

(b) Leon County

For urban drainage areas of less than 16 square miles in Leon County, Franklin and Losey
(1984) developed regression equations for areas inside and outside the Lake Lafayette
Basin.

The general form of both sets of equations is:

QT = C AB1 IAB2 (2.2-16)

where:
QT = Peak runoff rate for return period T, in cubic feet per second
C= Regression constant (See Appendix B, Table B-16)
A= Drainage area in square miles
IA = Impervious area, in percent of drainage area
B1, B2 = Regression exponents (See Appendix B, Table B-16)

These equations must not be used for watershed conditions outside the range of
applicability shown in Table B-16 (Appendix B). The following steps are used to apply
the Leon County regression equations:

1) Determine input parameters, including drainage area and impervious area.


2) Select the appropriate equations from Table B-16 (Appendix B), depending
on whether the area is inside or outside the Lake Lafayette Basin.
3) Calculate peak runoff rates for the desired return periods using the
equations in Table B-16 (Appendix B).

(c) West-Central Florida

For drainage areas in West-Central Florida, Hammett and DelCharco (2001) developed
regression equations for areas inside and outside the Southwest Florida Water
Management District. The general form of the regression equations are:

For Region 1:
QT = C AB1 ( LK + 0.6) B2 (2.2-17)

Chapter 2: Hydrology 2-16


January 2019

Drainage Design Guide


Chapter 2: Hydrology

For Regions 2 through 4:

QT = C AB1 ( LK + 3.0) B2 SLB3 (2.2-18)

where:
QT = Peak runoff rate for return period T, in cubic feet per second
C= Regression constant (See Appendix B, Table B-17)
A= Drainage area in square miles
LK = Drainage area covered by lakes, in percent of drainage area
SL = Channel slope (feet/mile) between points 10 percent and 85 percent of the
distance from the design point to the watershed boundary
B1, B2, B3 = Regression exponents (See Appendix B, Table B-17)

These equations must not be used for watershed conditions outside the range of
applicability shown in Table B-18 (Appendix B). The following steps are used to apply the
West-Central Florida regression equations:

1) Locate the appropriate region on Figure B-5 (Appendix B).


2) Select the appropriate equations (from Appendix B, Table B-17) for the
region in which your site is located.
3) Determine the input parameters for your selected regression equation.
4) Calculate peak runoff rates for the desired return periods.

(3) Water Management District and Local Drainage District Procedures

Some Water Management Districts (WMDs) in Florida set allowable discharge or removal
rates for specific watershed areas. WMDs also may have computer programs for surface
hydrology calculations available. Consult the appropriate WMD handbook and, if needed,
appropriate WMD or FDOT District drainage personnel for guidance. There are also local
drainage districts that control runoff amounts to particular streams or water bodies.

2.2.4 Flood Hydrographs


Because you will not be able to obtain observed data for deriving unit hydrograph
parameters in most cases, you will often need to use synthetic procedures. The two flood
hydrograph procedures that you can perform are the modified rational method and the
NRCS unit hydrograph. The Department’s rainfall distributions are included with the IDF
curves, which are available from the Department’s Internet site. Each Water Management
District has rainfall distributions appropriate for their respective regions.

Chapter 2: Hydrology 2-17


January 2019

Drainage Design Guide


Chapter 2: Hydrology

2.2.4.1 Modified Rational Method


Because of the assumptions and limitations of the rational method (see Section 2.2.3),
use of the modified rational method for flood hydrograph procedures is limited to small
basins having a time of concentration of 15 minutes or less. (See the Drainage Manual,
Section 5.4.2.)

Example: Using a drainage area of 0.981 acres, tc of 10 minutes, CA of 0.82, and IDF
Zone 5, calculate an inflow hydrograph for the 100-year two-hour rainfall.

From the Zone 5 IDF curves, the two-hour 100-year i = 2.7 inches/hour.
Therefore, Ptotal = 5.4 inches

(1) (2) (3) (4)


Time (hours) i/P total i (in/hr) Q (cfs)
0.2 0.50 2.70 2.21
0.4 0.75 4.05 3.31
0.6 1.00 5.40 4.41
0.8 1.25 6.75 5.51
1.0 0.50 2.70 2.21
1.2 0.30 1.62 1.32
1.4 0.25 1.35 1.10
1.6 0.20 1.08 0.88
1.8 0.15 0.81 0.66
2.0 0.00 0.00 0

Columns 1 & 2 are from the rainfall distribution curves


Column 3 is Column 2 times Ptotal
Column 4 is Column 3 times CA (0.82 for this example)

2.2.4.2 NRCS Hydrograph


Techniques developed by the NRCS, formerly the Soil Conservation Service (SCS), for
calculating rates of runoff require the same basic data as the Rational Method: drainage
area, a runoff factor, time of concentration, and rainfall. The NRCS approach also
considers the time distribution of the rainfall, initial losses to interception and depression
storage, and infiltration that decreases during the storm. Since NRCS hydrographs are
calculated using computers, the discussion in this guide will address the basic concepts
rather than computation methods.

Chapter 2: Hydrology 2-18


January 2019

Drainage Design Guide


Chapter 2: Hydrology

(A) Time of Concentration


Calculate the time of concentration using any of the methods in Section 2.2.2.

(B) Curve Number


The NRCS developed an empirical relationship for estimating rainfall excess that
accounts for infiltration losses and initial abstractions by using a site-specific runoff
parameter called the curve number (CN). The watershed CN is a dimensionless
coefficient that reflects watershed cover conditions, hydrologic soil group, land uses, and
antecedent moisture conditions.

Three levels of antecedent moisture conditions are considered by the NRCS relationship.
Antecedent Moisture Condition I (AMC-I) is the lower limit of antecedent rainfall or the
upper limit of the maximum soil storage (S). Antecedent Moisture Condition II (AMC-II)
represents average antecedent rainfall conditions, and Antecedent Moisture Condition III
(AMC-III) is the upper limit of antecedent rainfall or the lower limit of S. Only AMC-II
generally is selected for design purposes. The curve number values in the tables in the
Appendix B (Hydrology Design Aids) are based on AMC-II.

To determine the curve number:


1. Identify soil types using the appropriate county soil survey report.
2. Assign a hydrologic group (A, B, C, or D) to each soil type. (See Appendix B, Table
B-6.) In general:
A= deep sand, deep loess, aggregated silts
B= shallow loess, sandy loam
C= clay loams, shallow sandy loam, soils low in organic content, soils usually
high in clay
D= soils that swell significantly, heavy plastic clays, some saline soils
3. Identify drainage areas with uniform soil type and land use conditions.
4. Use tables B-7 through B-9 (Appendix B) or other references to select curve
number values for each uniform drainage area identified in Step 3.
5. Calculate a composite curve number using the equation:

Σ CN i Ai
CN C = (2.2-19)
AT

where:
CNC = Composite curve number
CNi = Curve number for sub-area I
Ai = Area for sub-area I
AT = Total area of watershed

Chapter 2: Hydrology 2-19


January 2019

Drainage Design Guide


Chapter 2: Hydrology

The curve number tables developed by the U.S. Department of Agriculture (USDA) are
based on the assumption that all impervious areas have a CN of 98 and are hydraulically
connected. If the rain on the roof of a house runs off onto the lawn, that roof area is not
hydraulically connected. If the roof drains into a gutter, which in turn flows onto the
driveway, then on to the street, that area is hydraulically connected.

If these assumptions don't fit the project area, there is an alternate method of predicting
curve number from Department-sponsored research on estimating coefficients for
hydrologic methods used for the design of hydraulic structures. The results were reported
in "Techniques for Estimating Hydrologic Parameters for Small Basins in Florida," by
Scott Kenner, et al, FDOT Project Number 99700-3542, April 1996.

The resulting equation for estimating the CN is:

CN = 58.38 - 8.2716 ln(A) + 0.50274 HCIA + 6.22971 ln(L) + 0.68079 ln( Lc ) - 0.14986 S
(2.2-20)
where:
A= Drainage area (acres)
HCIA = Hydraulically connected impervious area (percent of A)
L= Length of main flow channel (feet)
Lc = Length to centroid (feet)
S= Main channel slope (feet/mile)

(C) Rainfall-Runoff Relationship


The maximum soil storage and the CN value for a watershed are related by the following
expression:

1000
S= - 10 (2.2-21)
CN

where:
S= Maximum soil storage, in inches
CN = Watershed curve number, dimensionless

Chapter 2: Hydrology 2-20


January 2019

Drainage Design Guide


Chapter 2: Hydrology

When you know the maximum soil storage, you can calculate the rainfall excess using
the following NRCS relationship:
(P - 0.2S )2
R= (2.2-22)
P + 0.8S
where:
R= Accumulated rainfall excess (or runoff), in inches
P= Accumulated rainfall, in inches
S= Maximum soil storage, in inches

Additional information on the NRCS relationship is available in USDA, NRCS publications


TP-149 (1973) and NEH-4 (1972).

(D) Shape Factor


The hydrograph shape factor (B) generally is considered to be a constant characteristic
of a watershed. The NRCS dimensionless unit hydrographs are based on a B value of
484. However, since the value of B generally ranges from 600 in steep terrain to 300 or
less in flat swampy areas, you may need to make adjustments to the unit hydrograph
shape. You can make these adjustments by changing the percent of area under the rising
and recession limbs of the unit hydrograph to reflect the corresponding change in the
hydrograph shape factor. The B value of 484 reflects a hydrograph that has ⅜ of its area
under the rising limb. For mountainous terrain, a larger percentage of the area would
probably be under the rising limb, represented by a larger B value.

The South Florida Water Management District has a memorandum (dated June 25, 1993)
concerning hydrograph shape (peak rate) factors. For slopes less than 5 feet per mile, a
factor of 100 is recommended, and for slopes in South Florida greater than 5 feet per
mile, a factor of 256 is recommended.

Hal Wilkening of the St. Johns River Water Management District prepared a
memorandum for a "Procedure for Selection of SCS Peak Rate Factors for Use in MSSE
Permit Applications", dated April 25, 1990. The memorandum provides a summary of the
NRCS unit hydrograph methodology and information on research on, as well as
recommendations for the selection of, hydrograph shape (peak rate) factors. His
recommendations are outlined in the following table.

Chapter 2: Hydrology 2-21


January 2019

Drainage Design Guide


Chapter 2: Hydrology

Site Conditions Shape Factor


Represents watersheds with very mild slopes, recommended by
NRCS for watersheds with average slope of 0.5 percent or less.
Significant surface storage throughout the watershed. Limited onsite
drainage ditches. Typical ecological communities include: North 256 to 284
Florida flat woods, South Florida flat woods, freshwater marsh and
ponds, swamp hardwoods, cabbage palm flatlands, cypress swamp,
and similar vegetative communities.
Intermediate peak rate factor representing watersheds with moderate
surface storage in some locations due to depression areas, mild
slopes, and/or lack of existing drainage features. Typical ecological 323 to 384
communities include: oak hammock, upland hardwood hammock,
mixed hardwood and pine, and similar vegetative communities.
Standard peak rate factor developed for watersheds with little or no
storage. Represents watersheds with moderate to steep slopes
and/or significant drainage works. Typical ecological communities 484
include: long leaf pine, turkey oak hills, and similar vegetative
communities.

The Department sponsored research on estimating coefficients for hydrologic methods


used for the design of hydraulic structures. The results were reported in "Techniques for
Estimating Hydrologic Parameters for Small Basins in Florida," by Scott Kenner, et al.,
FDOT Project Number 99700-3542, April 1996. The resulting equation for estimating
the NRCS shape factor is:

B = exp[390 − 0.01396 A − 0.00473HCIA + 0.00064 L − 0.00053LC + 0.00567 S ]

(2.2-23)
where:
A= Drainage area (acres)
HCIA = Hydraulically connected impervious area (percent)
L= Length of main flow channel (feet)
Lc = Length to centroid (feet)
S= Main channel slope (feet/mile)

The designer should consult with district drainage personnel and, if necessary, WMD
personnel before using a shape (peak rate) factor other than the standard factor of 484.

Chapter 2: Hydrology 2-22


January 2019

Drainage Design Guide


Chapter 3: Open Channel

CHAPTER 3: OPEN CHANNEL

3. OPEN CHANNEL...................................................................................................3-1

3.1 Open Channel Flow Theory.................................................................................3-1


3.1.1 Mass, Energy, and Momentum .......................................................................3-1
3.1.1.1 Mass ........................................................................................................3-1
3.1.1.2 Energy......................................................................................................3-1
3.1.1.3 Momentum ...............................................................................................3-4
3.1.2 Uniform Flow ...................................................................................................3-5
3.1.2.1 Manning’s Equation .................................................................................3-5
3.1.3 Critical Flow...................................................................................................3-10
3.1.3.1 Specific Energy and Critical Depth ........................................................3-10
3.1.3.2 Critical Velocity ......................................................................................3-14
3.1.3.3 Super-Critical Flow.................................................................................3-14
3.1.3.4 Sub-Critical Flow....................................................................................3-14
3.1.3.5 Theoretical Considerations ....................................................................3-15
3.1.4 Non-Uniform Flow .........................................................................................3-15
3.1.4.1 Gradually Varied Flow............................................................................3-17
3.1.4.2 Gradually Varied Flow Profile Computation ...........................................3-17
3.1.4.3 Rapidly Varied Flow ...............................................................................3-24
3.1.5 Channel Bends..............................................................................................3-27

3.2 Open Channel Design ........................................................................................3-29


3.2.1 Types of Open Channels for Highways.........................................................3-29
3.2.2 Roadside Ditches ..........................................................................................3-31
3.2.3 Median Ditches .............................................................................................3-35
3.2.4 Interceptor Ditches ........................................................................................3-36
3.2.5 Outfall Ditches...............................................................................................3-37
3.2.6 Hydrology ......................................................................................................3-37
3.2.6.1 Frequency ..............................................................................................3-38
3.2.6.2 Time of Concentration............................................................................3-38
3.2.7 Tailwater and Backwater...............................................................................3-39
3.2.8 Side Drains....................................................................................................3-47
3.2.8.1 Design Analysis Requirements for Side Drains .....................................3-47
3.2.8.2 Material Requirements...........................................................................3-48
3.2.8.3 End Treatment .......................................................................................3-48

3.3 Channel Linings .................................................................................................3-52


3.3.1 Flexible Linings .............................................................................................3-52
3.3.1.1 Vegetation..............................................................................................3-53
3.3.1.2 Other Flexible Linings ............................................................................3-53
3.3.2 Rigid Linings..................................................................................................3-55

Chapter 3: Open Channel 3-i


January 2019

Drainage Design Guide


Chapter 3: Open Channel

3.3.2.1 Cast-in-Place Concrete..........................................................................3-56


3.3.2.2 Grout-Filled Mattresses..........................................................................3-56
3.3.3 Velocity and Shear Stress Limitations...........................................................3-57
3.3.4 Application Guidance for Some Common Channel Linings ..........................3-59
3.3.4.1 Rubble Riprap ........................................................................................3-59
3.3.4.2 Grout-Filled Mattresses..........................................................................3-61
3.3.4.3 Gabions..................................................................................................3-62
3.3.4.4 Soil Stabilizers .......................................................................................3-66

3.4 Drainage Connection Permitting and Maintenance Concerns.......................3-68


3.4.1 Drainage Connection Permitting ...................................................................3-68
3.4.1.1 Roadside Ditch Impacts .........................................................................3-68
3.4.1.2 Median Ditch Impacts ............................................................................3-70
3.4.1.3 Outfall Ditch Impacts..............................................................................3-72
3.4.2 Maintenance Concerns .................................................................................3-72
3.4.2.1 Ditch Closures........................................................................................3-72
3.4.2.2 Acquisition of Ditches from Local Ownership.........................................3-78
3.4.2.3 Addition of Sidewalks to Roadway Projects...........................................3-78

Chapter 3: Open Channel 3-ii


January 2019

Drainage Design Guide


Chapter 3: Open Channel

3. OPEN CHANNEL

3.1 OPEN CHANNEL FLOW THEORY


3.1.1 Mass, Energy, and Momentum
The three basic principles that generally apply to flow analysis, including open channel
flow evaluations, are:

 Conservation of mass
 Conservation of energy
 Conservation of linear momentum

3.1.1.1 Mass
You can mathematically express the conservation of mass for continuous steady flow in
the Continuity Equation as:

Q  v A (3.1-1)

where:
Q= Discharge, in cubic feet per second
A= Cross-sectional area, in square feet
v= Average channel velocity, in feet per second

For continuous unsteady flow, the Continuity Equation must include time as a variable.
You can find additional information on unsteady flow from Chow (1959) or Henderson
(1966).

3.1.1.2 Energy
The total energy head at a point in an open channel is the sum of the potential and kinetic
energy of the flowing water. The potential energy is represented by the elevation of the
water surface. The water surface elevation is the depth of flow, d, defined in Section 1.4,
added to the elevation of the channel bottom, z. The water surface elevation is a measure
of the potential work that the flow can do as it transitions to a lower elevation. The kinetic
energy is the energy of motion as measured by the velocity, v.

If you insert a straight tube down into the flow, the water level in the tube will rise to the
water surface elevation in the channel. If you insert a tube with a 90-degree elbow into
the flow with the open end pointing into the flow, then the water level will rise to a level
higher than the water surface elevation in the channel—this distance is a measure of the
ability of the water velocity to do work. Using Newton’s Laws of Motion, this distance is

Chapter 3: Open Channel 3-1


January 2019

Drainage Design Guide


Chapter 3: Open Channel

v2/2g, where g is the acceleration due to gravity. Therefore, the total energy head at a
point in an open channel is: d + z + v2/2g.

As water flows down a channel, the flow loses energy because of friction and turbulence.
You can set the total energy head between two points in a channel reach equal to one
another if the losses between the sections are added to the downstream total energy
head. This equality is commonly known as the Energy Equation, which is expressed as:

2 2
v1 v
d1   z1 d 2 2  z 2  hloss (3.1-2)
2g 2g

where:
d1, d2 = Depth of open channel flow at channel sections 1 and 2, respectively, in
feet
v1, v2 = Average channel velocities at channel sections 1 and 2, respectively, in
feet per second
z1, z2 = Channel elevations above an arbitrary datum at channel sections 1 and 2,
respectively, in feet
hloss = Head or energy loss between channel sections 1 and 2, in feet
g= Acceleration due to gravity, 32.174 ft/sec2

A longitudinal profile of total energy head elevations is called the energy grade line
(gradient). The longitudinal profile of water surface elevations is called the hydraulic grade
line (gradient). The energy and hydraulic grade lines for uniform open channel flow are
illustrated in Figure 3.1-1. For flow to occur in an open channel, the energy grade line
must have a negative slope in the direction of flow. A gradual decrease in the energy
grade line for a given length of channel represents the loss of energy caused by friction.
When considered together, the hydraulic and energy grade lines reflect not only the loss
of energy by friction, but also the conversion between potential and kinetic forms of
energy.

Chapter 3: Open Channel 3-2


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Figure 3.1-1: Characteristics of Uniform Open Channel Flow

Figure 3.1-2: Definition Sketch for Specific Head and Sub-Critical and Super-
Critical Flow

Chapter 3: Open Channel 3-3


January 2019

Drainage Design Guide


Chapter 3: Open Channel

For uniform flow conditions, the energy grade line is parallel to the hydraulic grade line,
which is parallel to the channel bottom (see Figure 3.1-1). Thus, for uniform flow, the
slope of the channel bottom becomes an adequate basis for the determination of
friction losses. During uniform flow, no conversions occur between kinetic and potential
forms of energy. If the flow is accelerating, the hydraulic grade line would be steeper than
the energy grade line, while decelerating flow would produce an energy grade line steeper
than the hydraulic grade line.

The Energy Equation presented in Equation 3.1-2 ignores the effect of a non-uniform
velocity distribution on the computed velocity head. The actual distribution of velocities
over a channel section are non-uniform (i.e., slow along the bottom and faster in the
middle). The velocity head for actual flow conditions generally is greater than the value
computed using the average channel velocity. Find guidance on kinetic energy
coefficients that account for non-uniform velocity conditions in Chapter 5 (Bridge
Hydraulics).

For typical prismatic channels with a fairly straight alignment, the effect of disregarding
the existence of a non-uniform velocity distribution is negligible, especially when
compared to other uncertainties involved in such calculations. Therefore, Equation 3.1-2
is appropriate for most open channel problems. However, if you know or suspect that
velocity distributions are non-typical, you should obtain additional information related to
velocity coefficients, as presented by Chow (1959) or Henderson (1966).

Equation 3.1-2 also assumes that the hydrostatic law of pressure distribution is
applicable. This law states that the distribution of pressure over the channel cross section
is the same as the distribution of hydrostatic pressure; that is, that the distribution is linear
with depth. The assumption of a hydrostatic pressure distribution for flowing water is valid
only if the flow is not accelerating or decelerating in the plane of the cross section. Thus,
restrict the use of Equation 3.1-2 to conditions of uniform or gradually varied non-uniform
flow. If you know that the flow will be varying rapidly, obtain additional information, as
presented by Chow (1959) or Henderson (1966).

3.1.1.3 Momentum
According to Newton's Second Law of Motion, the change of momentum per unit of time
is equal to all the resultant external forces applied to the moving body. Applying this
principle to open channel flow produces a relationship that is virtually the same as the
Energy Equation expressed in Equation 3.1-2. Theoretically, these principles of energy
and momentum are unique, primarily because energy is a scalar quantity (magnitude
only), while momentum is a vector quantity (magnitude and direction). In addition, the
head loss determined by the Energy Equation measures the internal energy dissipated in
a particular channel reach, while the Momentum Equation measures the losses due to
external forces exerted on the water by the walls of the channel. However, for uniform

Chapter 3: Open Channel 3-4


January 2019

Drainage Design Guide


Chapter 3: Open Channel

flow, since the losses due to external forces and internal energy dissipation are equal, the
Momentum and Energy Equations give the same results.

Applying the momentum principle has certain advantages for problems involving
substantial changes of internal energy, such as a hydraulic jump. Thus, you should use
the momentum principle for evaluating rapidly varied non-uniform flow conditions.
Theoretical details of the momentum principle applied to open channel flow are presented
by Chow (1959) and Henderson (1966). Section 3.1.4.3 provides a brief presentation of
hydraulic jump fundamentals.

3.1.2 Uniform Flow


Although steady uniform flow is rare in drainage facilities, it is practical in many cases to
assume that steady uniform flow occurs in appropriate segments of an open channel
system. The results obtained from calculations based on this assumption will be
approximate and general, but still can provide satisfactory solutions for many practical
problems.

3.1.2.1 Manning’s Equation


Determine the hydraulic capacity of an open channel by applying Manning's Equation,
which determines the average velocity when given the depth of flow in a uniform channel
cross section. Given the velocity, calculate the capacity (Q) as the product of velocity and
cross-sectional area (see Equation 3.1-1).

Manning's Equation is an empirical equation with values of constants and exponents


derived from experimental data of turbulent flow conditions. According to Manning's
Equation, the mean velocity of flow is a function of the channel roughness, the hydraulic
radius, and the slope of the energy gradient. As noted previously, for uniform flow,
assume that the slope of the energy gradient is equal to the channel bottom slope.
Manning's Equation is expressed mathematically as follows:

1.486 2 3 12
v R S (3.1-3)
n
or

1.486 2 1
Q AR 3 S 2 (3.1-4)
n

where:
v= Average channel velocity, in feet per second
Q= Discharge, in cubic feet per second
n= Manning's roughness coefficient

Chapter 3: Open Channel 3-5


January 2019

Drainage Design Guide


Chapter 3: Open Channel

𝐴
R= Hydraulic radius of the channel, in feet, calculated: 𝑅 = 𝑃
P= Wetted perimeter of channel, in feet
S= Slope of the energy gradient, in feet per feet
A= Cross-sectional area of the open channel, in square feet

You can find design values for Manning’s roughness coefficient for artificial channels (i.e.,
roadside, median, interceptor, and outfall ditches) in Chapter 2 (Section 2.7) of the
Drainage Manual. You can find guidance on methods for estimate Manning’s roughness
coefficient for natural channels in Chapter 5 (Bridge Hydraulics).

Example 3.1-1—Discharge given Normal Depth

Given: Depth = 0.6 ft


Longitudinal Slope = 0.005 ft/ft
Trapezoidal Cross Section shown below
Manning’s Roughness = 0.06

Calculate: Discharge, assuming normal depth

1 0.6 ft 1

4 6
5 ft

*Not to scale

Note: To make things easier, try breaking the drawing into three parts: two triangles and
a rectangle.

Step 1: Calculate Wetted Perimeter and Cross-Sectional Area


Wetted Perimeter (P):
Solve for the left triangle’s hypotenuse
x  0.6 2  (4  0.6) 2
x = 2.474 ft
Solve for the right triangle’s hypotenuse
x  0.6 2  (6  0.6) 2
x = 3.650 ft
Wetted Perimeter (P) = 2.474 + 3.650 + 5 = 11.124 ft
Cross-Sectional Area (A):
Solve for the left triangle’s area

Chapter 3: Open Channel 3-6


January 2019

Drainage Design Guide


Chapter 3: Open Channel

1
A1  (4  0.6)(0.6)
2
A1  0.72 ft2
Solve for the right triangle’s area
1
A2  (6  0.6)(0.6)
2
A2  1.08 ft2
Solve for the rectangle’s area
A3  5  0.6
A3  3 ft 2
Cross-Sectional Area (A) = 0.72  1.08  3  4.8 ft2

Step 2: Calculate Hydraulic Radius


A
Hydraulic Radius (R) =
P
4.8
Hydraulic Radius (R) =  0.4315 ft
11.124

Step 3: Calculate Average Velocity


1.486 2 1
Average Velocity (v) = ( R) 3 ( S ) 2
n
1.486 2 1
Average Velocity (v) = (0.4315) 3 (0.005) 2  1.00 ft/sec
0.06

Step 4: Calculate the Discharge


Discharge (Q) = v  A
Discharge (Q) = 1.00 ft/sec  4.8 ft 2  4.80 ft 3 /sec

As an alternative approach, Example C.1 of Appendix C solves this example problem


using equations from Figure C-4.

Example 3.1-1 has a direct solution because the depth is known. The next problem will
be more difficult to solve because the discharge will be given and the normal depth must
be calculated. The equations cannot be solved directly for depth, so an iterative process
is used to solve for normal depth. You also can solve Example 3.1-1 using the charts in
Appendix C.

Chapter 3: Open Channel 3-7


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Example 3.1-2—Normal Depth given Discharge


Given:
Discharge = 9 ft 3 /sec

Use the channel cross section shape, slope, and Manning’s roughness coefficient given
in Example 3.1-1

Calculate:
Normal Depth

1 1

4 6
5 ft

Note: The solution must use trial and error since you cannot solve the equations implicitly
for depth. Perform the first trial in the steps below and the remaining trials will be shown
in a table. The initial trial depth (i.e., the first guess) should be greater than the depth
given previously in Example 3.1-1 because the discharge is greater. So we will perform
our trial with an estimated depth of flow of 0.8 ft.

Step 1: Calculate Wetted Perimeter and Cross-Sectional Area


Wetted Perimeter (P):
Solve for the left triangle’s hypotenuse
x  0.8 2  (4  0.8) 2
x  3.298 ft
Solve for the right triangle’s hypotenuse
x  0.8 2  (6  0.8) 2
x  4.866 ft
Wetted Perimeter (P) = 3.298  4.866  5  13.164 ft

Cross-sectional Area (A):


Solve for the left triangle’s area
1
A1  (4  0.8)(0.8)
2
A1  1.28 ft 2
Solve for the right triangle’s area
1
A2  (6  0.8)(0.8)
2

Chapter 3: Open Channel 3-8


January 2019

Drainage Design Guide


Chapter 3: Open Channel

A2  1.92 ft 2
Solve for the rectangle’s area
A3  5  0.8
A3  4 ft 2
Cross-Sectional Area (A) = 1.28  1.92  4  7.2 ft 2

Step 2: Calculate Hydraulic Radius


A
Hydraulic Radius (R) =
P
7.2
Hydraulic Radius (R) =  .547 ft
13.164

Step 3: Calculate Average Velocity


1.486 2 1
Average Velocity (v) = ( R) 3 ( S ) 2
n
1.486 2 1
Average Velocity (v) = (0.547) 3 (0.005) 2  1.171 ft/sec
0.06

Step 4: Calculate the Discharge


Discharge (Q) = v  A
Discharge (Q) = 1.171 ft/sec  7.20 ft 2  8.43 ft 3 /sec

The discharge calculated in Step 4 is still less than 9 ft3/sec, so normal depth is greater
than 0.8 feet. Use a slightly higher depth of flow for the next guess. The following table
summarizes subsequent trials. The trial-and-error process continues until you achieve the
ideal level of accuracy.

Depth (ft) Area Perimeter Radius Velocity Discharge


0.8 7.2 13.16469 0.546917 1.171 8.433
0.85 7.8625 13.67499 0.574955 1.211 9.521
0.82 7.462 13.36881 0.558165 1.187 8.859
0.826 7.54138 13.43005 0.56153 1.192 8.989

The normal depth for the given channel and flow rate is 0.83 feet. You should perform
intermediate calculations using more significant digits than needed, and then round in the
last step to avoid rounding errors.

The Drainage Manual recommends that, where the flow depth is greater than 0.7 feet,
reduce the roughness value to 0.042. However, the normal depth using n = 0.042 is 0.69

Chapter 3: Open Channel 3-9


January 2019

Drainage Design Guide


Chapter 3: Open Channel

feet. The recommended roughness for flow depths less than 0.7 feet is 0.06. The abrupt
change in the recommended roughness values causes this anomaly. If the flow depth is
the primary concern, then using n = 0.06 will give a conservative answer. However, if the
velocity is the primary concern, then using n = 0.042 is conservative.

3.1.3 Critical Flow


The energy content of flowing water with respect to the channel bottom often is referred
to as the specific energy head, which is expressed by the equation:

v2
Ed (3.1-5)
2g

where:
E= Specific energy head, in feet
d= Depth of open channel flow, in feet
v= Average channel velocity, in feet per second
g= Acceleration due to gravity, 32.174 ft/sec2

Considering the relative values of potential energy (depth) and kinetic energy (velocity
head) in an open channel can help you with the hydraulic analysis of open channel flow
problems. Usually, you will perform these analyses using a curve that shows the
relationship between the specific energy head and the depth of flow for a given discharge
in a given channel that you can place on various slopes. Generally, you will use the curve
representing specific energy head for an open channel to identify regions of super-critical
and sub-critical flow conditions. This information usually is necessary to properly perform
hydraulic capacity calculations and evaluate the suitability of channel linings and flow
transition sections.

3.1.3.1 Specific Energy and Critical Depth


Figure 3.1-2 (Part B) illustrates a typical curve representing the specific energy head of
an open channel. The straight diagonal line on this figure represents points where the
depth of flow and specific energy head are equal. At these points, the kinetic energy is
zero; therefore, this diagonal line is a plot of the potential energy, or energy due to depth.
The ordinate interval between the diagonal line of potential energy and the specific energy
curve for the ideal discharge is the velocity head, or kinetic energy, for the depth in
question. The lowest point on the specific energy curve represents flow with the minimum
content of energy. The depth of flow at this point is known as the critical depth. Express
the general equation for determining the critical depth as:

Q 2 A3
 (3.1-6)
g T

Chapter 3: Open Channel 3-10


January 2019

Drainage Design Guide


Chapter 3: Open Channel

where:
Q= Discharge, in cubic feet per second
g= Acceleration due to gravity, 32.174 ft/sec2
T= Top width of water surface, in feet
A= Cross-sectional area, in square feet

You can calculate critical depth for a given channel through trial and error by using
Equation 3.1-6. Chow (1959) presents a procedure for the analysis of critical flow that
uses the Critical Flow Section Factor (Z), defined as the ratio of the cross-sectional area
and the square root of the hydraulic depth, expressed mathematically as:

A A
Z  (3.1-7)
D A
T

where:
Z= Critical flow section factor
A= Cross-sectional area of the flow perpendicular to the direction of flow, in
square feet
D= Hydraulic depth, in feet
T= Top width of the channel, in feet

Using the definition of the critical section factor and a velocity distribution coefficient of
one, the equation for critical flow conditions is:

Q
Z (3.1-8)
g

where:
Z= Critical flow section factor
Q= Discharge, in cubic feet per second
g= Acceleration due to gravity, 32.174 ft/sec2

When you know the discharge, Equation 3.1-8 gives the critical section factor and, thus,
by substitution into Equation 3.1-6, the critical depth. Conversely, when you know the
critical section factor, you can calculate the discharge with Equation 3.1-8.

It is important to note that the determination of critical depth is independent of the channel
slope and roughness, since critical depth simply represents a depth for which the specific
energy head is at a minimum. According to Equation 3.1-6, the magnitude of critical depth
depends only on the discharge and the shape of the channel. Thus, for any given size
and shape of channel, there is only one critical depth for the given discharge, which is

Chapter 3: Open Channel 3-11


January 2019

Drainage Design Guide


Chapter 3: Open Channel

independent of the channel slope or roughness. However, if Z is not a single-valued


function of depth, it is possible to have more than one critical depth. For a given value of
specific energy, the critical depth results in the greatest discharge, or conversely, for a
given discharge, the specific energy is a minimum for the critical depth.

Example 3.1-3—Critical Depth given Discharge


Given:
Discharge = 9 ft 3 /sec
Cross Section and Roughness from Example 3.1-1

Calculate:
Critical Depth

1 1

4 6
5 ft
Note: The solution must use trial and error since you cannot implicitly solve the equations
for depth. You can perform the first trial as shown in the steps below, with the remaining
trials shown in a table. Typically, the slope of a roadside ditch channel must exceed 2
percent to have a normal depth that is super-critical. Since the slope in Example 3.1-1
and Example 3.1-2 is 0.5 percent, the critical depth is probably much less than the normal
depth of 0.83 feet calculated in Example 3.1-2 for 9 cfs. So, we will perform our trial with
an estimated depth of flow of 0.4 ft.

Step 1: Calculate Cross-Sectional Area


Cross-Sectional Area (A):
Solve for the left triangle’s area
1
A1  (4  0.4)(0.4)
2
A1  0.32 ft 2
Solve for the right triangle’s area
1
A2  (6  0.4)(0.4)
2
A2  0.48 ft 2
Solve for the rectangle’s area
A3  5  0.4
A3  2 ft 2
Cross-Sectional Area (A) = 0.32  0.48  2  2.8 ft 2

Chapter 3: Open Channel 3-12


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Step 2: Calculate Top Width


Top Width (T):
Base Length of Left Triangle + Bottom Width + Base Length of Right Triangle
(4  0.4)  5  (6  0.4)  9 ft

Step 3: Rearrange Equation 3.1-6 to Solve for Discharge


Q 2 A3

g T
A3
Q2  g
T
A3
Q g
T
2.8 3
Q  32.174  8.86 ft 3 /sec
9

The discharge calculated in Step 3 is less than 9 ft3/sec, so critical depth is greater than
0.4 feet. Use a slightly higher depth of flow for the next guess. The following table
summarizes subsequent trials. The trial-and-error process continues until you achieve the
ideal level of accuracy.

Depth (ft) Area (sq. ft.) Top Width Discharge (cfs)


0.4 2.8 9 8.858665864
0.45 3.2625 9.5 10.84467413
0.41 2.8905 9.1 9.2404111
0.404 2.83608 9.04 9.010440628

You also can solve this problem by determining the minimum specific energy, as
discussed in the previous section. The following table solves Equation 3.1-5 for depths
bracketing the critical depth determined above, and shows that the critical depth has the
minimum specific energy.

Perimeter Velocity Specific


Depth (ft) Area (sq. ft.) V2/2g
(ft) (ft/sec) Energy
0.403 2.827045 9.112965 3.18354 0.1575 0.560501438
0.404 2.83608 9.123171 3.17339 0.1565 0.560499521
0.405 2.845125 9.133377 3.16331 0.15551 0.56050604

Chapter 3: Open Channel 3-13


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Most computer programs that solve water surface profiles for natural channels use the
minimum specific energy approach. For more information, refer to Chapter 5 (Bridge
Hydraulics).

3.1.3.2 Critical Velocity


The velocity at critical depth is called the critical velocity. An equation for determining the
critical velocity in an open channel of any cross section is:

vC  gd m (3.1-9)

where:
vc = Critical velocity, in feet per second
g= Acceleration due to gravity, 32.174 ft/sec2
dm = Mean depth of flow, in feet, calculated from:

A
dm  (3.1-10)
T

where:
A= Cross-sectional area, in square feet
T= Top width of water surface, in feet

3.1.3.3 Super-Critical Flow


For conditions of uniform flow, the critical depth, or point of minimum specific energy,
occurs when the channel slope equals the critical slope (i.e., the normal depth of flow in
the channel is critical depth). When channel slopes are steeper than the critical slope and
uniform flow exists, the specific energy head is higher than the critical value due to higher
values of the velocity head (kinetic energy). The specific head curve segment to the left
of critical depth in Figure 3.1-2 (Part B) illustrates this characteristic of open channel flow,
which is known as super-critical flow. Super-critical flow is characterized by relatively
shallow depths and high velocities, as shown in Figure 3.1-2 (Part A). If the natural depth
of flow in an open channel is super-critical, you can influence the depth of flow at any
point in the channel by an upstream control section. The relationship of super-critical flow
to the specific energy curve is shown in Figure 3.1-2 (Parts A and B).

3.1.3.4 Sub-Critical Flow


When channel slopes are flatter than the critical slope and uniform flow exists, the specific
energy head is higher than the critical value due to higher values of the normal depth of
flow (potential energy). The specific head curve segment to the right of critical depth in
Figure 3.1-2 (Part B) illustrates this characteristic of open channel flow, which is known

Chapter 3: Open Channel 3-14


January 2019

Drainage Design Guide


Chapter 3: Open Channel

as sub-critical flow. Sub-critical flow is characterized by relatively large depths with low
velocities, as shown in Figure 3.1-2 (Part C). If the natural depth of flow in an open
channel is sub-critical, a downstream control section can influence the depth of flow at
any point in the channel. The relationship of sub-critical flow to the specific energy curve
is shown in Figure 3.1-2 (Parts B and C).

3.1.3.5 Theoretical Considerations


There are several noteworthy points about Figure 3.1-2. First, at depths of flow near the
critical depth for any discharge, a minor change in specific energy will cause a much
greater change in depth. Second, the velocity head for any discharge in the sub-critical
portion of the specific energy curve in Figure 3.1-2 (Parts B and C) is relatively small when
compared to specific energy. For this sub-critical portion of the specific energy curve,
changes in depth of flow are approximately equal to changes in specific energy. Finally,
the velocity head for any discharge in the super-critical portion of the specific energy curve
increases rapidly as depth decreases. For this super-critical portion of the specific energy
curve, changes in depth are associated with much greater changes in specific energy.

3.1.4 Non-Uniform Flow


In locations where changes in the channel section or slope will cause non-uniform flow
profiles, you cannot directly solve Manning's Equation since the energy gradient for this
situation does not equal the channel slope. Three typical examples of non-uniform flow
are illustrated in Figures 3.1-3 through 3.1-5, below. The following sections describe these
non-uniform flow profiles and briefly explain how to use the total head line for
approximating these water surface profiles in a qualitative manner.

Chapter 3: Open Channel 3-15


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Figure 3.1-3: Non-Uniform Water Surface Profile for Downstream Control Caused
by a Flow Restriction

Figure 3.1-4: Non-Uniform Water Surface Profile Caused by a Change in Slope


Conditions

Figure 3.1-5: Non-Uniform Water Surface Profile Caused by a Hydraulic Jump

Chapter 3: Open Channel 3-16


January 2019

Drainage Design Guide


Chapter 3: Open Channel

3.1.4.1 Gradually Varied Flow


Figure 3.1-3 illustrates a channel on a mild slope (sub-critical) discharging into a reservoir
or pool. The figure exaggerates the vertical scale for clearer illustration.

Cross Section 1 is upstream of the pool, where uniform flow occurs in the channel. Cross
Section 2 is at the beginning of a level pool. The depth of flow between Sections 1 and 2
is changing, and the flow is non-uniform. The water surface profile between the sections
is known as a backwater curve and is characteristically very long.

Figure 3.1-4 illustrates a channel in which the slope changes from sub-critical (mild) to
super-critical (steep). The flow profile passes through critical depth near the break in slope
(Section 1). This is true whether the upstream slope is mild, as in the sketch, or the water
above Section 1 is ponded, as would be the case if Section 1 were the crest of a dam
spillway. If, at Section 2, you were to compute the total head, assuming normal depth on
the steep slope, it would plot above the elevation of total head at Section 1 (Point “a” in
Figure 3.1-4). This is physically impossible, because the total head line must slope
downward in the direction of flow. The actual total head line will take the position shown
and have a slope approximately equal to So, the slope of the channel bottom, at Section
1 and approaching So farther downstream. The drop in the total head line (hloss) between
Sections 1 and 2 represents the loss in energy due to friction.

At Section 2, the actual depth (d2) is greater than normal depth (dn) because sufficient
acceleration has not occurred, and the assumption of normal depth at this point would
clearly be in error. As you move Section 2 downstream, so that the total head for normal
depth drops below the pool elevation above Section 1, the actual depth quickly
approaches the normal depth for the steep channel. This type of water surface curve
(Section 1 to Section 2) is characteristically much shorter than the backwater curve
discussed previously.

Another common type of non-uniform flow is the drawdown curve to critical depth that
occurs upstream from Section 1 (Figure 3.1-4) where the water surface passes through
critical depth. The depth gradually increases upstream from critical depth to normal depth,
provided that the channel remains uniform over a sufficient distance. The length of the
drawdown curve is much longer than the curve from critical depth to normal depth in the
steep channel.

3.1.4.2 Gradually Varied Flow Profile Computation


Typically, you can compute water surface profiles using the Energy Equation (Equation
3.1-2). Given the channel geometry, flow, and the depth at one of the cross sections,
compute the depth at the other cross section.

Chapter 3: Open Channel 3-17


January 2019

Drainage Design Guide


Chapter 3: Open Channel

The losses between cross sections include friction, expansion, contraction, bend, and
other form losses. Expansion, contraction, bend, and other form losses will be neglected
in the computations presented in this design guide. Refer to Chapter 5 (Bridge Hydraulics)
for more information. Determine the remaining loss—the friction loss—which is express
as:

hf  S f L (3.1-11)

where:
hf = Friction head loss, in feet
Sf = Slope of the energy grade line, in feet per feet
L= Flow length between cross sections, in feet

Calculate the slope of the energy grade line at each cross section by rearranging
Manning’s Equation (Equation 3.1-4) into the following expression:

2
 Qn 
S   2

 (3.1-12)
 1.49 AR 3 

For uniform flow, the slope of the channel bed, the slope of the water surface (hydraulic
grade line), and the slope of the energy grade line are all equal. For non-uniform flow,
including gradually varied flow, each slope is different.

Use the slope determined at each cross section to estimate the average slope for the
entire flow length between the cross sections. You can use several different averaging
schemes to estimate the average slope, and these techniques are discussed in more
detail in the Chapter 5 (Bridge Hydraulics). The simplest estimate of slope of the energy
gradient between two sections is:

S1  S 2
Sf  (3.1-13)
2

where:
S1, S2 = Slope of the energy gradient at Sections 1 and 2, in feet per feet

Computing backwater curves in a quantitative manner can be quite complex. If you


require a detailed analysis of backwater curves, consider using computer software for this
purpose. Typical computer programs used for water surface profile computations include
HEC-RAS by the U.S. Army Corps of Engineers, HEC-2 by the U.S. Army Corps of
Engineers (1991), E431 by the USGS (1984), and WSPRO by the USGS (1986). In

Chapter 3: Open Channel 3-18


January 2019

Drainage Design Guide


Chapter 3: Open Channel

addition, textbooks by Chow (1959), Henderson (1966), or Streeter (1971), and


publications by the USGS (1976b), Brater and King (1976), or the USDA, SCS (NEH-5,
2008) may be useful.

Example 3.1.4—Gradually Varied Flow Example


Upon consultation, the District Drainage Engineer approved an exception to the
minimum ditch bottom width (5.0 ft.) due to a right-of-way constraint. The ditch cross
section previously used must be reduced to a 3.5-foot bottom width and a 1:3 back
slope for a distance of 100 feet. The transition length between the two ditch shapes is
15 feet.

Given:
Discharge = 25 ft 3 /sec
Roughness = 0.04
Cross Section from Example 3.1-1
Slope = 0.005 ft/ft

Calculate:
Depth of flow in narrower cross section

R/W

15’

100’

15’

Figure 3.1-6: Plan View

You can estimate the flow depths in the two cross sections using the slope conveyance
method, which solves Manning’s Equation and assumes that the ditch is flowing at normal
depth. Example C.2 (Appendix C) shows the computation of the normal depths for the

Chapter 3: Open Channel 3-19


January 2019

Drainage Design Guide


Chapter 3: Open Channel

ditch in this problem using the nomographs in Appendix C. The normal depth in the
standard ditch is 1.12 feet, and the normal depth in the narrowed ditch is 1.25 feet.

Although it is not standard practice to perform a standard step backwater analysis in a


roadside ditch, solving this example will illustrate how a gradually varied profile can be
computed using Equations 3.1-2 and 3.1-10 through 3.1-12.

The Froude Number (Fr) for normal depth flow at the first section is:

1 1
Area  (1.12  5)  (6  1.12)(1.12)  (4  1.12)(1.12)  11.87 sq. ft.
2 2
A 11.87
T  5  (6  4)1.12  16.2 ft. D   0.733
T 16.2
Q 25
v   2.11 fps
A 11.87
v 2.11
Fr  1
 1
 0.43
( gD) 2 (32.174  0.733) 2

Because Fr is less than one, the flow in the channel will be sub-critical. Therefore, you
will start the analysis at the downstream cross section and proceed upstream. Assume
normal depth in the standard ditch at a point just downstream of the downstream transition
(Section 1 in the figure above). This assumes that the ditch downstream is uniform for a
sufficient distance to establish normal depth at Section 1.

The water depth at Section 1 is 1.12 feet, as determined in Example C.2 (Appendix C).
The first row of the table on the next page shows this depth, along with other geometric
and hydraulic values needed for the computations. The elevation, z, is arbitrarily taken as
zero. Next, you will determine the depth at Section 2 from a trial-and-error procedure. The
first trial depth will be the normal depth at Section 2, which is 1.25 feet. Use Equations
3.1-10, 3.1-11, 3.1-12, and 3.1-2 to back calculate the depth at Section 2. The back-
calculated depth of 1.11 feet is shown in the last column. You can assume additional trial
depths until the trial and the back-calculated depths agree at the chosen level of accuracy.

After you have calculated the depth at Section 2, then calculate the depth at Section 3
using the same trial-and-error process. Repeat the same process to solve for the depth
at Section 4.

Chapter 3: Open Channel 3-20


January 2019

Drainage Design Guide


Chapter 3: Open Channel

1 2 3 4 5 6 7 8 9 10 11
Depth
XS Area Perimeter Radius Velocity V2 / 2g Z EGL Slope Loss Depth
Guess
# (ft2) (ft) (ft) (ft /s) (ft) (ft) (ft) (ft) (ft)
(ft)
1 1.12 11.872 16.43057 0.72256 2.105795 0.068912 0 1.188912 0.005

1.25 11.40625 15.0563 0.75757 2.191781 0.074655 0.075 1.399655 0.00504 0.075301 1.114558
1.1 9.295 13.66954 0.67998 2.689618 0.112421 0.075 1.287421 0.008766 0.103245 1.104737
1.104 9.348672 13.70652 0.68206 2.674177 0.111134 0.075 1.290134 0.00863 0.102228 1.105007
2 1.105 9.362113 13.71577 0.68258 2.670337 0.110815 0.075 1.290815 0.008597 0.101977 1.105075

1.25 11.40625 15.0563 0.75757 2.191781 0.074655 0.575 1.899655 0.00504 0.681854 1.323014
1.29 12.00345 15.4261 0.77812 2.082735 0.067411 0.575 1.932411 0.004392 0.649423 1.297827
1.296 12.09427 15.48157 0.78120 2.067094 0.066403 0.575 1.937403 0.004303 0.645002 1.294414
3 1.295 12.07911 15.47233 0.78069 2.069688 0.066569 0.575 1.936569 0.004318 0.645732 1.294977

1.12 11.872 16.43057 0.72256 2.105795 0.068912 0.65 1.838912 0.004955 0.069549 1.287206
1.28 14.592 18.06351 0.80781 1.713268 0.045616 0.65 1.975616 0.002827 0.053585 1.294538
1.294 14.84218 18.20639 0.81522 1.684389 0.044091 0.65 1.988091 0.002699 0.052628 1.295107
4 1.295 14.86013 18.2166 0.81575 1.682355 0.043985 0.65 1.988985 0.002691 0.052562 1.295147
Column 2. Use Area formula for trapezoid with the depth guessed in Column 1
Column 3. Use Wetted Perimeter formula for trapezoid with depth guessed in Column 1
Column 4. Column 2 ÷ Column 3
Column 5. Q ÷ Column 2
Column 8. Column 1 + Column 6 + Column 7
Column 9. Solve Equation 3.1-12 using Column 2 and Column 4 values
Column 10. Calculate Sf with Equation 3.1-13 using Column 9 from this row and last row of previous section. Calculate the loss with Equation 3.1-11
by multiplying Sf by the distance to the previous cross section.
Column 11. Back calculate Depth by calculating the Total Energy (Col. 8 of previous cross section + Col. 10) and subtracting the Datum and the
Velocity Head (Col. 7 + Col. 6).

Chapter 3: Open Channel 3-21


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Looking at the results of the profile analysis on the previous page, there are several things
you might not expect. First, the flow depth at Section 2 (1.105 feet) is less than the flow
depth at Section 1 (1.12 feet), which might be unexpected because the normal depth of
Section 2 is greater than Section 1. However, this is not an unusual occurrence in
contracted sections. The reason that the flow depth decreases is because the velocity,
and, therefore, the velocity head, increases. The increase in the velocity head is greater
than the losses between the sections; therefore, the depth must decrease to balance the
energy equation. The opposite can occur in an expanding reach, resulting in an
unexpected rise in the flow depth even though the normal depth decreases.

The next unusual result is that the flow depth at Section 3 is greater than the normal depth
in the narrow section. Since the flow depth is less than normal depth at Section 2, the
water surface profile should approach normal depth from below as the calculations
proceed upstream. Therefore, the flow depth at Section 3 should be less than the normal
depth. The reason that the profile jumps over the normal depth line is because of
numerical errors introduced by Equation 3.1-13. When the change in the energy gradient
between two cross sections is too large, Equation 3.1-13 does not accurately estimate
the average energy gradient between the sections. Cross sections must be added
between these cross sections to reduce the numerical errors to an acceptable amount.

This example was solved using HEC-RAS with the extra cross sections added. The
details are described below, but the results indicate that the flow depth essentially
converges to normal depth within the 100-foot distance between Sections 2 and 3. The
normal depth is 1.25 feet compared to the 1.24 feet computed by HEC-RAS at Section 3.
This Illustrates one of the primary reasons that water surface profiles are not
necessary in the typical roadside ditch design. The water depth does not significantly
vary from normal depth at any location. So, assuming that the design includes some
freeboard, the ditch will operate adequately when designed by assuming normal depth.

HEC-RAS Solution:

Four cross sections with the trapezoidal ditch shapes and slope were input into the
program. The expansion and contraction coefficients were changed to zero so that the
only the friction loss will be calculated. The friction loss method also was changed to the
Average Friction Loss to match Equation 3.1-13. The results of the analysis are shown
below.

Chapter 3: Open Channel 3-22


January 2019

Drainage Design Guide


Chapter 3: Open Channel

To compare the results with the spreadsheet solution, the depth of flow must be calculated
from the water surface elevation.

Flow Depth
Section River Station Water Surface Z
(Ft.)
1 0 1.12 0 1.12
2 15 1.18 0.075 1.11
3 115 1.87 0.575 1.30
4 130 1.94 0.65 1.29

The flow depths match the solution in Section 2. However, a conveyance ratio warning at
Section 3 indicates a possible error at that location. To improve the analysis, extra cross
sections were inserted between Section 2 and 3. Four cross sections are added by
interpolation and the profile is recomputed. The results are shown below:

Chapter 3: Open Channel 3-23


January 2019

Drainage Design Guide


Chapter 3: Open Channel

The new flow depth at Section 3 is 1.81 – 0.575 = 1.24 feet. The profile in the narrow
section has essentially converged to normal depth (1.25 feet). The depth of the complete
profile is shown below:

Flow Depth
Section River Station Water Surface Z
(Ft.)
1 0 1.12 0 1.12
2 15 1.18 0.075 1.11
35 1.35 0.175 1.18
55 1.48 0.275 1.21
75 1.60 0.375 1.23
95 1.71 0.475 1.24
3 115 1.81 0.575 1.24
4 130 1.90 0.65 1.25

3.1.4.3 Rapidly Varied Flow


A hydraulic jump occurs as an abrupt transition from super-critical to sub-critical flow. You
should consider the potential for a hydraulic jump in all cases where the Froude Number
is close to 1.0 and/or where the slope of the channel bottom changes abruptly from steep
to mild. For grass-lined channels, unless the erosive forces of the hydraulic jump are
controlled, serious damage may result.

It is important to know where a hydraulic jump will form, since the turbulent energy
released in a jump can cause extensive scour in an unlined channel. For simplicity, you
can assume that the flow in the channel is uniform except in the reach between the jump
and the break in the channel slope. The jump may occur in either the steep channel or
the mild channel, depending on whether the downstream depth is greater or less than the
depth sequent to the upstream depth.

Using the equation below, you can calculate the sequent depth:

2
d1 d 2v d
d2    1  1 1 (3.1-14)
2 4 g

where:
d2 = Depth below jump, in feet
d1 = Depth above jump, in feet
v1 = Velocity above jump, in feet per second
g= Acceleration due to gravity, 32.174 ft/sec2

Chapter 3: Open Channel 3-24


January 2019

Drainage Design Guide


Chapter 3: Open Channel

If the downstream depth is greater than the sequent depth, the jump will occur in the steep
region. If the downstream depth is lower than the sequent depth, the jump will move into
the mild channel (Chow). For more discussion on the location of hydraulic jumps, refer to
Open-Channel Hydraulics, by V.T. Chow, PhD.

When you have determined the location of the jump, you can determine the length using
Figure 3.1-7. This figure plots the Froude Number of the upstream flow against the
dimensionless ratio of jump length to downstream depth. The curve was prepared by V.T.
Chow from data gathered by the Bureau of Reclamation for jumps in rectangular
channels. You also can use the curve for approximate results for jumps formed in
trapezoidal channels.

Figure 3.1-7: Lengths of Hydraulic Jumps

When you have determined the location and the length of the hydraulic jump, you can
determine the need for alternative channel lining, as well as the limits the alternative lining
will need to be applied.

Detailed information on the quantitative evaluation of hydraulic jump conditions in open


channels is available in publications by Chow (1959), Henderson (1966), and Streeter
(1971), and in HEC-14 from USDOT, FHWA (1983). In addition, handbooks by Brater and
King (1976) and the USDA, SCS (NEH-5, 2008) may be useful.

Chapter 3: Open Channel 3-25


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Example 3.1-5—Hydraulic Jump Example

Given:
Q = 60.23 cfs
V1 = 13.81 fps
g = 32.2 ft/s2
d1 = 0.33 ft
d2 = 6.74 ft

You calculated the depths above using Manning’s Equation. The ditch has a 12.5-foot
bottom width with 1:2 side slopes. The longitudinal slopes are 10 percent and 0.001
percent, respectively. The roughness value for the proposed rubble riprap is 0.035.

Calculate:
Hydraulic Jump and the extent of rubble needed.

Step 1: Calculate Froude Number and the Length of the Hydraulic Jump
Froude Number, F1:
V
F1  1
gd1
13.81
F1 
(32.2)(0.33)
F1  4.24

Length of the Hydraulic Jump, L:


From Figure 3.1-7,
L
 5.85
d2
Therefore,
L  5.85d 2  (5.85)(6.74)
L  39.4 ft  40 ft

Step 2: Calculate the Upstream Sequent Depth


Upstream Sequent Depth, d1’:
d1 2V12 d1 d12
d1   
'

2 g 4

0.33 2(13.81) 2 (0.33) (0.33) 2


d 
'
1  
2 32.2 4

Chapter 3: Open Channel 3-26


January 2019

Drainage Design Guide


Chapter 3: Open Channel

d1'  1.81 ft

Since the downstream depth d2 (6.74 ft) is greater than the upstream sequent depth d1’
(1.81 ft), the hydraulic jump occurs in the steep region.

d2
d 1’
d1

Assuming a more conservative approach, you can split the length of the hydraulic jump
between the two regions and provide rubble riprap ditch protection for 20 feet
downstream.

3.1.5 Channel Bends

At channel bends, the water surface elevation increases at the outside of the bend
because of the super-elevation of the water surface. Additional freeboard is necessary in
bends, and you can calculate it using the following equation:

V 2T
d  (3.1-15)
gRC
where:
∆d = Additional freeboard required because of super-elevation, in feet
V= Average channel velocity, in feet per second
T= Water surface top width, in feet
g= Acceleration due to gravity, in feet per second squared
RC = Radius of curvature of the bend to the channel centerline, in feet

Chapter 3: Open Channel 3-27


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Example 3.1-6—Channel Bend Example

The channel of Example 3.1-2 takes a 45-degree bend with a radius of 30 feet. What is
the increased depth on the outside of the channel at the bend?

From Example 3.1-2, V = 1.192 ft/sec

Calculate Top Width

T  5  0.826(4  6)  13.26 ft.

V 2T 1.192 2 (13.26)
d    0.02 ft
gRC 32.174(30)

The depth of flow on the outside edge of the ditch is 0.86 + 0.02 = 0.88 ft.

The super-elevation is insignificant for this example problem, as it is for many ditches in
Florida. The variable that affects water surface super-elevation the most is the velocity
because it is squared in Equation 3.1-15. Ditches with a high velocity at a bend with a
small radius will have greater super-elevations.

Chapter 3: Open Channel 3-28


January 2019

Drainage Design Guide


Chapter 3: Open Channel

3.2 OPEN CHANNEL DESIGN


You were given the channel shape, slope, and roughness in the previous example
problems. From these example problems, you determined the flow depths and velocities
using the analysis methods described in this chapter. If a project incorporates existing
channels, then you can apply the analysis methods to those channels just as you applied
them to the example problems. However, many projects will require you to design new
channels. This section discusses how you select the channel geometry and channel
linings for FDOT projects.

3.2.1 Types of Open Channels for Highways


You can classify open channels generally as those that occur naturally and those that are
manmade, including improved natural channels. The latter, called artificial channels, are
used on most roadway projects. The types of channels commonly used on FDOT projects
are listed in Chapter 2 of the Drainage Manual:

 Roadside Ditch
 Median Ditch
 Interceptor Ditch
 Outfall Ditch
 Canals

Section 2.2 of the Drainage Manual recommends design frequencies for each of these
channel types.

The roadside ditch receives runoff from the roadway pavement and shoulders as directed
by the cross slope and shoulder slopes. The roadside ditch also may receive flow from
offsite drainage areas on adjacent properties. The roadside ditch also may intercept
ground water to protect the base of the roadway. The roadside ditch conveys the flow to
an outfall point, although the ditch may flow into other ditches or components of the
stormwater management system before reaching the ultimate outfall point from FDOT
right of way. Depressed medians will collect runoff and a median ditch will be needed to
convey runoff to an outfall point. In general, roadside and median ditches are relatively
shallow trapezoidal channels, while swales are shallow, triangular, zero-bottom-width
channels.

Interceptor ditches have various purposes. They provide a method for intercepting offsite
flow above cut slopes, thereby controlling slope erosion. They can also collect offsite flow
and keep it separate from the project stormwater. This flow can bypass the stormwater
treatment facilities, reducing their size and cost.

Chapter 3: Open Channel 3-29


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Design outfall ditches, in most cases, to receive runoff from numerous secondary
drainage facilities, such as roadside ditches or storm drains. The delineation between a
roadside ditch and an outfall ditch can become blurred. If the discharge from a stormwater
management facility is brought back to the roadside ditch to convey the flow to another
point on the project for ultimate discharge, then consider the roadside ditch to be an outfall
ditch for the purpose of selecting the design frequency. If you combine considerable flows
from offsite areas and onsite project flows together in the roadside ditch to become a
significant discharge, then consider the roadside ditch to be an outfall ditch for the
purpose of selecting the design frequency. It is unwise to use a roadside ditch as an outfall
ditch, since its probable depth and size could create a potential hazard.

Canals, like outfalls, also are large artificial channels that accept flows from other
drainage components. The added connotation of a canal is that there is always water in
the channel, unlike many outfalls that only flow immediately after a rainfall event. If the
canal, which always has water, is close to the road, then it can be a potential hazard. For
the purpose of identifying a hazard, the FDM defines a canal as an open ditch parallel to
the roadway for a minimum distance of 1,000 feet, and with a seasonal water depth in
excess of three feet for extended periods of time (24 hours or more). Water Management
Districts and local agencies may have a different definition for canals when determining
regulatory jurisdiction.

Other FDOT publications mention other types of ditches. Right-of-way ditches are
mentioned in the Standard Specifications and a detail is given on Standard Plans, Index
524-001. The right-of-way ditch often functions as a type of relief ditch, handling drainage
needs other than those for the roadway and thus freeing roadside ditches from carrying
anything except roadway runoff. You usually can consider right-of-way ditches as
interceptor ditches when selecting the design frequency.

The term “lateral ditch” is used in the FDM and the Standard Specifications. The term is
used to determine:

 How the ditch excavation will be paid for


 How the ditch is shown in the plans

A lateral ditch generally is perpendicular to the roadway and can flow either toward or
away from the road. However, a lateral ditch also can run parallel to the road right of way
if the ditch or channel is separate from the roadway template. Refer to the FDM for
guidance on selecting the excavation pay item. Consider the purpose of the lateral ditch
and associate it with one of the ditch types listed above to select the design frequency.

Several FDOT publications use the term roadway ditch rather than roadside ditch. These
two terms are interchangeable. Other FDOT publications or engineers performing work
for the Department also may use many other terms to refer to open channels. The

Chapter 3: Open Channel 3-30


January 2019

Drainage Design Guide


Chapter 3: Open Channel

definitions of most of these terms are self-explanatory because of their descriptive names.
Some examples are:

 Drainage ditch
 Stormwater ditch
 Bypass ditch
 Diversion ditch
 Conveyance channel
 Agricultural ditch
 Treatment swale

A swale is a special kind of artificial ditch that has become important in Florida. The
following legal definition of a swale as it relates to the regulation and treatment of
stormwater discharge is from Chapter 62-25.020, Florida Administrative Code (FAC):

"Swale" means a manmade trench which:


a) has a top width-to-depth ratio of the cross section equal to or greater than 6:1, or
side slopes equal to or greater than 3 feet horizontal to one-foot vertical; and
b) contains contiguous areas of standing or flowing water only following a rainfall
event; and
c) is planted with or has stabilized vegetation suitable for soil stabilization,
stormwater treatment, and nutrient uptake; and
d) is designed to take into account the soil erodibility, soil percolation, slope, slope
length, and drainage area so as to prevent erosion and reduce pollutant
concentration of any discharge.

3.2.2 Roadside Ditches


You can design roadside ditches using the following steps:

Step 1—Establish a Preliminary Drainage Plan. Roadside ditches will be components


of an overall drainage system. Since the roadside ditch generally will follow the grade of
the road, the high points in the roadway grade will be initial drainage boundaries.
However, you can adjust these boundaries by using special ditch grades so that the ditch
flows in a different direction than the roadway grade. You also can adjust the boundaries
significantly for projects in flat terrain. It is, however, best to keep existing drainage
patterns if possible. You also can adjust low points with special ditch grades if the ideal
discharge point is not at the low point of the roadway grade.

Most projects will have stormwater management facilities, so the roadside ditches will
connect with the conveyance components to the various facilities. Not all portions of the
roadside ditch can physically be directed to a stormwater management facility, so short

Chapter 3: Open Channel 3-31


January 2019

Drainage Design Guide


Chapter 3: Open Channel

segments may need to discharge to other points, such as streams or ditches near cross
drains and bridges, or other points along the roadway.

When determining initial ditch grades, provide a ditch slope with sufficient grade to
minimize ponding and sediment accumulation. The Drainage Manual requires a minimum
physical slope of 0.0005 feet/feet for ditches where positive flow is required. These flat
slopes are difficult to grade during construction and clumps of grass left behind by mowers
easily impede the flow.

Existing utilities also may control the grade of the ditch to maintain minimum cover over
the utility.

Step 2—Select Standard Ditch Components. The standard roadside ditch will be
shown in the plans on the typical section. You can find standard ditch sections in the FDM
for several roadway types, and in Figures 3.2-1 and 3.2-2, below. You may need to adjust
the standard ditch due to peculiarities that are consistent throughout the project. An
example might be a narrow border width and limited right of way.

The typical ditch shown in Figure 3.2-1 for two-lane roads is narrower than most mitered
end sections. In some situations, you can use a wider typical ditch section. If the wider
ditch is not used, then check the right of way at each mitered end to be sure the right of
way will be adequate to accommodate a wider ditch at the mitered end section.

Figure 3.2-1: Typical Ditch for Two-Lane Rural Roadway

Chapter 3: Open Channel 3-32


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Figure 3.2-2: Typical Roadside and Median Ditches

If the ditch size needs to be reduced due to right-of-way limitations, you can consider the
following options:

 The front slope must remain 1:6 in the clear zone, but can break over to 1:4 at the
clear zone
 You can narrow the bottom width. Five feet is an ideal minimum, but Maintenance
and Construction may have equipment to build and maintain a two-foot bottom
width. Avoid V-bottomed ditches with steep side slopes. Refer to Chapter 2 of the
Drainage Manual for criteria regarding V-bottomed ditches. Avoid using a bottom
width narrower than the side drain endwalls.
 You can steepen the back slope if the following is considered:
o Steeper slopes are harder to maintain, especially 1:3 and steeper
o Check the soils for stability
o Significant offsite drainage down a steep back slope will cause erosion on
the slope
 You can reduce the depth to the shoulder point if the following is considered:
o Check the ditch capacity
o Consider the type of facility and base clearance needs

Chapter 3: Open Channel 3-33


January 2019

Drainage Design Guide


Chapter 3: Open Channel

 You also can enclose the ditch with a pipe system, although a ditch or swale
usually still is needed to collect the roadway runoff into inlets. Enclosing the
system will increase construction costs, but may be less expensive than
obtaining more right of way.

Step 3—Check for locations where the standard ditch will not work. A good way to
check is to plot the standard ditch on the cross sections. Look for places where the ditch
extends beyond the right of way or conflicts with utilities and other obstructions. Also look
in the Plan View to check for obstructions between the cross sections.

You can adjust the size of the ditch while also considering the same issues identified in
the previous step. If the grade of the ditch must be adjusted, then you must develop a
special ditch profile and plot it in the plans. Some locations where the ditch grade may
need to be adjusted include:

 Outfall locations—The grade of the standard ditch will follow the grade of the road.
If the outfall location is not at the lowest point in the roadway profile, then you need
to develop a special ditch profile.
 Locations of high water table—These areas may require feedback to the roadway
designer to raise the roadway grade.
 Cross drains, median drains, and side drains—These structures may need to be
at a lower elevation than the standard ditch elevation. If the entrance end of the
culvert is depressed below the stream bed, more head is exerted on the inlet for
the same headwater elevation. Usually, the sump is paved, but for small
depressions, an unpaved excavation may be adequate.
 Locations where the top of the back slope creates a ditch that is too shallow—
Sometimes, you can use a berm to contain the ditch instead of changing the grade.
Be careful that offsite drainage is not blocked. If you use a berm, provide an
adequate top width and side slopes for ease of maintenance. A suggested
minimum top width is three feet, but five feet is ideal.

You will need to develop special ditch profiles if the profile grade is less than the minimum
ditch slope. Refer to the Drainage Manual for minimum ditch slope criteria. At vertical
curve crests, the ditch grade will be less than the minimum ditch grade criteria given in
the Drainage Manual. (In fact, the ditch grade will go to zero at the high point.) A special
ditch grade is not necessary at a vertical curve crest.

Step 4—Compute the Flow Depths and Velocities. Although some designers check
the ditch at regular intervals, it is not necessary. Checking at critical locations is adequate.
Check the ditch at the outfall point. The discharge will be greatest at this location, so it
may represent the worst-case conditions for the entire ditch. Other critical locations to
check are:

Chapter 3: Open Channel 3-34


January 2019

Drainage Design Guide


Chapter 3: Open Channel

 Changes in slope, specifically steeper slopes


 Changes in shape, specifically narrower sections
 Shallowest ditch depths
 Changes in lining (roughness)
 Changes in flow

Determine the maximum allowable depth of the ditch at these sections, including
freeboard. Section 2.4.5 of the Drainage Manual provides freeboard requirements. If the
actual depth exceeds the maximum allowable depth in the ditch, then the ditch does not
have enough capacity. Possible ways to increase the ditch capacity include:

 Increase bottom width


 Make ditch side slopes flatter
 Make longitudinal ditch slope steeper
 Provide a smoother ditch lining
 Install drop inlets and a storm drain pipe beneath the ditch
 Berm up the back slope of the ditch

Step 5—Check Lining Requirements. When the ditch geometry components are set
and the depth of flow is determined to be adequate, then the ditch needs to be checked
to determine if you need a ditch lining. Check the maximum velocity in the ditch against
the allowable velocities for bare earth shown in Table 2.3 of the Drainage Manual. If these
velocities are met, then you can use the standard treatment of grassing and mulching.

If the maximum ditch velocity exceeds the allowable velocity for bare earth, then you
should provide sodding, ditch paving, or other forms of ditch lining. See Section 3.3 for
more discussion of ditch linings.

3.2.3 Median Ditches


The design steps for median ditches are similar to those for roadside ditches.

Step 1—Establish a Preliminary Drainage Plan. As with roadside ditches, median


ditches also will be components of an overall drainage system. The grade of the median
ditch generally will follow the grade of the road. Generally, curbs are not provided on the
edge of the pavement and the median ditch drains part or all of the shoulder area in
addition to the median itself. Even where curbs are provided, it is preferable to slope
medians wider than 15 feet to a ditch. This keeps water in the median off the pavement.
Medians less than 15 feet wide generally are crowned for drainage, and, if they are less
than six feet in width, they usually are paved. Permitting agencies may request that the
median ditch be depressed.

Chapter 3: Open Channel 3-35


January 2019

Drainage Design Guide


Chapter 3: Open Channel

When the width of the median ditch is established, locate outfall points from the median.
If the travel lanes slope to the outside and the median is impervious, then the median
runoff may not need to be conveyed to a stormwater treatment facility. The median may
be able to discharge directly into cross drains via inlets.

Median cross overs, bridge piers, or other structures often interrupt continuous flow in
medians. Decide whether to convey around the obstruction or to one side of the roadway.
Consider the flow depth in the median, feasible means to convey the flow around the
obstruction, the size of pipe to convey the flow to the outside, the cover available, and the
elevation of the roadside ditch to which the flow will be conveyed. Also consider the actual
low point of the median ditch, which is usually at the low point of the roadway grade. This
may be affected by guardrail, turn lanes, etc. Turn lanes and other non-typical roadway
configurations also may create a depressed gore area. You will need to analyze these
areas with methods similar to those used for roadside ditches.

Considerations to determine which side of the roadside to discharge to include:

 Maintenance of traffic phasing and construction sequencing


 Which side the outfall or stormwater facility is located on
 Commingling with offsite runoff

Step 2—Select Standard Ditch Components. The standard median ditch will be shown
in the Plans on the Typical Section. Standard ditch sections are given in the FDM for
several roadway types, and one is shown in Figure 3.2-2.

Step 3—Compute the Flow Depths and Velocities. Determine critical locations to
check depth of flow and velocities, as outlined above. In addition to the critical areas for
the roadside ditch, you also should evaluate the median ditch in gore areas caused by
turn lanes or additional pavement. If the actual depth exceeds the maximum allowable
depth, then you will need to increase the capacity of the ditch. Use methods similar to
those for increasing the capacity of a roadside ditch. Be mindful of the additional clear
zone requirements for median ditches.

Step 4—Check Lining Requirements. After you establish the section of the ditch, check
the maximum velocities against the allowable velocities for bare soil. If those velocities
are exceeded, then you need to research further to determine the appropriate lining for
the ditch. See Section 3.3 of this design guide for further discussion.

3.2.4 Interceptor Ditches


Interceptor ditches run along the natural ground near the top edge of a cut slope or along
the edge of the right of way to intercept the runoff before it reaches the roadway.

Chapter 3: Open Channel 3-36


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Interceptor ditches along the edge of the right of way are commonly referred to as right-
of-way ditches.

The interceptor ditch generally will follow the grade of the natural ground adjacent to the
project, not the profile grade of the road. If possible, locate the high points in an interceptor
ditch at the drainage divides of the adjacent property to maintain existing drainage
patterns. Low points also typically follow the adjacent terrain, allowing the interceptor ditch
to discharge to points such as streams near cross drains and bridges.

Most projects will have stormwater management facilities. These facilities often are set
off from the project area, so it is important to consider conflicts that may arise where the
outfall ditch intersects the interceptor ditch.

The design steps for interceptor ditches are the same as those for the roadside ditch. See
Section 3.2.2 for the design procedure.

3.2.5 Outfall Ditches


Since outfall ditches receive runoff from numerous secondary drainage facilities, including
stormwater management facilities, design the standard ditch section for a larger capacity.
You should evaluate the standard ditch section against the clear zone criteria for the
project. Even though outfall ditches have a larger design event and carry larger flows, the
design steps are the same as those for the roadside ditch. See Section 3.2.2 for the
design procedure.

The design also should include consideration of the following:

 The drainage area that flows into the outfall ditch by overland flow. Designers often
forget to include this area in the total drainage area when determining the design
flow rates for the outfall ditch. Another concern is erosion down the side slope from
the sheet flow from these areas. You can use spoil from the ditch construction to
create berms to block and collect the flow in inlets to prevent this erosion.
 Check for existing outfall easements. Some easements may require a specific type
of conveyance, such as a ditch or a pipe system.

3.2.6 Hydrology
As stated in Section 2.3 of the Drainage Manual, hydrologic data used for the design of
open channels will be based on one of the following methods, as appropriate for the
particular site:

 Use a frequency analysis of observed (gage) data when available

Chapter 3: Open Channel 3-37


January 2019

Drainage Design Guide


Chapter 3: Open Channel

 Use the regional or local regression equation developed by the USGS


 Use the Rational Equation for drainage areas up to 600 acres
 Use the method applied for the design of the stormwater management facility in
the design of the outfall from this facility
 Request hydrologic data from the controlling entity for regulated or controlled
canals

For a more detailed discussion on procedure selection and method for calculating runoff
rates, refer to Chapter 2 (Hydrology).

3.2.6.1 Frequency
Roadside or median ditches or swales, including bypass and interceptor ditches, usually
are designed to convey a 10-year frequency storm without damage; outfall ditches or
canals should convey a 25-year frequency storm without damage. However, because the
risks and drainage requirements for each project are unique, site-specific factors may
warrant the use of an atypical design frequency. Regardless of the frequency selected,
you should always consider the potential for flooding that exceeds standard criteria. Pre-
development stages for all frequencies up to and including the 100-year event must not
be exceeded unless flood rights are obtained or the flow is contained within the ditch.

It also is important to consider sediment transport requirements for conditions of flow


below the design frequency. A low flow channel component within a larger channel can
reduce the maintenance effort by improving sediment transport in the channel.

Design temporary open channel facilities for use during construction to handle flood flows
commensurate with risks. The recommended minimum frequency for temporary facilities
and the temporary lining of permanent facilities is 20 percent of the standard frequency
for permanent facilities, which extrapolates as a two-year frequency for roadside ditches
and a five-year frequency for outfall ditches.

3.2.6.2 Time of Concentration


The time of concentration is defined as the time it takes runoff to travel from the most
remote point in the watershed to the point of interest. When using the Velocity Method,
calculate the time of travel for main channel flow using the velocity in the section and the
channel length. Segments used to determine the velocity should have uniform
characteristics. Use a new segment each time there is a change in the channel geometry,
such as cross section or channel slope. Calculate the time for each segment and then
add them together to determine the total time of concentration for the channel. See
Chapter 2 (Hydrology) for a discussion of methods and procedures to determine the time
of concentration.

Chapter 3: Open Channel 3-38


January 2019

Drainage Design Guide


Chapter 3: Open Channel

3.2.7 Tailwater and Backwater


The water depth at the downstream end of the ditch will affect the flow depth and velocities
in the ditch for some distance upstream. The downstream water depth, or tailwater, may
cause a backwater condition with a gradually varied water surface profile. In roadside
ditches, you can approximate the water surface profile as a flat water surface at the
tailwater (Tw) elevation that intercepts the normal depth (dn) of flow in the ditch, as shown
in Figure 3.2-3. If the tailwater depth is less than the normal depth in the ditch, then you
can approximate the water surface profile in the ditch as the normal depth in the ditch, as
shown in Figure 3.2-4. For the low tailwater condition, perform the velocity check for lining
requirements using the velocity for the tailwater depth, not the normal depth.

Actual water surface


profile
Assumed water surface
dn

Tw

Figure 3.2-3: Assumed Water Surface for Tw > dn

Actual water surface


profile
dn Assumed water surface

Tw

Figure 3.2-4: Assumed Water Surface for Tw < dn

Chapter 3: Open Channel 3-39


January 2019

Drainage Design Guide


Chapter 3: Open Channel

To summarize the water surface approximation, the water surface elevation at any point
in the ditch is the higher of the normal depth elevation or the tailwater elevation. You can
determine the frequency of the design tailwater elevation using the same
recommendations for storm drains in Section 3.4 of the Drainage Manual.

The same water surface profile assumptions illustrated above also apply to other
backwater conditions in the ditch. Side drains are an example. The water surface
elevation in the ditch at any point upstream of a side drain should be the greater of the
normal depth elevation or the headwater elevation of the culvert. The normal depth in the
ditch changes if the ditch slope, cross section, or roughness changes. If the downstream
normal depth is greater, then the assumed water surface is shown in Figure 3.2-5.

Actual water surface


profile
dn Assumed water
surface

dn

Figure 3.2-5: Assumed Water Surface for change in dn

Chapter 3: Open Channel 3-40


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Example 3.2-1—Roadside Ditch Design Example


The figures below show the plan and profile views of a proposed four-lane roadway.
Complete the design of the left roadside ditch.

400’ V.C.

1,500’ V.C.

3125 3130 3135 3145


3140

Pond

Mitered End Section

3125 3130 3135 3145


3140

Step 1—Drainage Plan. On the left side of the roadway near Station 3125+00A, there is
a stormwater pond to treat and attenuate the roadway runoff. Roadside ditches will collect
the runoff from the roadway and convey it to the cross drain, which empties into the pond.
The offsite drainage area is small; therefore, dual ditches are not needed to reduce the
size of the pond.

Chapter 3: Open Channel 3-41


January 2019

Drainage Design Guide


Chapter 3: Open Channel

The left roadside ditch will discharge into a mitered end section at Station 3126+50. The
design frequency for the ditch will be 10 years (refer to the Drainage Manual for the design
frequency). The pipe system and the pond may have different design frequencies than
the ditch, but you can determine a 10-year elevation in the pond and the 10-year hydraulic
grade line for the pipe system at the mitered end section. The hydraulic grade line of the
pipe system at this headwall will be the tailwater elevation for the ditch.

The design of the overall drainage system may be iterative. The design of one component,
such as the pond, can affect the design of other components, such as the left and right
roadside ditches, the cross drain, and even the median ditch. To simplify this example,
the tailwater elevation for the ditch will be given as 76.52 feet.

Step 2—Standard Ditch Components. The standard ditch shown in Figure 3.2-2 will be
used. The vertical distance from the profile grade line (PGL) to the ditch bottom elevation
of the standard ditch will be:

Elevation Difference = (24 ft. x 0.02) + (12 ft. x 0.06) + 3.5 ft. = 4.7 ft.

Step 3—Check for locations where the standard ditch will not work. Three reasons
why the standard ditch will not work are:

 The backslope tie in to natural ground extends beyond the right-of-way line and
acquiring additional right of way is not prudent.
 The natural ground elevation is lower than the standard ditch bottom elevation, or
low enough that the standard ditch is too shallow.
 The profile grade is less than the minimum ditch slope.

Plotting the standard ditch on the roadway cross sections is a good way to look for
locations where the standard ditch will not work. Also, starting at the downstream end of
the ditch and working upstream will afford an orderly approach to design the ditch. For
this example, the profile grade elevation will be 79.00 and the bottom of the standard ditch
will be 74.3 feet at Station 3127+00, as shown in the figure below.

Chapter 3: Open Channel 3-42


January 2019

Drainage Design Guide


Chapter 3: Open Channel

El. 79.00

El. 74.3 Sta.


3127+00

The PGL is flat (0.000 percent) between this cross section and the end section at Station
3126+50. The minimum slope of the ditch is 0.05 percent, and the ideal slope is at least
0.1 percent. Therefore, you will need a special ditch grade between these stations. If the
flowline at the headwall (Station 3126+50) is set at 74.2 feet, the ditch grade between
these stations will be 0.1/50 = 0.002, or 0.2 percent.

At this point in the design process, you would calculate the discharge at the downstream
end of the ditch. For this example, the discharge will be given as 12.7 cfs at the end
section. Refer to the Chapter 2 (Hydrology) for an explanation of how to calculate the
discharge. Solving Manning’s Equation with the standard ditch shape (five-foot bottom
width, 1:6 front slope, 1:4 back slope), the slope of 0.2 percent, n = 0.042, and the
discharge of 12.7 cfs gives a flow depth in the ditch of 1.03 feet. At the headwall, the
normal depth elevation would be 74.2 + 1.03 = 75.23 feet. This elevation is less than the
tailwater elevation. Therefore, the flow depth in the ditch is the tailwater elevation of 76.52
feet. The outside edge of the shoulder elevation is lower than the back of the ditch
elevation at this location and will, therefore, control the allowable flow depth in the ditch.
Since the tailwater elevation is lower than the allowable flow depth, the ditch depth is
adequate.

Proceed upstream to continue the design. Looking at the cross sections between Stations
3133+00 and 3136+00, the standard ditch bottom elevation will be higher than the natural
ground elevation for several hundred feet, as typified by the cross section shown below
for Station 3134+00.

Chapter 3: Open Channel 3-43


January 2019

Drainage Design Guide


Chapter 3: Open Channel

El. 87.08

El.
81.3
Sta.
3134+00

The standard ditch could be used if a berm was constructed. However, there are at least
two reasons not to construct the berm. First, some offsite flow to the ditch would be
blocked. Second, the cost of constructing the berm is unnecessary since you can use a
special ditch profile to lower the ditch into the natural ground.

The discharge needs to be determined at this point to continue the design. A conservative
assumption would be to use the discharge at the downstream end of the ditch. In this
case, the designer judges that the discharge might be significantly different and calculates
the discharge at this point. To simplify the example, the discharge at this location is given
as 10.2 cfs.

Assuming a ditch bottom elevation of about 79.3 ft (2 feet below natural ground), the slope
to Station 3127+00 would be (79.3 – 74.3)/700 = 0.007, or 0.07 percent. Selecting the
value of 2 feet was based on some preliminary calculations of the flow depth and including
some freeboard. Solving Manning’s Equation with the standard ditch shape, the slope of
0.7 percent, n = 0.042, and the discharge of 10.2 cfs gives a flow depth in the ditch of
0.68 feet. This would leave a freeboard of approximately 1.3 feet at this location, which is
more than needed. The flow depth of 0.68 feet is close enough to 0.7 feet that using n of
0.042 is reasonable given the amount of freeboard provided. A special ditch grade of 0.07
percent will be used between Stations 3127+00 and 3134+00.

The special ditch grade has to tie back into the standard ditch grade someplace further
upstream. The standard ditch bottom will return to an adequate depth into natural ground
to contain the flow at Station 3137+00. The PGL at Station 3137+00 is 91.17 feet. The
ditch bottom elevation for the standard ditch is 86.47 feet. The ditch grade will be (86.47
– 79.3)/300 = 0.0239, or 2.39 percent. Solving Manning’s Equation with the standard ditch
shape, the slope of 2.39 percent, n = 0.06, and the discharge of 10.2 cfs gives a flow
depth in the ditch of 0.59 feet and a velocity of 2.2 fps. Note that the roughness changes
because the flow depth is less than 0.7 feet. The velocity is low enough that ditch lining

Chapter 3: Open Channel 3-44


January 2019

Drainage Design Guide


Chapter 3: Open Channel

will not be needed. However, sod will be needed, instead of seed and mulch, to establish
grass during construction.

Checking the cross sections between 3134+00 and 3137+00, the ditch depth is at least
1.5 feet, which will provide acceptable freeboard.

To summarize, the special ditch grades will be:

 0.2 percent from Station 3126+50 to 3127+00


 0.07 percent from Station 3127+00 to 3134+00
 2.39 percent from Station 3134+00 to 3137+00

The standard ditch will provide an adequate depth from 3137+00 to the top of the hill.
Checking the cross section plots shows that the earthwork to construct the standard ditch
will not extend beyond the proposed right-of-way line.

Step 4—Compute the Flow Depths and Velocities. These values were calculated in
the description of the previous step. In most cases, the designer will be iterating through
Steps 3 and 4 as the ditch is designed.

Figure 3.2-6 shows the ditch checks appropriate for including in the Drainage
Documentation to prove the design.

Chapter 3: Open Channel 3-45


January 2019

Drainage Design Guide


Chapter 3: Open Channel

51
HYDRAULIC WORKSHEET FOR ROADSIDE DITCHES Sheet ___1____ of ___1_______
Road: New Road__________________________________________ Prepared by: _____XXX_________ Date: ___4/1/09_______
Project Number: 1234567___________________________________________
Checked by: _____YYY_________ Date: ___4/1/09_______

Ditch Section
Side
STATION TO % Drain Q Calculated Vel Ditch
SIDE "C" Tc I10 "n" "d" "dallowed" Drain Remarks
STATION Slope Area (cfs) F.S. B.W. B.S. Freeboard (fps) Lining
Pipe Dia

3126+50 LT 0.20 2.61 0.75 15 6.5 12.7 6 5.0 4 0.042 1.03 1.2 SOD TW El. will control

3127+00 LT 0.70 12.7 6 5.0 4 0.042 0.75 1.9 SOD TW El. will control

3134+00 LT 0.70 1.79 0.75 10 7.6 10.2 6 5.0 4 0.042 0.68 1.87 SDO

3134+00 LT 2.39 10.2 6 5.0 4 0.6 0.59 2.2 SOD

Note: F.S. = Front Slope B.W. = Bottom Width B.S. = Back Slope
Manning "N" is Transitioning as the depth Approaches 0.7'

Figure 3.2-6: Roadside Ditch Design Example

Chapter 3: Open Channel 3-46


January 2019

Drainage Design Guide


Chapter 3: Open Channel

3.2.8 Side Drains


Continuous flow in a roadside ditch can be interrupted by side street/road connections
and/or driveway connections to the project roadway. Even a limited access roadway, such
as an interstate highway, may have an occasional access driveway that will impede
roadside ditch flow, especially at or near adjacent stormwater pond locations. You can
maintain ditch flow continuity through such obstructions via roadside ditch culverts or side
drains.

A side drain is a class of culvert pipe that can transport flow through fill placed in a
roadside ditch. A side drain is normally aligned parallel or nearly parallel to the project
roadway and along the flowline of the ditch. Side drains located under public roads
connecting to the project roadway, are identified and hydraulically sized as a cross drains
(see Chapter 4, Culverts). Side drains and cross drains are similar in many ways, but
there are some differences in design analysis requirements, materials, and end treatment.
Cross drains have to meet more rigorous criteria for some parameters.

3.2.8.1 Design Analysis Requirements for Side Drains


You size a side drain for the storm frequency required to design the roadside ditch that
contains the side drain (usually the 10-year frequency, as mentioned in Section 3.2.6.1).
You can determine the side drain design flow by applying the same hydrologic method
used to compute the corresponding ditch design flows (usually the Rational Equation,
described in Section 2.2.3). Then, you can determine the side drain pipe dimensions via
the inlet-control/outlet-control procedure described in Section 4.5. (Note: The FHWA HY-8
computer software is one of several computer programs capable of applying this
procedure to the side drain design data.)

You will normally develop the design flow for a side drain in the design calculations
spreadsheet or worksheet for the roadside ditch that contains the side drain. (Figure 2-1
of the Drainage Manual depicts such a ditch design worksheet.) The design flow and
surface water depth for the ditch section at the upstream end of the side drain are
determined in the ditch calculations, and this ditch flow is the side drain design inflow as
well. This flow typically is also the design flow for the ditch section at the downstream end
of the side drain, and must be accounted for in the calculations for the remainder of the
downstream ditch length. Of course, if additional flow enters the side drain between its
upstream and downstream ends, this additional flow also must be appropriately
accounted for in both the side drain hydraulic design and in the downstream ditch design
calculations.

Determine the tailwater elevation at the culvert outlet. Since the culvert usually is placed
through fill in the roadside ditch, the ditch calculations downstream of the culvert are used
to determine the tailwater. The culvert tailwater will be the normal depth in the

Chapter 3: Open Channel 3-47


January 2019

Drainage Design Guide


Chapter 3: Open Channel

downstream ditch unless the tailwater for the ditch controls the water surface elevation at
the side drain outlet. Refer to Section 3.2.7 for more discussion on tailwater.

Then you can generate the hydraulic calculations for a side drain, using the procedure
described above to determine the pipe dimensions needed to safely pass the design flow
to the downstream ditch segment. Include these side drain calculations in the Drainage
Documentation Report as either a separate section or as part of the Ditch Calculations
section.

Note that the surface water depth computed for culvert flow at the upstream end of a side
drain generally will be larger than the depth computed for ditch flow at that location. If the
difference in this flow depth is not significant, evaluate the ditch flow depths upstream
from the side drain and adjust (if appropriate) for the “flat pool” that will be established in
the ditch by the higher of the two water surface elevations. If the difference in surface
water flow depth at the side drain is substantial and the ditch design is sensitive to actual
flow depths, a backwater analysis may be needed rather than the “flat pool” approximation
in determining the actual flow depth estimates.

3.2.8.2 Material Requirements


In general, side drains are not considered to be as critical as cross drains. Therefore,
material service life requirements for side drains are less stringent than for cross drains.
Consult Chapter 6 of the Drainage Manual, the FDOT Standard Specifications, Chapter
8 (Optional Pipe Materials) in this handbook, and the appropriate District Drainage
Engineer for any clarification needed on pipe materials acceptable for use as side drains.
Culvert and ditch calculations may show the need for two allowable pipe sizes, depending
on the Manning’s roughness coefficients of the optional pipe materials for the side drain.

3.2.8.3 End Treatment


The only allowable side drain end treatment is the mitered end section (Standard Plans,
Index 430-022). Due to the normal side drain alignment and close proximity to the project
roadway (usually within the clear zone), Standard Plans, Index 430-022 specifies that
grates be installed for the larger pipe sizes. The grates are intended to provide a measure
of safety for errant vehicles that encounter the end treatment. The grates, however, will
potentially collect debris and will increase the entrance loss coefficient, Ke, from 0.7 to
1.0 for the mitered end section. When a grate is likely to be used, consider the following
items:

 Recognize that the specification of a grate could increase the required side drain
size (due to the increase in Ke).
 In critical hydraulic locations, evaluate the potential debris transport prior to using
grates. Vegetated ditch grades in excess of 3 percent, pipe with less than 1.5 feet

Chapter 3: Open Channel 3-48


January 2019

Drainage Design Guide


Chapter 3: Open Channel

of cover, or paved ditch grades in excess of 1 percent will require such an


evaluation.
 Determine highly corrosive locations and specify in the plans when the grates need
to be hot-dipped galvanized after fabrication.

Example 3.2-2 – Side Drain Design

Problem Statement:
A driveway is included in the design of the left roadside ditch for a new two-lane rural
roadway segment. Figure 3.2-1 depicts the typical section for the left side of the roadway.
The ditch extends and flows from Station 10+00 to Station 45+00, with the centerline of
the driveway located at Station 40+00. The width of the proposed driveway base at the
ditch flowline is 40 feet, and the ditch section is uniform throughout its length with a 2-foot
allowable depth below the left top-of-bank. At its upstream and downstream ends, the
ditch flowlines must match elevations of 100.0 feet and 96.0 feet, respectively. The
following sketch shows the ditch longitudinal slopes are 0.1 percent from Station 10+00
to Station 35+00, and 0.15 percent from Station 35+00 to Station 45+00. The site is
located in FDOT IDF Curve Zone 7, and the natural ground slopes away from the left top-
of-bank of the ditch section.

Determine the required side drain diameter.

Design Approach:
First, develop the ditch design calculations to determine the side drain design inflow at
Station 39+80. These calculations are shown on Figure 3.2-7, and identify a side drain
design flow of 4.60 cfs.

Next, refer to Section 4.5 for the side drain hydraulic design procedure. Use either the
inlet control and outlet control nomographs from FHWA HDS-5, or software such as HY-
8, to develop the required side drain size.

Chapter 3: Open Channel 3-49


January 2019

Drainage Design Guide


Chapter 3: Open Channel

El. 100

0.1% 40+00
0.15%
Culver
t
El. 96
10+00
35+00

40’
45+00

Chapter 3: Open Channel 3-50


January 2019

Drainage Design Guide


Chapter 3: Open Channel

51
HYDRAULIC WORKSHEET FOR ROADSIDE DITCHES Sheet ___1____ of ___1_______
Road: New Road__________________________________________ Prepared by: _____XXX_________ Date: ___4/1/09_______
Project Number: 1234567___________________________________________
Checked by: _____YYY_________ Date: ___4/1/09_______

Ditch Section
Side
STATION TO % Drain Q Calculated Vel Ditch
SIDE "C" Tc I10 "n" "d" "dallowed" Drain Remarks
STATION Slope Area (cfs) F.S. B.W. B.S. Freeboard (fps) Lining
Pipe Dia

10+00 - Seed &


LT 0.10 2.75 0.47 60.1 3.24 4.19 6 5.0 4 0.042 0.7 0.7
35+00 Mulch

35+00 - Seed & Drain Area includes 1/2 of


LT 0.15 3.31 0.47 69.5 2.96 4.60 6 5.0 4 0.042 0.67 0.83
39+80 Mulch driveway width

39+80 - See Side Drain Calcs for


LT 4.60 18"
40+20 details

40+20 - Seed & Drain Area includes 1/2 of


LT 0.15 3.86 0.47 78.6 2.72 4.93 6 5.0 4 0.042 0.69 0.85
45+00 Mulch driveway width

Note: F.S. = Front Slope B.W. = Bottom Width B.S. = Back Slope
Manning "N" is Transitioning as the depth Approaches 0.7'

Figure 3.2-7: Side Drain Design Example

Chapter 3: Open Channel 3-51


January 2019

Drainage Design Guide


Chapter 3: Open Channel

3.3 CHANNEL LININGS


As stated in Section 2.4.3 of the Drainage Manual, when designing open channels, you
need to consider channel linings. Erosion and sloughing cause most maintenance
problems in channels. Channel linings often solve these problems. The Standard Plans,
and the Standard Specifications identify standard lining types. The two main
classifications of open channel linings are flexible and rigid. Flexible linings include
vegetative linings such as grass, rubble riprap, and geotextile or interlocking concrete
grids. Rigid linings include concrete, asphalt, and soil-cement. From an erosion control
standpoint, the primary difference between rigid and flexible channel linings is their
response to changes in channel shape (i.e., width, depth, and alignment). For most
artificial channels, the ideal lining is natural, emerging vegetation, with grass used to
provide initial and long-term erosion resistance.

The following are examples of lining materials in each classification.

1. Flexible Linings:
a. Grasses or natural vegetation
b. Rubble riprap
c. Wire-enclosed riprap (gabions)
d. Turf reinforcement (non-biodegradable)
2. Rigid Linings:
a. Cast-in-place concrete or asphaltic concrete
b. Soil cement and roller-compacted concrete
c. Grout-filled mattresses
d. Partially grouted riprap
e. Articulated concrete blocks

3.3.1 Flexible Linings


Flexible linings have several advantages compared to rigid linings. They generally are
less expensive, permit infiltration and exfiltration, and can be vegetated to have a natural
appearance. Flow in channels with flexible linings is similar to that found in natural small
channels. Natural conditions offer better habitat opportunities for local flora and fauna. In
many cases, flexible linings are designed to provide only transitional protection against
erosion while vegetation establishes and becomes the permanent lining of the channel;
flexible channel linings are best suited to conditions of moderate shear stresses. Channel
reaches with accelerating or decelerating flow (expansions, contractions, drops, and
backwater) and waves (transitions, flows near critical depth, and shorelines) will require
special analysis and may not be suitable for flexible channel linings.

Chapter 3: Open Channel 3-52


January 2019

Drainage Design Guide


Chapter 3: Open Channel

3.3.1.1 Vegetation
Vegetative linings consist of seeded or sodded grasses placed in and along the channel,
as well as naturally occurring vegetation. Vegetation is one of the most common and most
ideal channel linings for an artificial channel. It stabilizes the body of the channel,
consolidates the soil mass of the bed, checks erosion on the channel surface, and
controls the movement of soil particles along the channel bottom. Vegetative channel
lining also is recognized as a best management practice for stormwater quality design in
highway drainage systems. The slower flow of a vegetated channel helps the uptake of
highway runoff contaminants (particularly suspended sediments) before they leave the
highway right of way and enter streams.

There are conditions for which vegetation may not be acceptable, so you will need to
consider other linings. These conditions include, but are not limited to:

 Standing or continuous flowing water


 Areas which do not receive the regular maintenance necessary to prevent
domination by taller vegetation
 Lack of nutrients and excessive soil drainage
 Areas where sod will be excessively shaded

The Department operates on the premise that, with proper seeding and mulching during
construction, maintenance of most ditches on normal sections and grades can be handled
economically until a growth of grass becomes established. The use of temporary erosion
control measures in ditches with low velocities will provide time for grassing and mulching
to establish a vegetative ditch. When velocities exceed those for bare soils, seeding and
mulching should not be used.

Sodding is recommended when the design velocity exceeds the value permitted for the
bare base soil conditions, but is less than 4 feet per second. Lapped or shingle sod is
recommended when the design velocity exceeds that for sod (4 feet per second), and is
suitable with velocities up to 5.5 feet per second.

3.3.1.2 Other Flexible Linings


Flexible linings usually are less expensive than rigid linings, provide a safer roadside, and
have self-healing qualities that reduce maintenance. They also allow the infiltration and
exfiltration of water.

(A) Rubble Riprap


After grass, rubble riprap is the most common type of flexible lining. It presents a rough
surface that can dissipate energy and mitigate velocity increases. There are two standard

Chapter 3: Open Channel 3-53


January 2019

Drainage Design Guide


Chapter 3: Open Channel

types of rubble riprap. Use ditch lining rubble riprap in standard or typical ditches or
channels. It consists of smaller stone sizes, which reduces construction costs over bank
and shore rubble. Limit bank and shore rubble riprap to uses such as revetments and
linings along stream banks and shorelines where extreme flows or wave action occurs.

Limited right of way and availability of material may restrict the use of this type of flexible
lining. Place rubble riprap on a filter blanket and prepared slope to form a well-graded
mass with a minimum of voids. Riprap and gabion linings can perform in the initial range
of hydraulic conditions where you would use rigid linings. Stones used for riprap and
gabion installations preferably have an angular shape that allow them to interlock. These
linings usually require a filter material between the stone and the underlying soil to prevent
soil washout and migration of fine grained soils. Sometimes you will need a bedding stone
layer to protect the filter fabric from larger stone.

(B) Gabion Mats


Gabions are made of riprap enclosed in a wire container or closed structure that binds
units of the riprap lining together. The wire enclosure normally consists of a rectangular
container made of steel wire woven in a uniform pattern, and reinforced on corners and
edges with heavier wire. The containers are filled with stone, connected together, and
anchored to the channel side slope. The forms of wire-enclosed riprap vary from thin
mattresses to boxlike gabions. Use gabions typically when rubble riprap is either not
available or not large enough to be stable. Although flexible, wire mesh restricts gabion
movement. The wire mesh must provide an adequate service life. If the wire mesh fails,
the individual stones will migrate.

(C) Articulating Concrete Block (ACB) Revetment Systems


ACB systems consist of a precast block matrix connected together by cables. The
articulating properties of the matrix allow the system to accommodate changes in the
ground surface that may occur due to settling. The block configuration varies with the
manufacturer. The systems typically are manufactured in units of multiple precast blocks
that can be lifted easily and placed with construction equipment. HEC-23 and the National
Concrete Masonry Association’s Design Manual for Articulating Concrete Block
Revetment Systems provide guidance for the design of these systems.

(D) Turf Reinforcement


Depending on the application, materials, and method of installation, turf reinforcement
may serve a transitional or long-term function. The concept of turf reinforcement is to
provide a structure to the soil/vegetation matrix that will both assist in the establishment
of vegetation and provide support to mature vegetation. Two types of turf reinforcement
commonly are available: soil/gravel methods and turf reinforcement mats (TRMs).

Chapter 3: Open Channel 3-54


January 2019

Drainage Design Guide


Chapter 3: Open Channel

To create soil/gravel turf reinforcement, you mix gravel mulch into on-site soils and seed
the soil-gravel layer. The rock products industry provides a variety of uniformly graded
gravels for use as mulch and soil stabilization. A gravel/soil mixture provides a non-
degradable lining that is created as part of the soil preparation and is followed by seeding.

A TRM is a non-degradable rolled erosion control product (RECP) composed of UV-


stabilized synthetic fibers, filaments, netting, and/or wire mesh processed into a three-
dimensional matrix. TRMs provide sufficient thickness, strength, and void space to permit
soil filling and establishment of grass roots within the matrix. One limitation to the use of
TRMs is in areas where siltation is a problem. When the ditch is cleaned by maintenance,
it is likely that the geofabric will be snagged and pulled out by the equipment.

3.3.2 Rigid Linings


Rigid linings generally are constructed of concrete, asphalt, or soil-cement pavement
whose smoothness offers a higher capacity for a given cross-sectional area. Higher
velocities, however, create the potential for scour at channel lining transitions from the
rigid lining back to the grass lining. A rigid lining can be destroyed by flow undercutting
the lining, channel headcutting, or the buildup of hydrostatic pressure behind the rigid
surfaces. When properly designed, rigid linings may be appropriate where the channel
width is restricted. Rigid linings are useful in flow zones where high shear stress or rapidly
varied or turbulent flow conditions exist, such as at transitions in channel shape or at an
energy dissipation structure.

Rigid linings are particularly vulnerable to a seasonal rise in the water table that can cause
a static uplift pressure on the lining. If you need a rigid lining in such conditions,
incorporate a reliable system of under drains and weep holes as a part of the channel
design. Evaluate the migration of fine grained soils into filter layers to ensure that the
ground water is being discharged without filter clogging or collapse of the underlying soil.
A related case is the buildup of soil pore pressure behind the lining when the flow depth
in the channel drops quickly. Using watertight joints and backflow preventers on weep
holes can help to reduce the buildup of water behind the lining.

Section 2.4.3.1.2 of the Drainage Manual states that you should consider the potential for
buoyancy due to the uplift water pressure when concrete linings are to be used where
soils may become saturated. The total upward force is equal to the weight of the water
displaced by the channel. The total weight of the lining helps to resist the uplift pressure.
When the weight of the lining is less than the uplift pressure, the channel is unstable.

Chapter 3: Open Channel 3-55


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Acceptable countermeasures include:

 Increasing the thickness of the lining to add additional weight


 For sub-critical flow conditions, specifying weep holes at appropriate intervals in
the channel bottom to relieve the upward pressure on the channel
 For super-critical flow conditions, using sub-drains in lieu of weep holes

3.3.2.1 Cast-in-Place Concrete


Refer to Standard Plans, Index 524-001 for typical ditch pavement details. Asphalt linings
have limited use since routine maintenance activities often damage or destroy them. Use
filter fabric to prevent soil loss through pavement cracks.

Despite the non-erodible nature of concrete linings, they are susceptible to failure from
foundation instability. The major cause of failure is undermining that can occur in a
number of ways. Inadequate erosion protection at the outfall, at the channel edges, and
on bends can initiate undermining by allowing water to carry away the foundation material
and leaving the channel to break apart. Concrete linings also may break up and
deteriorate due to conditions such as a high water table or swelling soils that exert an
uplift pressure on the lining. When a rigid lining breaks and displaces upward, the lining
continues to move due to dynamic uplift and drag forces. The broken lining typically forms
large, flat slabs that are particularly susceptible to these forces.

3.3.2.2 Grout-Filled Mattresses


Grout-filled mattresses are the result of pumping a concrete mix into fabric envelopes or
cases. The advantage of using grout-filled mattresses is that they reduce construction
time by eliminating the need for wooden forms and expensive lifting machines and also
allow the concrete to be pumped and cured below the water line.

Filter point grout-filled mattresses consist of a dual wall fabric that is injected with
concrete. This type of grout-filled mattress is characterized by a deeply cobbled surface.
The filter points woven into the fabric provide a means for groundwater to escape and to
provide release for the hydrostatic pressure. Filter point fabrics provide a higher
coefficient of friction to promote energy dissipation.

Note: As of May 2016, FDOT does not have standard specifications for grout-filled
mattresses. A technical specification will need to be prepared if grout-filled mattresses
are proposed for a project.

Chapter 3: Open Channel 3-56


January 2019

Drainage Design Guide


Chapter 3: Open Channel

3.3.3 Velocity and Shear Stress Limitations


HEC-15 provides a detailed presentation of stable channel design concepts for roadside
and median channels. This section provides a brief summary of significant concepts.

Stable channel design concepts provide a means of evaluating and defining channel
configurations that will perform within acceptable limits of stability. Most highway drainage
channels cannot tolerate bank instability and lateral migration. When the material forming
the channel boundary effectively resists the erosive forces of the flow, then you have
achieved stability. You can apply principles of rigid boundary hydraulics to evaluate this
type of system.

Apply both velocity and tractive force methods to help determine channel stability.
Permissible velocity procedures are empirical in nature, so they have been used to design
numerous channels in Florida and throughout the world. However, tractive force methods
consider actual physical processes occurring at the channel boundary and represent a
more realistic model of the detachment and erosion processes.

The hydrodynamic force that water flowing in a channel creates causes a shear stress on
the channel bottom. The bed material, in turn, resists this shear stress by developing a
tractive force. Tractive force theory states that the flow-induced shear stress should not
produce a force greater than the tractive resisting force of the bed material. This tractive
resisting force of the bed material creates the permissible or critical shear stress of the
bed material. In a uniform flow, the shear stress is equal to the effective component of the
gravitational force acting on the body of water parallel to the channel bottom. The average
shear stress is equal to:

=RS (3.3-1)

where:
= Average shear stress, in pounds per square feet
= Unit weight of water,62.4 lb/ft3
R= Hydraulic radius, in feet
S= Average bed slope or energy slope, in feet per feet

The maximum shear stress for a straight channel occurs on the channel bed and is less
than or equal to the shear stress at maximum depth. Compute the maximum shear stress
as follows:

d = d S (3.3-2)

Chapter 3: Open Channel 3-57


January 2019

Drainage Design Guide


Chapter 3: Open Channel

where:
d = Maximum shear stress, in pounds per square feet
d= Maximum depth of flow, in feet
S= Channel bottom slope, in feet per feet

Velocity limitations for artificial open channels should be consistent with stability
requirements for the selected channel lining. As indicated above, use seed and mulch
only when the design velocity does not exceed the allowable velocity for bare soil. Table
2.3 of the Drainage Manual presents maximum shear stress values and allowable
velocities for different soils. When design velocities exceed those acceptable for bare soil,
sod, or lapped sod, consider flexible or rigid linings. Table 2.4 of the Drainage Manual
summarizes maximum velocities for these lining types.

Side Slope Stability


The shear stress on the channel sides generally is less than the maximum shear stress
calculated on the channel bottom, but you should consider this issue when determining
the height of a channel lining along the side slope of the channel. The maximum shear
stress on the side of a channel is given by:

s= d (3.3-3)

where:
s = Side shear stress on the channel, in pounds per square feet
 = Ratio of channel side to bottom shear stress
d = Shear stress in channel at maximum depth, in pounds per square feet

The value K1 depends on the size and shape of the channel. For parabolic channels, the
shear stress at any point on the side slope is related to the depth at that point and you
can calculate it using Equation 3.3-2. For trapezoidal and triangular channels, K1 is based
on the horizontal dimension 1: Z (V: H) of the side slopes.

K1 = 0.77 Z ≤ 1.5
K1 = 0.066Z + 0.67 1.5 < Z < 5
K1 = 1.0 5≤Z

Avoid using side slopes steeper than 1:3 for flexible linings other than riprap or gabions
because of the potential for erosion at the side slopes. Steep side slopes are allowable
within a channel if cohesive soil conditions exist.

Maintenance Considerations
Also consider maintenance of the channel when choosing a channel lining. The channel
will need to be accessible by mowers and trucks.

Chapter 3: Open Channel 3-58


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Mowing

Side slopes of vegetated channels will need to be traversable for mowing equipment and
crews. The maximum traversable slope for this equipment is 1:4.

Access Across Channel

If there is rubble riprap lining the channel and a vegetated buffer on the backside of the
channel along the right of way, the irregularity of the riprap typically prevents access. In
this situation, it may become impractical to maintain the vegetation.

3.3.4 Application Guidance for Some Common Channel Linings


3.3.4.1 Rubble Riprap
Types
 Ditch Lining—Flexible layer or facing of rock placed on a filter blanket and
prepared slope used to line a ditch or channel for protection from erosion.
 Bank and Shore—Flexible layer or facing of rock placed on a bank or shore to
prevent erosion or scour of the embankment or a structure.

What is its purpose?


Use rubble riprap in channels, along embankments, or around structures that are
vulnerable to erosion or scour.

Where and how is it commonly used?


 Ditch Lining—In this case, use rubble riprap to line ditches and channels to
protect slopes from erosion.
 Bank and Shore—In this case, use it as a flexible revetment to line banks and
shores subject to erosion.

When should it be installed?


 Ditch Lining—Install rubble riprap in channels with moderate shear stresses. To
prevent uplifting forces on the lining, the filter requires adequate permeability.
 Bank and Shore—Use rubble riprap to protect banks or shores with flows that
generally are greater than 50 ft3/s or that are subject to wave action.

When should it not be installed?


 Bank and Shore—Do not install rubble riprap when ditch lining methods are
applicable.

Chapter 3: Open Channel 3-59


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Advantages and disadvantages


ADVANTAGES
 Flexible
 Not weakened by minor shifting caused by settlement
 Easily repaired by additional rock placement
 Simple construction method
 Recoverable/reusable
 Long-term or temporary installations

DISADVANTAGES
 Hauling and installation costs
 Prohibits maintenance equipment from traversing channels
 If hand placement is required, then labor is intensive
 Vegetation growth can hinder inspections

North Carolina Erosion and Sediment Planning and Control Manual

Figure 3.3-1: Riprap-lined Channel Cross Sections

Chapter 3: Open Channel 3-60


January 2019

Drainage Design Guide


Chapter 3: Open Channel

3.3.4.2 Grout-Filled Mattresses


Types
Grout-filled mattresses for concrete with filtering points that provide for the relief of
hydrostatic pressures.

What is the purpose?


Use grout-filled mattresses—filter point or articulating—for slopes or areas that are
subject to severe to moderate erosion problems.

Where and how are they commonly used?


Use grout-filled mattresses in ditches, channels, canals, streams, rivers, ponds, lakes,
reservoirs, marinas, and ports/harbors to reduce the impact of erosion.

When should they be installed?


Install grout-filled mattresses where there are moderate to severe erosion problems and
where the channel is subjected to hydrostatic uplift pressures. Also, install these where
there is a need to allow water to permeate into the soil and not remain wet.

When should they not be installed?


Do not use grout-filled mattresses in ditches or channels that are subject to changes in
soil conditions such as erosion under the mat or consolidation.

Advantages and disadvantages


ADVANTAGES
 Adapts easily to contours
 Easy to install
 Permeable
 Reduces uplift pressure
 Can be installed under the water line

DISADVANTAGES
 Needs to be installed on a prepared slope
 Not aesthetically pleasing
 Easily undermined if not toed properly

Chapter 3: Open Channel 3-61


January 2019

Drainage Design Guide


Chapter 3: Open Channel

(Source: http://www.fabriform1.com)
Construction Techniques, Inc.

Figure 3.3-2: Grout-Filled Mattress with Filter Point Linings

3.3.4.3 Gabions
Types
 Gabion Mats—Wire mesh mats filled with stones
 Gabion Baskets—Wire mesh baskets filled with stones

What is the purpose?


Rock-filled baskets or mattresses that are used to line large ditches, channels, canals,
and coastal shores for stabilization and protection.

Where and how are they commonly used?


Gabion Mats—Use gabion mats in ditches, channels, canals, streams, rivers, ponds,
lakes, reservoirs, marinas, and ports/harbors to reduce the impact of erosion.

When should they be installed?


Gabion Mats—Install gabion mats in large areas where there are moderate to severe
erosion problems due to extreme velocities. Also where there is a need to allow water to
permeate into the soil and not remain wet.

When should they not be installed?


Gabion Baskets—Small areas subject to low velocities and when a temporary situation
exists.

Advantages and disadvantages


ADVANTAGES
 Protects seed mix from eroding when used
 Permeable
 Increases retention of soil moisture
 Permits the growth of vegetation
 Able to span minor pockets of bank subsidence without failure

Chapter 3: Open Channel 3-62


January 2019

Drainage Design Guide


Chapter 3: Open Channel

DISADVANTAGES
 Cost of installation
 Susceptibility of the wire baskets to corrosion and abrasion damage
 More difficult and expensive to repair
 Less flexible than standard riprap

Chapter 3: Open Channel 3-63


January 2019

Drainage Design Guide


Chapter 3: Open Channel

(Source: http://www.gabions.net/downloads.html)
Modular Gabion Systems, a division of C.E. Shepherd Company

Figure 3.3-3: Gabion Dimensions


Chapter 3: Open Channel 3-64
January 2019

Drainage Design Guide


Chapter 3: Open Channel

(Source: http://www.gabions.net/downloads.html)
Modular Gabion Systems, a division of C.E. Shepherd Company

Figure 3.3-4. Gabion Binding


Chapter 3: Open Channel 3-65
January 2019

Drainage Design Guide


Chapter 3: Open Channel

3.3.4.4 Soil Stabilizers


Types
 Turf Reinforcement Mats—A long-term non-degradable mat composed of UV
stabilized synthetic fibers, nettings, and/or filaments.
 Erosion Control Blankets—A temporary degradable mat composed of
processed natural or polymer fibers mechanically, structurally, or chemically bound
together to form a continuous matrix.

What is the purpose?


To protect disturbed slopes and channels from wind and water erosion. The blanket
materials are natural materials, such as straw, wood excelsior, coconut, or are geotextile
synthetic woven materials, such as polypropylene.

Where and how are they commonly used?


 Turf Reinforcement Mats—Use them on ditch slopes and fill slopes to reduce the
impact of erosion for long periods of construction.
 Erosion Control Blankets—Use them on ditch slopes and fill slopes to reduce
the impact of erosion during short periods of construction.

When should they be installed?


 Turf Reinforcement Mats—Where there are low velocities of flow
 Erosion Control Blankets—Where there are low velocities of flow and where
there are sensitive environmental areas

When should they not be installed?


 Turf Reinforcement Mats—Do not install for permanent situations and where
there are high velocities of flow.
 Erosion Control Blankets—Do not install for permanent situations and where
there are high velocities of flow.

Advantages and disadvantages


ADVANTAGES
 Adapts easily to contours
 Easy to install
 Permeable
 Reduced uplift pressure
DISADVANTAGES
 Cost
 Maintenance equipment can damage or pull out

Chapter 3: Open Channel 3-66


January 2019

Drainage Design Guide


Chapter 3: Open Channel

(Source: http://propexglobal.com/) (Source: http://propexglobal.com/)


Propex Geosynthetics Propex Geosynthetics

Figure 3.3-5: Erosion Control Mat in Channel Figure 3.3-6: Initial Anchor
(Downstream)

(Source: http://propexglobal.com/)
Propex Geosynthetics

Figure 3.3-7: Longitudinal Anchor Trench Detail (Trapezoidal Channel)

Chapter 3: Open Channel 3-67


January 2019

Drainage Design Guide


Chapter 3: Open Channel

3.4 DRAINAGE CONNECTION PERMITTING AND MAINTENANCE


CONCERNS
3.4.1 Drainage Connection Permitting
Adjacent property owners must obtain a Drainage Connection Permit from FDOT
according to Section 334.044(15), Florida Statute (F.S.), Chapter 14-86, Florida
Administrative Code (F.A.C)., Rules of the Department of Transportation, when developing
their property. In general terms, the Drainage Connection Permit ensures that the
development will not overload the Department’s stormwater conveyance systems and
cause flooding on either the roadway or other downstream properties. For more
information on Drainage Connection Permits, refer to the Drainage Connection Permitting
Handbook. This section will discuss several aspects of the Department’s ditches that you
should consider during the Drainage Connection Permitting process.

3.4.1.1 Roadside Ditch Impacts


Discharges to the roadside ditch from the proposed development will be limited by the
Permit so that the ditch flow will not be increased. However, the proposed development
can physically impact the roadside ditch by placing or widening driveways to the property
or by widening the roadway to add turn lanes.

If the roadside ditch is a linear treatment pond, then any reduction in the volume of the
ditch could violate the conditions of the permit obtained for the facility. The simplest way
to resolve this issue is to rework the ditch so that any volume lost as a result of the
development is replaced. This may require that the property owner donate some property
to the Department to provide an area to rework the ditch.

Even if the roadside ditch is not a linear treatment facility, you must maintain the capacity
of the ditch. Include a side drain to convey the ditch flow from one side of the turnout to
the other, unless the turnout is located at a high point in the ditch and the flow is away
from the turnout in both directions. An added turn lane may require that the roadside ditch
be relocated. The relocated portion of the ditch should have the same capacity or more
than the existing ditch. If the existing right of way is not wide enough to accommodate the
relocated ditch, then right of way may need to be donated to FDOT for the ditch. A turnout
requiring a side drain and a turn lane requiring donated right of way for the ditch relocation
are shown in Figure 3.4-1.

Chapter 3: Open Channel 3-68


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Figure 3.4-1: Effect of Adjacent Development on a Roadside Ditch

In some cases, the developer may need to add a left-turn lane. Widening the road to
accommodate the left-turn lane also may affect the ditch on the opposite side of the road
from the development. Often, the developer will not own the property on both sides of the
road. In this case, the roadside ditches and roadway must be redesigned to accommodate
the new turn lanes in such a way as to require donated right of way on the new
development’s side of the road.

The flow lines of the side drain should match the existing ditch. Also ensure that the flow
lines of the new side drain are higher than the next side drain downstream and lower than
the next side drain upstream to avoid temporary ponding in the ditch.

Make sure to size the side drain properly. You can make some judgments about the size
of the pipe by looking at the side drains upstream and downstream of the new drive.
Analyze the side drain to ensure the new pipe does not cause the water levels to pop out
of the ditch. In some cases, you can obtain the design discharge for the ditch from the old
plans for the roadway. Or you can calculate the flow by determining the drainage area
and performing the proper hydrologic calculations; typically, the Rational Equation. You
can find more details on these hydrology calculations in Chapter 2. Calculate the losses
through the pipe using methods given in Chapter 4. Additional sizing considerations are
discussed in Section 3.2.8.

When adding new side drains, another consideration is the proximity of other existing side
drains. If side drains are too close to each other, then the hydraulic losses can be too
large. The general requirement is that the end sections of two side drains in series should
be at least 25 feet apart. If the distance is less than 25 feet, then you should enclose the
area and add an inlet to collect the runoff from the area between the driveways.

Chapter 3: Open Channel 3-69


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Evaluate potential erosion at the infall point of the connection, especially for pipe
connections. Chapter 4 explains how to calculate the outlet velocity from a pipe. Refer to
Section 3.3 for channel linings. You can find outlet erosion protection criteria in the
Drainage Manual.

3.4.1.2 Median Ditch Impacts


A new development can impact the median ditch if the Department allows a new median
opening or left-turn lane.

Unless you can place a new median opening at the high point in the median ditch, or
close enough to the high point that it is possible to regrade the ditch to flow away from
the new median opening in both directions, then the new opening will block the flow in
the ditch. Figure 3.4-2 shows a typical situation where there is an existing median opening
at the high point in the median ditch and the ditch flows to a median drain, which consists
of a ditch bottom inlet, pipe, and endwall. The median drain discharges runoff from the
median to keep the median from filling with water and spilling across the roadway.

Figure 3.4-2: Existing Median Ditch

If you add a new median opening to accommodate an adjacent development, the opening
may block the flow in the median ditch. Include a new drainage structure with the opening
to discharge the flow from the median. Figure 3.4-3 shows a side drain included to convey
the ditch flow from one side of the new median opening to the other. This often will be the
most economical method to provide adequate drainage for the median. However, in many
cases, the median ditch will be too shallow and the side drain will not have adequate
cover over the pipe. Refer to Appendix C of the Drainage Manual for the minimum cover
needed over the pipe.

Chapter 3: Open Channel 3-70


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Figure 3.4-3: New Median Opening with Side Drain

Figure 3.4-4 includes a new median drain to accommodate the median flow. If you choose
to use this option, check the capacity of the roadside ditch with the added discharge from
the median. Unless you jack and bore the pipe, the existing pavement would have to be
cut and patched to install the pipe. Make sure to consider the cutting and patching
operations in maintenance of traffic plans.

Figure 3.4-4: New Median Drain

Another option that might avoid the expense of jacking and boring or the concerns of
cutting and patching the existing roadway is shown in Figure 3.4-5. You could connect
the new ditch bottom inlet (DBI) to the existing median drain with a pipe beneath the new
median opening.

Chapter 3: Open Channel 3-71


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Figure 3.4-5: New Median Drainage System

Adding a turn lane in the median often will reduce the size of the median ditch adjacent
to the new turn lane. Check the reduced ditch for capacity, and add extra median drainage
structures if needed. Super-elevated roadways that drain to the median can worsen the
capacity problems in areas where the ditch has been reduced.

3.4.1.3 Outfall Ditch Impacts


Requested connections or crossing may physically impact outfall ditches. Usually, the
permitted flow will not be greater than the existing flow rate because of the requirements
of the connection permit. However, you need to evaluate losses associated with the
physical impacts to ensure there is no compromise to the capacity of the outfall ditch.

Overland flow connections can cause bank erosion and sloughing if the flow becomes
concentrated. To avoid this problem, use point connections through pipes or ditches.
Erosion problems also can occur at the connections to an outfall ditch. Refer to Section
3.4.1.1 for guidance to protect the infall point.

3.4.2 Maintenance Concerns

3.4.2.1 Ditch Closures


Residents or other property owners occasionally will request that the roadside ditch in
front of their property be filled and replaced with a pipe system. Piping a ditch can increase
the energy loss and reduce infiltration. Under storm conditions, open ditching is an
efficient method of accommodating a significantly greater quantity of drainage than a pipe.
Therefore, any piping or filling of a roadside ditch generally is of no benefit to the
Department and may reduce operational and maintenance aspects of the road.

Chapter 3: Open Channel 3-72


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Drainage connection applicants should perform a hydraulic assessment to determine


ditch piping or filling impacts on the area drainage system. These impacts should adhere
to Rule 14-86 requirements, as consistent with the Drainage Manual. Unless you acquire
flood rights, any increase over pre-development stages should not change land use
values significantly.

Do not consider filling an open ditch if the basis for the modification is for aesthetic
purposes, for landscaping, or to benefit the abutting private property owner only. Table
3.4-1 lists criteria and other considerations for converting existing drainage ditches to
closed drainage systems.

Table 3.4-1
Capacity of Closed System

Criteria Comments
Design Storms: Primary considerations:
The more stringent of:  Minimize adverse impact on
 Rule 14-86, F.A.C. Storms: Department & other facility users
 Original Ditch Design Storms:  Maximize capacity of facility
 Drainage Manual Design Storms:  Maximize life of facility
o Evacuation route? o Avoid need to reconstruct
o Upstream owner constraints? for later foreseeable
 Potential for flooding projects
upstream?  Minimize maintenance cost
o Downstream constraints?
 Tailwater
 Planned work program improvements:
Pipe Size: Check various scenarios and use the
The more stringent of: criteria that most satisfies the
 Rule 14-86, F.A.C. Criteria: Department’s interests.
 Original Ditch Design Criteria:
 Drainage Manual Criteria:
 Future Work Program Requirements:

Chapter 3: Open Channel 3-73


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Table 3.4-1 (continued)


Capacity of Closed System (continued)
Criteria Comments
Method: Do not rely solely on the size of existing
Prove that the headwater elevation for the upstream systems for designing capacity of
design storms shall not be increased ditch systems downstream. While
immediately upstream of the proposed knowledge of upstream systems is useful in
system. many ways, these existing systems:
 May be undersized due to:
Base design on hydrologic conditions in the o Design errors
field, not the size of existing pipe systems. o Under estimated watershed
area
Base design on condition that entire length of o Subsequent land
the ditch will eventually have a closed system. development activity
o Subsequent system changes
or diversion
 May not reflect current design
standards
 May not be adequate for current or
future needs
o Existing flooding conditions
o Future road improvements
Other considerations: The applicant usually hopes to reduce the
Remember that the Department owns not only Department’s easement area by closing the
the current capacity of its outfall easements, open ditch with pipe or other structures.
but also the right to use any potential excess
capacity available in the outfall. This usually represents a false economy
when one adds the requirements necessary
Any proposed piped outfall must be sized for to maintain the closed system at minimum
the Design Frequency noted in Chapter 2 of the expense.
Drainage Manual.
Oftentimes, you can eliminate or greatly
Select solutions that maximize preservation of reduce major risk of damage due to system
the Department’s ability to expand its system to failure by careful attention to the failure
the full use of its facility for future needs. mode and addition of details to re-route
overflows or provide protective measures
Consider the consequences that result when such as curbs, berms, emergency
the proposed system fails and make any spillways, etc.
reasonable adjustments to minimize damage
and liability for the Department.

Chapter 3: Open Channel 3-74


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Table 3.4-1 (continued)


Work Program
Criteria Comments
Considerations: The possibility exists that the applicant can
 In Work Program: simply build the outfall already under
o If already designed & approved design by the Department, especially if the
– use the design applicant cannot wait for the Department’s
o If not designed – coordinate future construction job to complete the
design for approval by DOT work.
project engineer
 Not in Work Program:
o Route design submittal for
review and approval by District
Drainage Engineer among
others

Erosion Control
Considerations: May result in failure of the pipe outfall
 Erosion at outlet system.
 Erosion when flows exceed system
capacity Possible turbid discharge downstream.
 Soils
 Flow velocity
 Slopes
Methods:
 Drainage Manual
 Erosion and Sediment Control Designer
and Reviewer Manual
 Protective measures
o Structural solutions
o Non-structural methods

Maintenance
Responsibility: Define this carefully in the agreement.
 Applicant (local government)
responsible
o When concession needed from
Department in negotiation
o When special structures require
more maintenance attention or
expense
 DOT responsible
o At DOT discretion

Chapter 3: Open Channel 3-75


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Table 3.4-1 (continued)


Maintenance (continued)
Criteria Comments
Considerations: Consider these factors when negotiating
 Reasonable & Safe Access the terms of agreement.
o For equipment
o For personnel Remember: If the new facility cannot be
o For operations – spoil, staging, reasonably maintained in a safe and cost
etc. effective manner, then perhaps the
 Other facilities in easement easement should remain an open ditch.
o Above ground - trees, fences,
sheds, etc.
o Underground - utilities,
drainage, etc.
 Potential to damage adjacent facilities
o Above ground structures,
buildings, etc.
o Overhanging structures,
utilities, etc.
 Limitations:
o Depth of work - shoring
needed?
o Groundwater
Right-of-way
Considerations: Consult with right-of-way attorney to
 Additional right-of-way required: determine:
o To maintain access  the appropriate style of easement
o To enable maintenance  relation to downstream owners not
o To minimize Cost of involved in the transaction
Maintenance o Where to end the easement
o To preserve or secure drainage when drainage exits
rights applicant’s property and
 Donation of right-of-way falls onto another person’s
 Reduction of right-of-way: property?
o Only when fully justified  special terms to add into the
o Must meet Drainage Manual easement document
requirements for dimension,
etc.

Chapter 3: Open Channel 3-76


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Table 3.4.1 (continued)


Permitting
Criteria Comments
Document: A Drainage Connection Permit is not the
 Contractual Agreement appropriate form for approval of this
 Easement Agreement category of work, unless the work is
 Easement Donation / Exchange performed as part of a larger scope of
 Drainage Connection Permit property improvements that require the
permit and there is no need to alter the
existing easement in any way.

A contractual agreement with appropriate


terms and conditions is the preferred
method of approval.
Process: Some typical contract terms:
 If easement relocation or exchange  Review and approval of plans
required:  Party responsible for maintenance
o Follow “Property Management  Failure-to-perform provisions
Related Reconstruction  Responsibility to obtain all required
Process” chart permits
 If no change needed to existing  Review of plans
easement :  Notice of changes
1. Consult early with Legal  As-built plans & computations
Department to determine form of  Final certification by engineer
agreement  May waive need for other permits, if
2. Perform review proposed design to practicable
determine any special conditions or
 Other conditions as needed
terms required in the agreement
3. Legal Department to draft
agreement
4. Maintenance to review draft
agreement and resolve any issues.
5. Deliver agreement to applicant for
signature.
6. Obtain Department signature
7. Administer terms of agreement
Construction
Considerations:
 Pre-construction meeting
 All permits in hand
 Erosion control measures in place
 Oversight & Inspection

Chapter 3: Open Channel 3-77


January 2019

Drainage Design Guide


Chapter 3: Open Channel

Table 3.4-1 (continued)


Construction (continued)
Inspection:
 Administer contract
 Obtain approval from engineer for
changes
 Erosion control
Acceptance:
 Follow contract terms for completion of
contract
 File as-built plans & design
computations

3.4.2.2 Acquisition of Ditches from Local Ownership


When roadways pass from local ownership to FDOT, it is not unusual for issues to arise.
Often, the roadside ditches on these roadways do not meet FDOT standards. They often
were designed for a lesser design frequency and do not contain enough capacity. Other
ditches have substandard slopes located within the clear zone. When safety concerns
force these roadways to be updated, evaluate the existing conditions to bring the ditches
up to current standards.

In some cases, there may be enough right of way available to reconstruct the ditch to
standards. More frequently, though, right of way is not sufficient to provide these
upgrades. Then, it may be practical to purchase additional right of way or drainage
easements in which to upgrade the current ditch system. If additional right of way proves
to be too costly, consider a closed system with a series of inlets and storm drain pipes.
The least ideal but often unavoidable option will consist of obtaining exceptions or
variances of the current standards for the existing ditch.

3.4.2.3 Addition of Sidewalks to Roadway Projects


In an ongoing attempt to connect communities with pedestrian walkways, existing
roadways often have sidewalks added. The sidewalks often are located outside of the
existing ditch system along the right-of-way line. When designing these sidewalks, ensure
that the sidewalk does not impede flow from offsite runoff. Place it so that offsite runoff
can sheet flow over the sidewalk into the existing ditch or that the system can collect
runoff and pipe it under the sidewalk into the ditch or an existing storm drain system. In
many cases, you can construct a simple pedestrian bridge to cross over existing ditches
without impacts to the ditch.

Chapter 3: Open Channel 3-78


January 2019

Drainage Design Guide


Chapter 4: Culvert

CHAPTER 4: CULVERT

4. CULVERT ............................................................................................................. 4-1

4.1 General ................................................................................................................. 4-1


4.1.1 Cross Drain Design ........................................................................................ 4-1
4.1.2 Scour Estimate............................................................................................... 4-1
4.1.3 Flood Definition .............................................................................................. 4-2

4.2 Design Frequency ............................................................................................. 4-11

4.3 Backwater .......................................................................................................... 4-12


4.3.1 Backwater Consistent with the Flood Insurance Study Requirements ......... 4-12

4.4 Tailwater ............................................................................................................ 4-13

4.5 Hydraulic Analysis ............................................................................................ 4-13

4.6 Specific Standards Relating To All Cross Drains Except Bridges................ 4-27
4.6.1 Culvert Materials .......................................................................................... 4-27
4.6.2 Scour Estimates ........................................................................................... 4-27

4.7 Recommended Design Procedure ................................................................... 4-28


4.7.1 Culvert Extensions ....................................................................................... 4-28
4.7.2 Small Cross Drains ...................................................................................... 4-43
4.7.3 Large Cross Drains ...................................................................................... 4-47

Chapter 4: Culvert 4-i


January 2019
January 2019

Drainage Design Guide


Chapter 4: Culvert

4. CULVERT

4.1 GENERAL
4.1.1 Cross Drain Design
Section 4.2 of the Drainage Manual states, "All cross drains shall be designed to have
sufficient hydraulic capacity to convey the selected design frequency flood without
damage to the structure and approach embankments, with due consideration to the
effects of greater floods." This requires evaluation of the following:
Backwater
Refer to Section 4.3 of this design guide and Section 4.4 of the Drainage Manual.

Tailwater
Refer to Section 4.4 of this design guide and Section 4.5 of the Drainage Manual.

Scour
Refer to Sections 4.1.2 and 4.6.2 of this design guide and Section 4.9.2 of the
Drainage Manual.

You may need to perform a risk analysis to evaluate damage to structures and/or
embankments caused by backwater and/or scour. Refer to Appendix G, Risk Evaluations.

4.1.2 Scour Estimate


When producing scour estimates for bridge culvert foundation designs, it is best not to
use the methods in FHWA’S HEC-18. Instead, consider the outlet velocity and
degradation of the stream, discussed in Section 4.6.2 of this document.

To use bridge culverts with no bottom slab and toe wall, you need to get the following
approval/evaluation:

a) Prior approval from the District Drainage Engineer.

b) An analysis of the degradation that could take place through the bridge culvert.
This would require you to recommend the toe wall depths of the bridge culvert and
the need for scour protection for the design-year frequency, 100-year frequency,
and 500-year frequency.

Chapter 4: Culvert 4-1


January 2019

Drainage Design Guide


Chapter 4: Culvert

4.1.3 Flood Definition


Design Flood
The “design flood” is defined as the flood or storm surge associated with the
probability of exceedance (frequency) selected for the design of a highway
encroachment. This frequency, known also as the "design-year frequency," is
discussed in Section 4.2.

Base Flood
The “base flood” (100-year frequency flood event) is defined as the flood or storm
surge having a 1-percent chance of being exceeded in any given year. The base
flood is the standard in Federal Emergency Management Agency (FEMA) flood
insurance studies and many agencies have adopted it to comply with regulatory
requirements.

Greatest Flood
The “greatest flood” (500-year frequency flood event) is defined as the flood or
storm surge having a 0.2-percent chance of being exceeded in any given year.
This event is used to define the possible consequences of a flood occurrence
significantly greater than the 1-percent flood event. While it is seldom possible to
compute the discharge for the 500-year frequency flood with the same accuracy
that you would compute the discharge for the base flood, it serves to draw attention
to the fact that floods greater than the base flood can occur. In some cases, FEMA
and other agencies compute the 500-year frequency flood.

Overtopping Flood
The “overtopping flood” is described by the probability of exceedance and water
surface elevation at which water begins to flow over the highway, a watershed
divide, or through structure(s) providing for emergency relief.

The overtopping flood is of particular interest because it will indicate one of the
following:

1. When a highway will be inundated

2. The limit (stage) at which the highway, ditch, or some other control point will
act as a significant flood relief for the structure of interest

Carefully compare roadside ditch elevations with respect to the water surface
elevation for the structure being designed or analyzed. There may be instances
where the ditch elevation will provide significant relief to the structure for a certain
flood. This ditch elevation will define the overtopping flood stage.

Example 4.1-1 shows how the overtopping flood is determined.

Chapter 4: Culvert 4-2


January 2019

Drainage Design Guide


Chapter 4: Culvert

Example 4.1-1—Computing the Overtopping Flood

Given the information below, determine the discharge and frequency for the overtopping
flood.

Q (25) = 31 ft3/sec Stage (25) = 134.3 ft.


Q (100) = 55 ft3/sec Stage (100) = 139.0 ft.
Q (Overtopping) =? Stage (Overtopping) = 140.9 ft.

Solution

Step 1:

a. To determine the overtopping discharge, plot stage versus discharge


on algebraic scale graph paper for the 25-year and 100-year floods,
as shown on Figure 4.1-1.

Note: Graphical estimation methods are explained in Hydraulic


Design Series No. 2 (HDS-2), Publication No. FHWA-NHI-
02-001, October 2002.

b. Draw the best-fit line through these points.

c. Knowing what the overtopping stage is, you can conservatively


approximate the overtopping discharge. The overtopping discharge
was found to be 64 ft3/sec.

Note: For stages above overtopping, the overtopping flow can


provide significant relief. The stage versus discharge
relationship usually flattens out after overtopping.

Step 2:

a. To determine the overtopping frequency, plot frequency versus


discharge on log-normal probability paper for the 25-year and 100-
year floods, as shown in Figure 4.1-2.

b. Draw the best-fit line through these points.

c. Knowing the overtopping discharge from Step 1c, you can determine
the probability of the overtopping flood being exceeded in any year.
In this case, the probability is 0.65 percent. This corresponds to a
frequency of 154 years (i.e., 100/0.65).

Chapter 4: Culvert 4-3


January 2019

Drainage Design Guide


Chapter 4: Culvert

Figure 4.1-1: Example I - Computing Overtopping Flood (cont.)

Chapter 4: Culvert 4-4


January 2019

Drainage Design Guide


Chapter 4: Culvert

Figure 4.1-2: Example I - Computing Overtopping Flood (cont.)

Chapter 4: Culvert 4-5


January 2019

Drainage Design Guide


Chapter 4: Culvert

Flood Data Summary Box

For culverts other than bridge culverts, include hydraulic data in a Flood Data
Summary Box similar to the example shown in Figure 4.1-3. Include these data for
those conditions discussed in FDM 305.
STRUCTURE

DESIGN FLOOD BASE FLOOD


STATION

OVERTOPPING GREATEST
FLOOD FLOOD
NO.

___% PROB. ____ YR FREQ ___% PROB. ____ YR FREQ

PROB FREQ. PROB FREQ


DISCHARGE STAGE DISCHARGE STAGE DISCHARGE STAGE DISCHARGE STAGE
% YR % YR

Note: The hydraulic data are shown for informational purposes only, to indicate the flood discharges and water
surface elevations that may be anticipated in any given year. These data were generated using highly variable
factors determined by a study of the watershed. Many judgments and assumptions are required to establish
these factors. The resultant hydraulic data are sensitive to changes, particularly of antecedent conditions,
urbanization, channelization, and land use. Users of these data are cautioned against the assumption of
precision, which cannot be attained. Discharges are in cubic feet per second and stages are in feet.

Definitions:
Design Flood The flood selected by FDOT to be utilized to assure a standard level of hydraulic
performance
Base Flood: The flood having a 1-percent chance of being exceeded in any given year (100-year
frequency)
Overtopping Flood: The flood that causes water to flow over the highway, over a watershed divide, or through
emergency relief structures
Greatest Flood: The most severe flood that can be predicted, where overtopping is not practicable;
normally, one with a 0.2-percent chance of being exceeded in any given year (500-year
frequency)

Figure 4.1-3: Flood Data Summary Box

Chapter 4: Culvert 4-6


January 2019

Drainage Design Guide


Chapter 4: Culvert

Fill out the hydraulic flood data sheet according to the Federal Aid Policy Guide (23 CFR
650A). You can find this policy guide at
http://www.fhwa.dot.gov/legsregs/directives/fapg/cfr0650a.htm. In general, the following
applies.

a. If the overtopping flood is less than the standard design frequency, perform a
risk assessment to define the design flood as the overtopping flood. Fill out the
information for the design flood, base flood, and overtopping flood.

b. If the overtopping flood is between the standard design frequency and the base
flood (100-year flood), then fill out the information for the design flood, base
flood, and overtopping flood.

c. If the overtopping flood is between the base flood (100-year flood) and the
greatest flood (500-year flood), then fill out the information for the design flood,
base flood, and overtopping flood.

d. If the overtopping flood is larger than the greatest flood (500-year flood), then
fill out the information for the design flood, base flood, and greatest flood.

Example 4.1-2 shows you how to complete the Flood Data Summary Box when the
overtopping flood is less than the greatest flood (500-year flood).

Example 4.1-3 shows you how to complete the Flood Data Summary Box when the
overtopping flood occurs at a 10-year frequency.

Example 4.1-2—Completing the Flood Data Summary Box

Referring back to Example 4.1-1, assume the design flood is the 25-year frequency. Fill
out the Flood Data Summary Box.

Solution
Since the overtopping flood is between the base flood (100-year flood) and the greatest
flood (500-year flood), then fill out the information for the design flood, base flood, and
overtopping flood.

Q (25) = 31 ft3/sec
Stage (25) = 134.3 ft.

Q (100) = 44 ft3/sec
Stage (100) = 136.4 ft.

Chapter 4: Culvert 4-7


January 2019

Drainage Design Guide


Chapter 4: Culvert

Q (Overtopping) = 64 ft3/sec
Stage (Overtopping) = 140.9 ft.

Put these values in the corresponding column, as shown in Figure 4.1-4. From Example
4.1-1, the overtopping flood was found to have a 0.65 percent chance of being exceeded
in any year, or a frequency of 154 years.
STRUCTURE

DESIGN FLOOD BASE FLOOD


STATION

OVERTOPPING GREATEST
FLOOD FLOOD
NO.

4% PROB. 1% PROB.
25-YR FREQ 100-YR FREQ
PROB FREQ. PROB FREQ
DISCHARGE STAGE DISCHARGE STAGE DISCHARGE STAGE DISCHARGE STAGE
% YR % YR

S-1 30+50 31 134.3 44 136.4 64 140.9 0.65 154

Note: The hydraulic data are shown for informational purposes only, to indicate the flood discharges and water
surface elevations that may be anticipated in any given year. These data were generated using highly variable
factors determined by a study of the watershed. Many judgments and assumptions are required to establish
these factors. The resultant hydraulic data are sensitive to changes, particularly of antecedent conditions,
urbanization, channelization, and land use. Users of these data are cautioned against the assumption of
precision, which cannot be attained. Discharges are in cubic feet per second and stages are in feet.

Definitions:
Design Flood: The flood selected by FDOT to be utilized to assure a standard level of hydraulic
performance
Base Flood: The flood having a 1-percent chance of being exceeded in any given year (100-year
frequency)
Overtopping Flood: The flood that causes water to flow over the highway, over a watershed divide, or through
emergency relief structures
Greatest Flood: The most severe flood that can be predicted, where overtopping is not practicable;
normally, one with a 0.2-percent chance of being exceeded in any given year (500-year
frequency)

Figure 4.1-4: Flood Data Summary Box

Chapter 4: Culvert 4-8


January 2019

Drainage Design Guide


Chapter 4: Culvert

Example 4.1-3—Completing the Hydraulic Flood Data Sheet

Given the information below, fill out the Hydraulic Flood Data Sheet.

The standard frequency for Structure 1 is 50 years, based on the criteria from
Section 4.3 of the Drainage Manual.

The structure overtops during a 10-year frequency flood.

Perform a risk assessment to define the design flood as the overtopping flood.

Q (Overtopping) = 20 ft3/sec
Stage (Overtopping) = 45 ft.

Q (100) = 37 ft3/sec
Stage (100) = 50.5 ft.

Solution
Since the overtopping flood is less than the standard design frequency and you performed
a risk assessment to define the design flood as the overtopping flood, fill out the
information for the design (overtopping) flood, base flood, and overtopping flood. Put
these values in the corresponding columns, as shown in Figure 4.1-5.

Q (Overtopping) = 20 ft3/sec
Stage (Overtopping) = 45 ft.

Q (100) = 37 ft3/sec
Stage (100) = 50.5 ft.

Chapter 4: Culvert 4-9


January 2019

Drainage Design Guide


Chapter 4: Culvert

Example 4.1-3—Completing the Flood Data Summary Box


STRUCTURE

STATION DESIGN FLOOD BASE FLOOD


OVERTOPPING GREATEST
FLOOD FLOOD
NO.

10% PROB. 1% PROB.


10-YR FREQ 100-YR FREQ
PROB FREQ. PROB FREQ
DISCHARGE STAGE DISCHARGE STAGE DISCHARGE STAGE DISCHARGE STAGE
% YR % YR

S-1 30+50 20 45 37 50.5 20 45 10 10

Note: The hydraulic data are shown for informational purposes only, to indicate the flood discharges and water
surface elevations that may be anticipated in any given year. These data were generated using highly variable
factors determined by a study of the watershed. Many judgments and assumptions are required to establish
these factors. The resultant hydraulic data are sensitive to changes, particularly of antecedent conditions,
urbanization, channelization, and land use. Users of these data are cautioned against the assumption of
precision, which cannot be attained. Discharges are in cubic feet per second and stages are in feet.

Definitions:
Design Flood: The flood selected by FDOT to be utilized to assure a standard level of hydraulic
performance
Base Flood: The flood having a 1-percent chance of being exceeded in any given year (100-year
frequency)
Overtopping Flood: The flood that causes water to flow over the highway, over a watershed divide, or through
emergency relief structures
Greatest Flood: The most severe flood that can be predicted, where overtopping is not practicable;
normally, one with a 0.2-percent chance of being exceeded in any given year (500-year
frequency)

Figure 4.1-5: Flood Data Summary Box

Chapter 4: Culvert 4-10


January 2019

Drainage Design Guide


Chapter 4: Culvert

4.2 DESIGN FREQUENCY


“Design frequency” means a frequency that accommodates an adopted design criterion.
After you determine the design frequency, you can then determine a discharge for the
selected frequency. This discharge is known as the "design discharge." By definition, the
design discharge does not overtop the road. After you determine the design discharge,
you can determine a headwater. This headwater also is known as the "design discharge
headwater.” The design discharge headwater may be at an elevation lower than the
road's profile grade to meet other design criteria, such as protection of property,
accommodating land use needs, lowering velocities, reducing scour, or complying with
regulatory mandates.

To provide an acceptable standard level of service against flooding, the Department


typically employs widely used pre-established design frequencies, which are based on
the importance of the transportation facility to the system and allowable risk for that
facility. Selecting the appropriate design storm from these standards is a matter of
professional judgment since it is rarely either possible or practical to provide for the
greatest possible flood. The design flood frequency standards for cross drains listed in
Section 4.3 of the Drainage Manual provide an engineering consensus on reasonable
values. The actual design must consider the consequences of greater events, such as
the 100-year flood for culverts and bridges and even the 500-year flood for bridges.

Under certain conditions, it may be appropriate to establish a level of risk allowable for a
site and to design to that level. When the risks associated with a particular project are
significant for floods of greater magnitude than the standard design flood, evaluate a
greater return interval design flood by using a risk analysis. Risk analysis procedures are
provided in FHWA’s HEC 17 and discussed briefly in Appendix G, Risk Evaluations. In
addition, consider incorporating or addressing design standards of other agencies that
have control or jurisdiction over the waterway or facility of concern in the design.

Chapter 4: Culvert 4-11


January 2019

Drainage Design Guide


Chapter 4: Culvert

4.3 BACKWATER
Backwater is defined as the increase of water surface elevation induced upstream from
a bridge, culvert, dike, dam, another stream at a higher stage, or other similar structures;
or conditions that obstruct or constrict a channel relative to the elevation occurring under
natural channel and floodplain conditions.

4.3.1 Backwater Consistent with the Flood Insurance Study


Requirements
Backwater Effects on Land Use

Backwater effects are important to consider in the design/analysis of cross drains in rural
and urban areas.

In rural areas, the concern centers on increased flood stages. The degree and duration
of an increased flood stage could affect present and future land uses. You certainly must
evaluate agricultural land use for increased risks due to flooding. As an example,
inundation may impact crops or livestock.

In urban areas, the effects of increased flood stages or increased velocities become an
important consideration. In addition to the impact on future land use, the existing property
may suffer extensive physical damage. Many urban areas have stream or watershed
management regulations or are part of the National Flood Insurance Program (NFIP).
These regulations may dictate limits on changes that can be made to flow characteristics
of a watershed.

You may need to perform a risk evaluation to determine damage to surrounding property.
Refer to Appendix G, Risk Evaluations.

Obtaining Flood Rights

The Department does not encourage obtaining flood rights; however, it is recognized that,
in some instances, it may be necessary. Evaluate all possible alternatives before
recommending that the Department obtains flood rights.

Alternatives to obtaining flood rights for upstream flooding include:

• Prior approval from the property owner


• Purchase of the property
• Upsizing the structure as long as there is no increased flooding to the
downstream owner

Chapter 4: Culvert 4-12


January 2019

Drainage Design Guide


Chapter 4: Culvert

Consider performing a risk analysis for situations where you are evaluating acquiring flood
rights. Appendix G briefly discusses risk analysis, whereas the topic is extensively
covered in HEC 17 (USDOT, FHWA, 1981).

Further discussion about obtaining flood rights is included in Appendix B of the Drainage
Manual.

4.4 TAILWATER
Section 4.5 of the Drainage Manual states: "For the sizing of cross drains and the
determination of headwater and backwater elevations, use the highest tailwater elevation
that can reasonably be expected to occur coincident with the design storm event."

Additional guidelines for tailwater elevations are provided in Section 4.5. For cross drains
subject to tidal conditions, include in the tailwater determination a sea-level rise analysis,
as described in Section 3.4.1 of the Drainage Manual.

4.5 HYDRAULIC ANALYSIS


During a storm event, a culvert may operate under inlet control, outlet control, or both.
Different variables and equations determine the culvert capacity for each type of control.
For more detailed information on theory, refer to Federal Highway Administration
Hydraulic Design Series No. 5 (HDS-5), Hydraulic Design of Highway Culverts. You can
find the publication on FHWA’s website at:
http://www.fhwa.dot.gov/engineering/hydraulics/library_arc.cfm?pub_number=7&id=13.

Guidelines that pertain to the hydraulic analysis of bridge culverts and other culverts are
presented below.

• Allowable Headwater

You can determine the allowable headwater elevation by evaluating land use
upstream of the culvert and the proposed or existing roadway elevation. The
criteria in Section 4.4 of the Drainage Manual apply, but other factors that may limit
the allowable headwater are:

• Identify non-damaging or permissible upstream flooding elevations (e.g.,


existing buildings or flood insurance regulations). Keep headwater below these
elevations.
• Identify state regulatory constraints (e.g., Water Management District).
• Address other site-specific design considerations, as required.

Chapter 4: Culvert 4-13


January 2019

Drainage Design Guide


Chapter 4: Culvert

In general, the constraint that gives the lowest allowable headwater elevation
should establish the basis for hydraulic calculations.

• Inlet Control

Nomographs

FHWA has developed inlet nomographs, shown in FHWA HDS-5, to provide


graphical solutions to headwater equations for various culvert materials, cross
sections, and inlet combinations. Because of the low velocities in most entrance
pools and the difficulty in determining the velocity head for all flows, ignore the
approach velocity and assume the water surface and energy line at the entrance
are coincident. The headwater depths obtained by using the nomographs can be
higher than will occur in some instances because of this factor.

You can determine the headwater elevation for inlet control by taking the culvert invert
elevation at the entrance and adding the headwater depth.

• Outlet Control
Nomographs

Outlet control nomographs have been developed and are shown in FHWA HDS-5
to provide graphical solutions to the head loss equations for various culvert
materials, cross sections, and inlet combinations.

Culvert Entrance Loss Coefficients

Appendix F, Applications Guide for Pipe End Treatments presents culvert entrance
loss coefficients (ke) for the end treatments. For other types of end treatments,
refer to FHWA HDS-5.

Critical Depth

Use FHWA HDS-5 or other suitable methods to determine the critical depth for
various sizes and types of culverts.

Equivalent Hydraulic Elevation

For culverts flowing partially full, the distance from the invert of the culvert outlet to
the equivalent hydraulic grade line is termed the equivalent hydraulic elevation and
is expressed as:

Chapter 4: Culvert 4-14


January 2019

Drainage Design Guide


Chapter 4: Culvert

D+ dc
ho = (Equation 4.5-1)
2

where:
ho = Equivalent hydraulic elevation, in feet, for an unsubmerged outlet condition
D = Depth of the culvert, in feet
dc = Critical depth at the culvert outlet, in feet

If the value for dc from the figures of FHWA HDS-5 is greater than D, then ho will
equal D.

The equivalent hydraulic elevation is valid as long as the headwater is not less
than 0.75D. For headwaters lower than 0.75D, perform backwater calculations to
obtain headwater elevations.

Tailwater

Tailwater (TW) is the depth of water measured from the invert of the culvert at the
outlet to the water surface elevation due to downstream conditions. Evaluate the
hydraulic conditions downstream of the culvert site to determine a tailwater depth
for the discharge and frequency under consideration. Determine tailwater as
follows:

a. If an upstream culvert outlet is near the inlet of a downstream culvert, the


headwater elevation of the downstream culvert may define the tailwater depth
for the upstream culvert.

b. For culverts that discharge to an open channel, the tailwater may be equal to
the normal depth of flow in that channel. Calculate normal depth using a trial-
and-error solution of the Manning’s equation. The known inputs are channel
roughness, slope, and geometry.

For bridge culverts that discharge to an open channel, you may have to
determine the tailwater by performing a standard backwater calculation.
Consider this analysis if the open channel does not have constant channel
roughness, slope, and geometry or if there is a control structure downstream
that could cause backwater.

c. If the culvert discharges to a lake, pond, or other major water body, the
expected high-water elevation of the particular water body may establish the
culvert tailwater. However, it is probably not appropriate to use a 25-year lake

Chapter 4: Culvert 4-15


January 2019

Drainage Design Guide


Chapter 4: Culvert

stage for a cross drain that uses a 25-year design frequency, due to the
difference in time relationship between occurrences. Usually, the mean annual
stage would be appropriate.

d. If tidal conditions occur at the outlet, the mean high water, as determined by
sources such as the National Oceanic and Atmospheric Administration
(NOAA), usually establishes the initial basis for tailwater conditions. Adjust the
mean high water for sea level rise as described in Section 3.4.1 of the Drainage
Manual.

Design Tailwater

The tailwater condition that prevails during the design event is called the design
tailwater (DTW). The design tailwater may be a function of either downstream or
culvert outlet conditions.

Two tailwater conditions can affect the selection of a design tailwater:

a. For the submerged outlet condition shown in Figure 4.5-1, TW is greater


than ho and, thus, TW becomes DTW.

b. For the unsubmerged outlet shown in Figure 4.5-2, TW is less than ho, so
the ho elevation becomes DTW.

Chapter 4: Culvert 4-16


January 2019

Drainage Design Guide


Chapter 4: Culvert

Figure 4.5-1: Tailwater for Submerged Outlet Conditions

Figure 4.5-2: Tailwater for Unsubmerged Outlet Conditions

Chapter 4: Culvert 4-17


January 2019

Drainage Design Guide


Chapter 4: Culvert

Headwater Depth
Having established the total head loss (H) and the design tailwater depth (DTW),
compute the headwater depth (HW), as follows:

HW = H + DTW - LS o (Equation 4.5-2)

where:
HW = Headwater depth for outlet control, in feet
H = Total head, in feet
DTW = Design tailwater depth, in feet
L = Length of culvert barrel, in feet
So = Barrel slope, in feet/feet

The difference in elevation between the culvert inlet and the culvert outlet is equal
to LSo. You may use it directly in Equation 4.5-2.

Determine the headwater elevation for outlet control by taking the culvert invert
elevation at the entrance and adding the headwater depth.

• Controlling Headwater Depth or Elevation

The controlling headwater depth or elevation is defined as the greatest headwater


depth or elevation between the inlet and outlet control conditions.

• Outlet Velocity
Inlet Control

In inlet control, you may need to make backwater calculations to determine the
outlet velocity. These calculations begin at the culvert entrance and proceed
downstream to the exit. Obtain the flow velocity from the flow and the cross
sectional area at the exit:

Q
V= (Equation 4.5-3)
A
where:
V = Average velocity in the culvert, in feet per second
Q = Flow rate, in cubic feet per second
A = Cross sectional area of the flow, in square feet

Chapter 4: Culvert 4-18


January 2019

Drainage Design Guide


Chapter 4: Culvert

To avoid backwater calculations in determining outlet velocity, you may use an


approximation. Since the water surface profile converges toward normal depth as
calculations proceed downstream, you can assume the normal depth and use it to define
the area of flow at the outlet. Then you can use the normal depth obtained to determine
the outlet velocity (see Figure 4.5-3). The velocity obtained may be higher than the actual
velocity at the outlet.

Calculate normal depth using a trial-and-error solution of the Manning equation. The
known inputs are barrel resistance, slope, and geometry. Then, determine the area of
flow prism based on the culvert barrel geometry and depth equal to normal depth. You
also can determine normal depth and area of flow using the charts for various pipe cross
section shapes in Appendix E.

Figure 4.5-3: Outlet Velocity for Inlet Control

Example 4.5-1 illustrates computing outlet velocity for inlet control.

Chapter 4: Culvert 4-19


January 2019

Drainage Design Guide


Chapter 4: Culvert

Example 4.5-1—Computing Outlet Velocity for Inlet Control

Given the information below, determine the outlet velocity for inlet control.

where:
Qdesign = 18 ft3/sec
Diameter of Pipe (D) = 24 in.
Slope of Pipe (S) = 0.01 ft./ft.
Roughness Coefficient (n) = 0.012

Solution
Step 1: Determine area, wetted perimeter, and hydraulic radius of the pipe flowing
full.

( π D 2 ) π × (24 inches /12)2


Area (A) = = = 3.14 ft 2
4 4

Wetted perimeter (WP) = πD = π * (24 in./12) = 6.28 ft.

Hydraulic radius (R) = A/WP = 3.14 ft2/6.28 ft. = 0.5 ft.

Step 2: Using Manning's Equation, determine the discharge and velocity of the pipe
flowing full.

1.49
Q Full = A R 2/3 S 1/2
n

1.49
Q Full = (3.14 ft 2 ) (0.5 ft. )2/3 (0.01 ft./ft. )1/2 = 24.56 ft 3 /s (say 25 ft 3 /s)
0.012

3 2
V Full = Q Full / AFull = 25 ft /s / 3.14 ft = 7.96 ft/s (say 8.0 ft/s)

Step 3: Using Figure 4.5-4, determine the area of flow for the design discharge
using the following relationship:

Q Design 18 ft 3 /s
= = 0.72 or 72 % of value for section
Q Full 25 ft 3 /s

Chapter 4: Culvert 4-20


January 2019

Drainage Design Guide


Chapter 4: Culvert

1. Enter on Figure 4.5-4 the value of 0.72 on the horizontal axis.


2. Project vertically up until the flow curve is met.
3. Project horizontally from the flow curve to the area of the flow curve.
4. Project vertically down from the area of the flow curve and read from
the horizontal axis a value of 0.66 or 66 percent of value for full
section.
5. You can make a relationship between the full flow area and the
normal depth area (A Design):

ADesign
= 0.66 ; ADesign = 0.66 × 3.14 ft 2 = 2.07 ft 2
AFull

Step 4: Determine the outlet velocity using QDesign and ADesign:

Q Design 18 ft 3 /s
V Design = = = 8.70 ft/s
ADesign 2.07 ft 2

Chapter 4: Culvert 4-21


January 2019

Drainage Design Guide


Chapter 4: Culvert

Figure 4.5-4: Example 4.5-1

Chapter 4: Culvert 4-22


January 2019

Drainage Design Guide


Chapter 4: Culvert

Outlet Control:

In outlet control, the cross sectional area of the flow (Ap) is defined by the geometry of
the outlet and either critical depth, tailwater depth, or the height of the culvert (see Figure
4.5-5).

Use critical depth when the tailwater is less than critical depth; use the tailwater depth
when tailwater is greater than critical depth but below the top of the barrel. The total barrel
area is used when the tailwater exceeds the top of the barrel.

Q
V= (Equation 4.5-4)
Ap

where:
V = Average velocity in the culvert, in feet per second
Q = Flow rate, in cubic feet per second
Ap = Cross sectional area of the flow defined by the geometry of the outlet and either
critical depth, tailwater depth, or the height of the culvert, in square feet

You can determine the area of flow prism based on barrel geometry and depth of flow (d)
using the charts for various pipe cross section shapes in Appendix E.

Example 4.5-2 illustrates computing outlet velocity for outlet control.

Chapter 4: Culvert 4-23


January 2019

Drainage Design Guide


Chapter 4: Culvert

Figure 4.5-5: Outlet Velocity for Outlet Control

Chapter 4: Culvert 4-24


January 2019

Drainage Design Guide


Chapter 4: Culvert

Example 4.5-2—Computing Outlet Velocity

Given the information below, determine the outlet velocity for outlet control.

where:
QDesign = 18 ft3/s
Diameter of Pipe = 36 in.
Slope of Pipe = 0.01 ft./ft.
Roughness Coefficient (n) = 0.012
Critical Depth (dc) = 1.4 ft. (determined from FHWA HDS-5, for QDesign = 18 ft3/s)
Tailwater (TW) = 2.0 ft.

Solution
Step 1: Determine the area of the pipe flowing full:

π D 2 π × (36 inches /12)2


Area (A) = = = 7.07 ft 2
4 4

Step 2: Since D > TW > dc, then d = TW Depth or d = 2.0 ft

Step 3: Using Figure 4.5-6 determine the depth of flow to full depth flow (TW/D) or
2 ft./3 ft. = 0.67, or 67 percent of the full depth.

1. Enter on Figure 4.5-6 this value of 0.67 on the horizontal axis.

2. Project horizontally to the area of flow curve.

3. Project vertically down from the area of flow curve and read from the
horizontal axis a value of 0.73, or 73 percent for full section.

4. Use this relationship to determine the normal depth area (A Design):

ADesign
= 0.73 ; ADesign = 0.73 × 7.07 ft 2 = 5.61 ft 2
AFull

Step 4: Determine the outlet velocity using QDesign and ADesign:

18 ft 3 /s
V Design = Q Design / ADesign = = 3.2 ft/s
5.61 ft 2

Chapter 4: Culvert 4-25


January 2019

Drainage Design Guide


Chapter 4: Culvert

Figure 4.5-6: Example 4.5-2

Chapter 4: Culvert 4-26


January 2019

Drainage Design Guide


Chapter 4: Culvert

Culvert Capacity Calculations

a. Worksheet for manual calculations

FHWA’s HDS-5 presents a worksheet for doing culvert capacity


calculations.

b. Computer programs

FHWA’s HY-8 computer program is only one of several programs that are capable of
culvert capacity calculations. The Department has accepted the computer program for
use and it is available through FHWA’s website
(http://www.fhwa.dot.gov/engineering/hydraulics/software/hy8/). Before you use other
computer programs in the design of a Department project, the District Drainage Engineer
should approve them.

4.6 SPECIFIC STANDARDS RELATING TO ALL CROSS DRAINS


EXCEPT BRIDGES
4.6.1 Culvert Materials
Chapter 6 of the Drainage Manual provides standards for suitable optional culvert
materials.

When the vertical distance from invert to roadway is limited, arch culverts may be
appropriate. When the rise of a culvert exceeds four feet, consider the use of box culverts
since they may offer cost advantages.

4.6.2 Scour Estimates


Scour prediction at culvert outlets depends on the following characteristics:
• Channel bed and bank material
• Velocity and depth of flow in the channel and at the culvert outlet
• Velocity distribution
• Amount of sediment and other debris in the flow
• Culvert end section and treatment

A method for estimating the dimensions of a scour hole at a culvert outlet is available in
HEC 14, Chapter 5, linked below:
http://www.fhwa.dot.gov/engineering/hydraulics/pubs/06086/hec14ch05.cfm.

Chapter 4: Culvert 4-27


January 2019

Drainage Design Guide


Chapter 4: Culvert

A good guide in estimating potential scour at the outlet of proposed culverts is to look for
scour developed at the outlet of similar existing culverts. Scour does not develop at all
suspected locations because the susceptibility of the stream to scour is difficult to assess
and the flow conditions that will cause scour do not occur at all flow rates. At locations
where you expect scour to develop only during relatively rare flood events, the most
economical solution may be to repair or retrofit the damage after it occurs.

At many locations, using simple outlet treatments—such as aprons of concrete or riprap—


will provide adequate protection against scour. At other locations, using a rougher culvert
material may be sufficient to prevent damage from scour.

When the outlet velocity is greater than or equal to 12 ft/sec, consider energy dissipation
devices, such as those shown in the Standard Plans.

4.7 RECOMMENDED DESIGN PROCEDURE


The following procedures normally will result in acceptable, cost-effective designs.
However, you are not exempt from developing an appropriate design. You are
responsible for identifying which standards are not applicable to a particular design and
for obtaining variances as necessary to achieve proper design.

The design procedures below do not account for structures within regulatory floodways;
therefore, you may need to deviate from these procedures to satisfy regulatory agencies.
Evaluate and determine the level of effort needed to produce an acceptable design.

Design procedures for three categories of cross drains are provided, including:

1. Culvert extensions (including side drain pipes)


2. Small cross drains (up to 48 inches round or equivalent other shape)
3. Large cross drains (more than 48 inches, but less than a 20-foot bridge)

4.7.1 Culvert Extensions


• Contact the appropriate FDOT Maintenance Office to determine if there is any
history of problems associated with the existing culvert (e.g., flooding, scour, etc.).

• Conduct a field review to evaluate the condition/adequacy of the existing culvert.


Review for condition, signs of scour, and sedimentation. Check the available right
of way to see if there is room to transition ditches to meet the culvert extension.
You can use a review checklist (see the following suggested format) to document
the field review.

Chapter 4: Culvert 4-28


January 2019

Drainage Design Guide


Chapter 4: Culvert

Review Checklist

Date: _________________________________

Project: _________________________________

Location: ________________________ Size/Type___________________________

Road surface/Leaking joints?____________________________________________

Recent development in basin?___________________________________________

Overtopping? Roadway Basin divide In roadway ditch

___________________________________________________________________

Concerns with culvert extension? Limited right of way Wetlands

Normal high water marks:_______________________________________________

Tailwater: Ditch Piped outfall Overland flow Swamp

___________________________________________________________________

Erosion/Sedimentation: ________________________________________________

Misc. Comments: _____________________________________________________

___________________________________________________________

___________________________________________________________

___________________________________________________________

___________________________________________________________

Chapter 4: Culvert 4-29


January 2019

Drainage Design Guide


Chapter 4: Culvert

• Method 1: No Known Historical Problems

If there are no signs of undesirable scour at inlet and outlet ends, no excessive
sedimentation, and no history of problems, you may extend the existing culvert.
The hydrologic and hydraulic analysis would follow the procedure shown below:

a. Estimate discharges as follows:

i. 25 yr. Q = AV where A = Existing Culvert Area


V = 6 feet per second (Confirm this
value with the District Drainage
Engineer; some districts use a
lower velocity)

ii. 100 yr. Q = 1.4 x (25 yr Q)

iii. 500 yr. Q = 1.7 x (100 yr Q)

b. Estimate tailwater. If the outlet is in a free-flowing condition, assume the crown


of the pipe at the outlet is the tailwater.

c. Conduct hydraulic analysis to compute stages using FHWA HDS 5 techniques.

d. Document as required in the Drainage Manual.

General Concerns
Make sure enough right of way exists beyond the ends of the extended culvert to
tie in the roadside ditches and provide for outlet treatment if necessary. There is a
detail in the Standard Plans for ditch transitions at culvert locations. If right of way
is inadequate, consider adjusting the ditch cross-section. If there is not enough
room for the transition shown in the Standard Plans, you may design a sharper
transition, but evaluate the need for channel lining to prevent erosion of the ditch
side slopes. Example 4.7-1 illustrates this method.

Chapter 4: Culvert 4-30


January 2019

Drainage Design Guide


Chapter 4: Culvert

Example 4.7-1—Culvert Extension

Existing: Two-lane rural road


ADT = 2,000 vehicles
36-inch diameter round concrete pipe (RCP)
Length of pipe is 59 feet
Straight end walls

Elevations are as follows:


Allowable headwater (Edge of travel lane) = 105.0 feet
Flow line (upstream) = 100.0 feet
Flow line (downstream) = 99.8 feet

• Contact the appropriate FDOT Maintenance Office to determine if there is any


history of problems associated with the existing culvert (e.g., flooding, scour, etc.).

Spoke with Mr. Steve Smith from the FDOT Maintenance Office on November 18,
1993. From our discussion, we found that there has been no history of problems
in overtopping of the roadway and no complaints of flooding from upstream
property owners have been found.

• Conduct a field review to evaluate condition/adequacy of existing culvert. Review


for condition, signs of scour, and sedimentation.

We performed a field review with Mr. Smith on November 21, 1993. From our
review, we determined the culvert was in good condition, with no signs of
sedimentation or scour.

• No Known Historical Problems


Since there were no known historical problems, use Method 1. Recommend that
the existing 36-inch RCP be extended four feet in both directions with 36-inch RCP
and straight endwalls (Standard Plans, Index 430-030). The proposed flow line
elevations are as follows:

Flow line (upstream) = 100.1 feet


Flow line (downstream) = 99.7 feet

a. Estimate discharges as follows:

Area of 36-inch RCP = (πD2)/4 = (π(36 inch/12)2)/4 = 7.07 ft2

Q(25) = AV = 7.07 ft2 x 6 ft/sec = 42 ft3/sec


Q(100) = 1.4 x Q(25) = 59 ft3/sec
Q(500) = 1.7 x Q(100) = 100 ft3/sec

Chapter 4: Culvert 4-31


January 2019

Drainage Design Guide


Chapter 4: Culvert

Since this roadway has an ADT > 1,500, the design frequency is 50 years
(determined from the Drainage Manual). To determine the 50-year
discharge, a procedure similar to that used in Example 4.1-1 is appropriate.
For this example, the Q(50) is 50 ft3/sec.

b. Estimate tailwater as discussed in Chapter 3 (Open Channel) or if outlet is in a


free-flowing condition, assume the crown of the pipe at the outlet is the
tailwater.

For this example, the 50-year tailwater elevation to be used will be:

TW (50 year) = 2.7 ft.

c. Conduct hydraulic analysis using the procedures in FHWA HDS 5.

For this example, only the hydraulic analysis for the 50-year frequency will be
computed. However, you also would need to compute an analysis for the other
frequencies. The analysis is for the proposed conditions. Figure 4.7-1
summarizes the following calculations.

Chapter 4: Culvert 4-32


January 2019

Drainage Design Guide


Chapter 4: Culvert

Figure 4.7-1: Culvert Capacity Worksheet for Example 4.7-1

Chapter 4: Culvert 4-33


January 2019

Drainage Design Guide


Chapter 4: Culvert

• Inlet Control

Nomographs:

Using Chart 1 in FHWA HDS 5, HW/D = 1.27


Therefore,
HW = 1.27 x D = 1.27 x 3 ft. = 3.81 ft., say 3.8 ft.

Determine the headwater elevation by taking the culvert invert at the entrance
and adding the headwater depth:

HW Elevation = 100.1 ft. + 3.8 ft. = 103.9 ft.

• Outlet Control

Nomographs:

Using Chart 5 in FHWA HDS 5 with a pipe length of 67 feet (existing 59 ft. + 8
ft. of extension) and an entrance loss coefficient of 0.2 feet (as determined
below), the headwater (H) for the 50-year discharge is 1.55.

Culvert Entrance Loss Coefficients (Ke):

Culvert entrance loss is 0.2 as determined from the Application Guidelines for
Pipe End Treatment, Appendix F, based on the structure having a standard
end wall treatment.

Critical Depth (dc):

Using Chart 4 in FHWA HDS 5, the critical depth was found to be 2.3 feet.

Equivalent Hydraulic Elevation (ho):

D + d c 3 ft + 2.3 ft
ho = = = 2.65 ft.
2 2

Chapter 4: Culvert 4-34


January 2019

Drainage Design Guide


Chapter 4: Culvert

Design Tailwater (DTW):

Since the TW > ho, then the DTW = TW = 2.7 ft.

Headwater Depth (HW):

Having established the total head loss (H) and the design tailwater depth
(DTW) as described above, compute the headwater depth (HW), as follows:

HW = H + DTW - LSo
HW = 1.55 ft. + 2.70 ft. - (0.4 ft.)
HW = 3.85 ft., say 3.9 ft.

Determine the headwater elevation by taking the culvert invert at the entrance
and adding the headwater depth:

HW Elevation = 100.1 ft. + 3.9 ft. = 104.0 ft.

• Controlling Headwater (HW) Depth or Elevation

Since the HW depth or elevation for outlet control (HW Elevation = 104.0 feet)
is greater than that of inlet control (HW Elevation = 103.9 feet), then the
controlling HW Elevation is 104.0 feet.

• Outlet Velocity

Outlet velocity for a culvert for this type of problem does not need to be
computed since the discharges were estimated using a 25-year velocity of 6
ft/sec.

d. Document as required in the Drainage Manual.

End of Example 4.7-1

Chapter 4: Culvert 4-35


January 2019

Drainage Design Guide


Chapter 4: Culvert

• Method 2: Known Historical Problems or If the Analysis Yields Unrealistic


Results
If scour, sedimentation, or other known historical problems exist, or if Method 1
yields unrealistic results, conduct complete hydrologic and hydraulic analysis and
evaluate alternatives.

a. Conduct a complete hydrologic analysis using one of the following methods, as


appropriate (see Section 4.7 of the Drainage Manual):

- Frequency analysis of observed data


- Regional or local regression equation
- Rational Equation (up to 600 acres)

b. Determine tailwater conditions.

c. Conduct hydraulic analysis using procedures in FHWA HDS 5.

d. Assess cause of problem and investigate/evaluate alternative solutions. Final


recommended design should address the problem with consideration to design
standards.

e. Document as required in the Drainage Manual.

General Considerations
The ditch transition concerns in the previous section also apply here. In addition,
any problems such as scour, sedimentation, etc., should be limited to within the
right of way or not extend any further outside the right of way than they currently
extend. Example 4.7-2 illustrates this procedure.

Chapter 4: Culvert 4-36


January 2019

Drainage Design Guide


Chapter 4: Culvert

Example 4.7-2—Culvert Extension

Existing: Two-lane rural road


ADT = 2,000
2 foot x 2 foot concrete box culvert cross drain
Length of pipe is 50 feet
Straight end walls

Elevations are as follows:


Allowable headwater (edge of travel lane) = 104.6 feet
Flow line (upstream) = 100.0 feet
Flow line (downstream) = 99.8 feet

• Contact appropriate FDOT Maintenance Office to determine if there is any


history of problems associated with the existing culvert (e.g., flooding, scour, etc.).

Spoke with Mr. Steve Smith from the FDOT Maintenance Office on November 18,
1993. We found that there has been history of overtopping of the roadway.

• Conduct a field review to evaluate condition/adequacy of existing culvert. Review


for condition, signs of scour, and sedimentation.

We performed a field review with Mr. Smith on November 21, 1993. From our
review, we discovered that the area around the outlet end of the culvert showed
signs of scouring.

• Known Historical Problem


Since the area around the outlet end of the culvert showed signs of scouring,
analyze the structure using Method 2.

a. Conduct a complete hydrologic analysis using one of the following methods, as


appropriate (see Section 4.7 of the Drainage Manual):

- Frequency analysis of observed data


- Regional or local regression equation
- Rational Equation (up to 600 acres)

Chapter 4: Culvert 4-37


January 2019

Drainage Design Guide


Chapter 4: Culvert

From the field review and hydrologic calculations, the following design
information is known:

Q(50) = 35 ft3/sec
Q(100) = 52 ft3/sec
Q(500) = 88 ft3/sec

b. Determine tailwater as discussed in Chapter 3 (Open Channel) or if outlet is in


a free-flowing condition, the crown of the pipe at the outlet may be assumed.

For this example, the 50-year tailwater elevation to be used will be TW (50
year) = 2.5 ft.

c. Conduct hydraulic analysis using the procedures in FHWA HDS 5.

For this example, only a hydraulic analysis for the 50-year frequency will be
computed. The other frequencies also would need to be analyzed for an actual
project. The analysis is for the existing conditions. Figure 4.7-2 summarizes the
following calculations.

• Inlet Control

Nomographs

Using Chart 8 in FHWA HDS 5, Q/B = (35 ft3/sec)/2 ft. = 17.5 ft3/sec.
Therefore, HW/D = 2.4 and HW = 2.4 ft. x D = 2.4 ft. x 2 ft = 4.8 ft.

Determine the headwater elevation by taking the culvert invert at the


entrance and adding the headwater depth:

HW Elevation = 100.0 ft + 4.8 ft = 104.8 ft

Chapter 4: Culvert 4-38


January 2019

Drainage Design Guide


Chapter 4: Culvert

• Outlet Control

Nomographs

Using Chart 15 in FHWA HDS 5, the headwater (H) for the 50-year
discharge is 2.2 feet based on the pipe length of 50 feet and entrance loss
coefficient of 0.2, as determined below.

Culvert Entrance Loss Coefficients (Ke)

Culvert entrance loss is 0.2 as determined from the Application Guidelines


for Pipe End Treatments, Appendix F, based on the structure having a
straight end wall treatment.

Critical Depth (dc)

Using Chart 14 in FHWA HDS 5, the critical depth was found to be 2 feet.

Equivalent Hydraulic Elevation (ho)

D + d c 2 ft. + 2 ft.
ho = = = 2 ft.
2 2

Design Tailwater (DTW)

Since the TW > ho, then the DTW = TW = 2.5 ft.

Chapter 4: Culvert 4-39


January 2019

Drainage Design Guide


Chapter 4: Culvert

Headwater Depth (HW):

Having established the total head loss (H) and the design tailwater depth
(DTW) as described above, compute the headwater depth (HW), as follows:

HW = H + DTW - LSo
HW = 2.2 ft. + 2.5 ft. - (0.2 ft.)
HW = 4.5 ft.

Determine the headwater elevation by taking the culvert invert at the


entrance and adding the headwater depth:

HW Elevation = 100.0 ft. + 4.5 ft. = 104.5 ft.

• Controlling Headwater (HW) Depth or Elevation

Since the HW depth or elevation for inlet control (HW elevation = 104.8 feet)
is greater than that of outlet control (HW elevation = 104.5 feet), then the
controlling HW elevation is 104.8 feet.

• Outlet Velocity

Since the existing structure was found to be inlet control, the outlet velocity
was determined as discussed earlier in this section.

Chapter 4: Culvert 4-40


January 2019

Drainage Design Guide


Chapter 4: Culvert

Figure 4.7-2: Worksheet for Example 4.7-2

Chapter 4: Culvert 4-41


January 2019

Drainage Design Guide


Chapter 4: Culvert

d. Assess the cause of the problem and investigate/evaluate alternative solutions.


Final recommended design should address the problem with consideration to
design standards.

Review of Figure 4.7-2 indicates that the roadway is overtopped for a 50-year
design frequency. Therefore, recommend replacing the structure. It is
anticipated that a cross drain no larger than a 48-inch diameter would be
appropriate for this location. The procedure in Section 4.7.2, following, could
be used. Example 4.7-3 illustrates this using the information from this example.

e. Document as required in the Drainage Manual.

End of Example 4.7-2

Chapter 4: Culvert 4-42


January 2019

Drainage Design Guide


Chapter 4: Culvert

4.7.2 Small Cross Drains


This information applies to cross drains having an area of opening up through a 48-inch-
diameter round culvert or the equivalent.

• Conduct hydrologic analysis


Estimate discharges for design year frequency, base flood, and greatest flood. Use
one of the following procedures as appropriate (see Section 4.7 of the Drainage
Manual):

- Rational Equation (up to 600 acres)


- Regional or Local Regression Equation

• Select trial culvert size based on the following:

A = Q/V

Where:
A = Culvert area (square feet)
Q = Design discharge (e.g., 50 year)
V = Average velocity (feet per second); use an average velocity of four feet per second

• Estimate tailwater. If the outlet is in a free-flowing condition, the crown of the pipe
at the outlet may be assumed.

• Conduct hydraulic analysis using techniques provided in FHWA HDS 5.


Compute headwater conditions for the selected size for the design flood, base
flood, and greatest flood or overtopping flood as appropriate.

• Check hydraulic results against design standards for backwater, minimum size,
and scour. If these standards are satisfied, the trial culvert size is acceptable.

• Determine the most economical culvert size that satisfies all standards. If the
trial selected size does not satisfy all design standards, obtain a variance.

• Document as required in the Drainage Manual.

Example 4.7-3 illustrates this procedure.

Chapter 4: Culvert 4-43


January 2019

Drainage Design Guide


Chapter 4: Culvert

Example 4.7-3—Design of Small Cross Drain

Referring back to Example 4.7-2, you determined that the two-foot x two-foot concrete
box culvert should be replaced. A design frequency of 50 years was determined as the
minimum for this roadway. The existing length of the two-foot x two-foot concrete box
culvert was 50 feet. However, since the structure will have to be extended four feet on
each side, the design length of the proposed structure will be 58 feet.

Proposed Elevations are as follows:


Allowable headwater (edge of travel lane) = 104.6 ft
Flow line (upstream) = 100.1 ft
Flow line (downstream) = 99.7 ft

• Conduct hydrologic analysis


Estimate discharges for design-year frequency, base flood, and greatest flood. Use
one of following procedures as appropriate (see Section 4.7 of the Drainage
Manual):

- Rational Equation (up to 600 acres)


- Regional or Local Regression Equation

Use the same discharges from Example 4.7-2:

Q(50) = 35 ft3/sec
Q(100) = 52 ft3/sec
Q(500) = 88 ft3/sec

• Select trial culvert size

Q 35 ft 3 /s
A= = = 8.8 ft 2
V 4 ft/s

D = 3.3 ft., so try D = 36-inch pipe and 42-inch pipe

Chapter 4: Culvert 4-44


January 2019

Drainage Design Guide


Chapter 4: Culvert

• Conduct hydraulic analysis using FHWA HDS 5 procedures.

The hydraulic analysis would be similar to what was done in Example 4.7-1 and
Example 4.7-2. A worksheet of the calculations for the 50-year frequency is shown
in Figure 4.7-3. The other frequencies also would need to be analyzed for an actual
project. The analysis shown in Figure 4.7-3 is for the proposed conditions.

• Check hydraulic results against design standards.

Review of the worksheet in Figure 4.7-3 indicates that the roadway will not overtop
for the 50-year frequency for either culvert size. There is very little difference
between the 36-inch and 42 inch pipe as far as controlling headwater. Therefore,
either pipe size would be adequate. However, it is recommended that the 36-inch
pipe be installed since it would be slightly less in cost than the 42-inch pipe. In
addition, it would be recommended that a rubble ditch lining design be installed at
the outlet end due to velocities exceeding six feet per second.

• If design does not meet standards or if you can use more economical culvert size
that satisfies the standards, then perform new computations for that design.

Document as required in the Drainage Manual.

Chapter 4: Culvert 4-45


January 2019

Drainage Design Guide


Chapter 4: Culvert

Figure 4.7-3: Worksheet for Example 4.7-3

Chapter 4: Culvert 4-46


January 2019

Drainage Design Guide


Chapter 4: Culvert

4.7.3 Large Cross Drains


This information applies to cross drains having an area of opening greater than a 48-inch
diameter pipe and less than a 20-foot bridge. The procedure for large cross drains is
similar to that for small cross drains except that a greater level of effort and detail is
expected in developing the hydrologic estimates and the determination of tailwater
conditions.

• Conduct hydrologic analysis


Estimate discharges for design-year frequency, base flood, and greatest flood. Use
one of following procedures as appropriate (see Section 4.7 of the Drainage
Manual):

- Frequency analysis of observed conditions


- Regional or Local Regression Equation
- Rational Equation (up to 600 acres)

The remaining steps are the same as those identified in Section 4.7.2 for small cross
drains.

Chapter 4: Culvert 4-47


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

CHAPTER 5: BRIDGE HYDRAULICS

5. BRIDGE HYDRAULICS.........................................................................................5-1

5.1 PROJECT APPROACH and MISCELLANEOUS CONSIDERATIONS................5-1


5.1.1 Identify Hydraulic Conditions...........................................................................5-1
5.1.2 Floodplain Requirements ................................................................................5-5
5.1.2.1 FEMA Requirements................................................................................5-5
5.1.2.2 Other Government Agency Requirements...............................................5-9
5.1.3 Design Frequencies ........................................................................................5-9
5.1.4 Clearances ....................................................................................................5-11
5.1.4.1 Debris.....................................................................................................5-11
5.1.4.2 Navigation ..............................................................................................5-11
5.1.4.3 Waves ....................................................................................................5-12
5.1.5 Bridge Length Justification ............................................................................5-12
5.1.6 Berms and Spill-Through Abutment Bridges.................................................5-13
5.1.7 Design Considerations for Dual Bridges .......................................................5-16
5.1.8 Design Considerations for Bridge Widenings ...............................................5-18

5.2 RIVERINE ANALYSIS .........................................................................................5-21


5.2.1 Data Requirements .......................................................................................5-21
5.2.1.1 Geometric Data......................................................................................5-21
5.2.1.2 Geotechnical Data .................................................................................5-23
5.2.1.3 Historical Data........................................................................................5-24
5.2.1.4 Drainage Basin Information ...................................................................5-32
5.2.1.5 FEMA Maps ...........................................................................................5-32
5.2.1.6 Upstream Controls .................................................................................5-33
5.2.1.7 Site Investigation....................................................................................5-33
5.2.2 Hydrology ......................................................................................................5-34
5.2.3 Model Selection.............................................................................................5-34
5.2.3.1 One-Dimensional versus Two-Dimensional...........................................5-35
5.2.3.2 Steady versus Unsteady Flow ...............................................................5-35
5.2.3.3 Commonly Used Programs....................................................................5-35
5.2.4 Model Setup ..................................................................................................5-36
5.2.4.1 Defining the Model Domain....................................................................5-37
5.2.4.2 Roughness Coefficient Selection ...........................................................5-40
5.2.4.3 Model Geometry ....................................................................................5-41
5.2.4.4 Boundary Conditions..............................................................................5-63
5.2.4.5 Bridge Model..........................................................................................5-64
5.2.5 Simulations....................................................................................................5-70
5.2.5.1 Calibration..............................................................................................5-70

Chapter 5: Bridge Hydraulics 5-i


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.2.5.2 Existing Conditions ................................................................................5-76


5.2.5.3 Design Considerations...........................................................................5-76

5.3 TIDAL ANALYSIS................................................................................................5-77


5.3.1 Data Requirements .......................................................................................5-77
5.3.1.1 Survey Data ...........................................................................................5-77
5.3.1.2 Geotechnical Data .................................................................................5-79
5.3.1.3 Historical Information .............................................................................5-79
5.3.1.4 FEMA Maps ...........................................................................................5-91
5.3.1.5 Inland Controls.......................................................................................5-91
5.3.1.6 Site Investigation....................................................................................5-91
5.3.2 Hydrology (Hurricane Rainfall) ......................................................................5-92
5.3.3 Model Selection.............................................................................................5-93
5.3.3.1 Storm Surge Model ................................................................................5-95
5.3.3.2 Wave Model ...........................................................................................5-96
5.3.3.3 Model Coupling ......................................................................................5-96
5.3.4 Model Setup ..................................................................................................5-97
5.3.4.1 Defining the Model Domain....................................................................5-97
5.3.4.2 Roughness Selection .............................................................................5-98
5.3.4.3 Model Geometry ....................................................................................5-99
5.3.4.4 Boundary Conditions..............................................................................5-99
5.3.4.5 Bridge...................................................................................................5-106
5.3.5 Simulations..................................................................................................5-107
5.3.5.1 Model Calibration .................................................................................5-107
5.3.5.2 Storm Surge Simulations .....................................................................5-109
5.3.5.3 Design Considerations.........................................................................5-110
5.3.5.4 Wave Simulations ................................................................................5-111
5.3.6 Wave Forces on Bridge Superstructures ....................................................5-112

5.4 MANMADE CONTROLLED CANALS...............................................................5-115


5.4.1 Introduction .................................................................................................5-115
5.4.2 Watershed Description & Flow....................................................................5-115
5.4.3 Channel Excavation, Clearance, and Other Owner Requirements.............5-116
5.4.4 Scour Estimation .........................................................................................5-116
5.4.5 Abutment Protection....................................................................................5-116
5.4.6 Bridge Deck Drainage .................................................................................5-116
5.4.7 Appendix .....................................................................................................5-116

5.5 BRIDGE SCOUR ...............................................................................................5-117


5.5.1 Scour Components .....................................................................................5-117
5.5.1.1 Long-Term Channel Processes ...........................................................5-117
5.5.1.2 Contraction Scour ................................................................................5-120
5.5.1.3 Local Scour (Pier and Abutment).........................................................5-124
5.5.1.4 Scour Considerations for Waves .........................................................5-129
5.5.2 Scour Considerations for Ship Impact.........................................................5-130

Chapter 5: Bridge Hydraulics 5-ii


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.5.3 Florida Rock/Clay Scour Procedure............................................................5-132


5.5.3.1 Pressure Scour ....................................................................................5-134
5.5.3.2 Debris Scour ........................................................................................5-134
5.5.4 Scour Countermeasures .............................................................................5-135
5.5.4.1 Abutment Protection ............................................................................5-135
5.5.4.2 Scour Protection at Existing Piers........................................................5-145

5.6 DECK DRAINAGE .............................................................................................5-146


5.6.1 Bridge End Drainage...................................................................................5-146
5.6.2 No Scuppers or Inlets (Option 1).................................................................5-147
5.6.3 Scuppers (Option 2) ....................................................................................5-151
5.6.4 Closed Collection Systems (Option 3) ........................................................5-167

5.7 BRIDGE HYDRAULICS REPORT FORMAT and DOCUMENTATION ...............5-169


5.7.1 Bridge Hydraulics Report Preparation.........................................................5-170
5.7.1.1 Executive Summary .............................................................................5-170
5.7.1.2 Introduction ..........................................................................................5-171
5.7.1.3 Floodplain Requirements .....................................................................5-172
5.7.1.4 Hydrology.............................................................................................5-172
5.7.1.5 Hydraulics ............................................................................................5-172
5.7.1.6 Scour....................................................................................................5-186
5.7.1.7 Deck Drainage .....................................................................................5-186
5.7.1.8 Appendices ..........................................................................................5-186
5.7.2 Bridge Hydraulics Report Process ..............................................................5-186
5.7.3 Common Review Comments ......................................................................5-191
5.7.4 Bridge Hydraulics Recommendations Sheet (BHRS) .................................5-194

Chapter 5: Bridge Hydraulics 5-iii


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5. BRIDGE HYDRAULICS

5.1 PROJECT APPROACH AND MISCELLANEOUS


CONSIDERATIONS
The material in this section addresses background information and initial decision making
needed in preparation for a bridge hydraulic design. The following sections present more
detailed design guidance.

Most bridge projects in Florida receive funding from FHWA. Even if the project is not
planned to receive federal funding, the funding situation may change before the project
is complete. As a result, much of the hydraulic analyses and documentation required by
the Department’s standards are tailored to satisfy federal regulations and requirements.

FHWA 23 CFR 650A outlines the principal hydraulic analysis and design requirements
that you must satisfy to qualify bridge projects (as well as any other project involving
floodplain encroachments) for Federal Aid. The requirements in 23 CFR 650A are very
comprehensive, so you, as the drainage engineer, should become familiar with them. The
document is available at: (http://www.fhwa.dot.gov/legsregs/directives/fapg/cfr0650a.htm).

5.1.1 Identify Hydraulic Conditions


Before beginning any hydraulic analysis of a bridge, first you must determine the mode
of flow for the waterway. For purposes of bridge hydraulics, the Department separates
the mode of flow into three categories of tidal influence during the bridge design flows:

1. Riverine flow—Crossings with no tidal influence during the design storm, such as:
(a) inland rivers, or (b) controlled canals with a salinity structure oceanward
intercepting the design hurricane surge. Bridges identified as riverine dominated
require only examination of design runoff conditions.

2. Tidally dominated flow—Crossings where the tidal influences are dominated by the
design hurricane surge. Flows in tidal inlets, bays, estuaries, and interconnected
waterways are characterized by tide propagation evidenced by flow reversal
(Zevenbergen et al., 2004). Large bays, ocean inlets, and open sections of the
Intracoastal Waterway typically are tidally dominated, so much so that even
extreme rainfall events have little influence on the design flows in these systems.
Tidally dominated areas with negligible upland influx require only examination of
design storm surge conditions.

3. Tidally influenced flow—Both river flow and tidal fluctuations affect flows in tidally
influenced crossings, such as tidal creeks and rivers opening to tidally dominated
waterways. Tidally affected river crossings do not always experience flow reversal;

Chapter 5: Bridge Hydraulics 5-1


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

however, backwater effects from the downstream tidal fluctuation can induce water
surface elevation fluctuations up through the bridge reach. Tidally influenced
bridges require you to examine both design runoff and surge conditions to
determine which hydraulic (and scour) parameter will dictate design. For example,
a bridge located near the mouth of a river that discharges into a tidal bay (see
Figure 5.1-2) may experience a high stage during a storm surge event. However,
high losses through the bridge and a relatively small storage area upstream may
limit the flow (and velocities) through the bridge. In fact, the design flow parameters
(and thus scour) may occur during the design runoff event while the design stage
(for clearance) and wave climate occurs during the storm surge event. Given that
tidally influenced crossings may require both types of analyses, you should plan to
include a coastal engineer for these bridge projects.

The level of tidal influence is a function of several parameters, including distance from the
open coast, size of the upstream watershed, elevation at the bridge site, conveyance
between the bridge and the open coast, upstream storage, and tidal range.

By far, the best indicator is distance from the coast. Comparisons of gage data or tidal
benchmarks with distance from the coast will illustrate the decrease in tidal influence with
increasing distance (see Figure 5.1-2). The figure shows that with increasing distance,
the tidal range decreases, the flow no longer reverses, and, eventually, the tidal signal
dies out completely. This illustrates the transition from tidally controlled (gage 2323592),
to tidally influenced (gages 2323590, and 2323567, and 2323500), and finally to a riverine
dominant system (gage 2323000).

Chapter 5: Bridge Hydraulics 5-2


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

USGS Gages on Suwanee River


6

5
2323592 (7 miles from coast)
4 2323590 (13 miles from coast)
2323567 (20 miles from coast)
2323500 (28 miles from coast)
Stage in ft-NGVD

3
2323000 (47 miles from coast)

-1

-2
10/31/2010 11/2/2010 11/4/2010 11/6/2010 11/8/2010 11/10/2010
Date

Figure 5.1-1: USGS Gage Data from the Suwannee River with Increasing Distance
from the Coast

Chapter 5: Bridge Hydraulics 5-3


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

US 90 over the
US-90
Escambia River

Escambia Bay

Figure 5.1-2: Example of a Bridge Requiring both Riverine and Tidal Analyses
(US-90 over Escambia Bay)

For the purposes of Department work, a coastal engineer is an engineer who holds a
Master of Science or doctoral degree in coastal engineering or a related engineering field
and/or has extensive experience (as demonstrated by publications in technical journals
with peer review) in coastal hydrodynamics and sediment transport processes.

Chapter 5: Bridge Hydraulics 5-4


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.1.2 Floodplain Requirements


Address potential floodplain impacts during the Project Development and Environment
(PD&E) phase of the project. Usually, you will not prepare a Bridge Hydraulics Report
(BHR) during PD&E studies. However, if you do not prepare a BHR for a bridge, then the
Location Hydraulic Study should address:

 Conceptual bridge length


 Conceptual scour considerations
 Preliminary vertical grade requirements
 The need, if any, for the input of a coastal engineer during final design

Refer to the PD&E or environmental documents and the Location Hydraulic Report for
commitments made during the PD&E phase. Refer to Part 2, Chapter 24 of the FDOT
Project Development and Environment Manual for more information on floodplain
assessment during PD&E.

5.1.2.1 FEMA Requirements


All bridge crossings must be consistent with the National Flood Insurance Program
(NFIP), which will depend on the presence of a floodway and the participation status of
the community. To determine these factors, review:

 Flood maps for the bridge site, if available, to determine if the floodplain has been
established by approximate methods or by a detailed study, and if a floodway has
been established.
 Community Status Book Report to determine the status of the community’s
participation in the NFIP.

Both the flood maps and the Status Book are available at the Federal Emergency
Management Agency (FEMA) website: http://www.fema.gov/.

The Special Flood Hazard Area (SFHA) is the area within the 100-year floodplain (refer
to Figure 5.1-3). If a floodway has been defined, it will include the main channel of the
stream or river, and usually a portion of the floodplain. The remaining floodplain within the
SFHA is called the floodway fringe. The floodway is established by including simulated
encroachments in the floodplain that will cause the 100-year flood elevation to increase
one foot (refer to Figure 5.1-4).

Figure 5.1-5 shows an example of a floodway on the flood map. The floodway, as well as
other map features, may have a different appearance on different community flood maps.
Each map will have a legend for the various features on the map.

Chapter 5: Bridge Hydraulics 5-5


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.1-3: Special Flood Hazard Area

Figure 5.1-4: Floodway Definitions

Chapter 5: Bridge Hydraulics 5-6


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Base Flood Elevation


Elevations for the 100-year flood

Floodway
The cross-hatched area. Includes
the most conveyance and highest
velocities.

Zone AE and Zone A


Zone AE: Subject to flooding by
the 100-year flood as determined
by a detailed study.
Zone A: Flooding area determined
by approximate methods.

Zone X (shaded)
Subject to flooding by the 500-
year flood. Zone B on some maps.

Zone X (unshaded)
Outside 500-year floodplain.

Figure 5.1-5: Example Flood Map

The simplest way to be consistent with the NFIP standards for an established floodway
is to design the bridge and approach roadways so that you exclude their components
from the floodway. If a project element encroaches on the floodway but has a very minor
effect on the floodway water surface elevation (such as piers in the floodway), the project
may be considered consistent with the standards if hydraulic conditions can be improved
so that no water surface elevation increase is reflected in the computer printout for the
new conditions. You will prepare a No-Rise Certification and support it by technical data.
Base the data on the original model used to establish the floodway. The FEMA website
has contact information to obtain the original model.

A Flood Insurance Study (FIS) documents methods and results of the detailed hydraulic
study. The report includes the following information:

 Name of community
 Hydrologic analysis methods
 Hydraulic analysis methods

Chapter 5: Bridge Hydraulics 5-7


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

 Floodway data, including areas, widths, average velocities, base flood elevations,
and regulatory elevations
 Water surface profile plots

The FIS can be obtained from the FEMA website. Note that the report does not include
the original hydraulic model.

For some rivers and streams, a detailed study was performed, but a floodway was not
established (refer to Figure 5.1-6). In this case, the bridge and roadway approaches may
be designed to allow no more than a one-foot increase in the base flood elevation
depending on local regulations and if offsite land use values will not be significantly
impacted (see Section 4.4 of the Drainage Manual). Use information from the FIS and the
original hydraulic model to model the bridge, and submit technical data to the local
community and FEMA.

Zone AE
Subject to flooding by the 100-year flood as
determined by a detailed study.

Base Flood Elevation


Elevations for the 100-year flood

Zone A
Flooding area determined by approximate
methods.

Figure 5.1-6: Example Flood Map

Chapter 5: Bridge Hydraulics 5-8


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

If the encroachment is in an area without a detailed study (Zone A on Figure 5.1-5 and
Figure 5.1-6), then generate technical data for the project. You should give base flood
information to the local community. Pursuant to NFIP regulations in CFR 60.3(c)(10), no
more than a foot of increase in base flood elevation is allowed for cumulative development
within the floodplain.

5.1.2.2 Other Government Agency Requirements


Many government agencies (cities, counties, Water Management Districts, etc.) will have
additional limitations on backwater conditions in floodplains. The agency may designate
the limitations at multiple distances upstream from the bridge. For example, backwater
increase immediately upstream must be limited to one foot, and backwater increase 1,000
feet upstream must be limited to 0.1 foot.

Many of these agencies also have implemented mitigation requirements for fill within the
floodplain, because it reduces the storage capacity in the floodplain and may increase
discharges downstream. Therefore, other agencies may require a compensation area that
creates the amount of storage lost due to the roadway approach fill.

5.1.3 Design Frequencies


Design frequency requirements are given in Section 4.3 of the Drainage Manual. These
design frequencies are based on the importance of the transportation facility to the system
and allowable risk for that facility. They provide an acceptable standard level of service
against flooding.

Criteria that are based on the design frequency include:

 The bridge must convey the design frequency without damage (Section 4.2 of the
Drainage Manual).
 Backwater for the design frequency must be at or below the travel lanes (Section
4.4 of the Drainage Manual).
 The bridge must have adequate debris clearance.

Figure 5.1-7 showsthe relationship between these design frequency criteria and the
geometric design. These criteria tend to create a crest curve on the bridge, with the profile
of the approach roadway lower than the bridge profile. This is a desirable profile because
the roadway will overtop before the bridge is inundated. Losing the roadway is preferable
to losing the bridge.

Backwater criteria also apply for floods other than the design flood:

 Backwater must be consistent with the NFIP.


 Backwater must not change the land use of affected properties without obtaining
flood rights.

Chapter 5: Bridge Hydraulics 5-9


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

When the risks associated with a particular project are significant for floods of greater
magnitude than the standard design flood, a greater return interval design flood should
be evaluated by use of a risk analysis. Risk analysis procedures are provided in FHWA
HEC 17 and discussed briefly in Appendix G, Risk Evaluations. Discuss changing the
design frequency with the District Drainage Engineer before making a final decision. In
addition, incorporate or address in the design hydraulic design frequency standards of
other agencies that have control or jurisdiction over the waterway or facility.

Figure 5.1-7: Bridge and Cross Drain Roadway Grade Controls

Chapter 5: Bridge Hydraulics 5-10


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Scour analysis and design has a separate design frequency, discussed in Section 4.9 of
the Drainage Manual. You will find national standards for scour design in FHWA HEC 18,
Evaluating Scour at Bridges.

The worst-case condition for scour usually will occur at overtopping of the approach
roadway or another basin boundary. Overtopping flow at the bridge often provides flow
relief, and scour conditions will be a maximum at overtopping.

For more guidance on scour computation and design, refer to Section 5.5 of this
document and the FDOT Bridge Scour Manual.

5.1.4 Clearances
The span lengths of a bridge affect the cost of the bridge, with longer spans generally
increasing the cost. Increased height above the ground increases the cost of the
foundations and the earthen fill of the approach roadways. However, minimum vertical
and horizontal clearance requirements must be maintained to ensure the hydraulic
crossing functions in conformance with the design criteria. Minimum clearances are
addressed in the FDM 210.

5.1.4.1 Debris
The two-foot minimum debris drift clearance used by the Department traditionally has
provided an acceptable level of service. Though this clearance usually is adequate for
facilities of all types, review bridge maintenance records for the size and type of debris
that may be expected. For example, if the watershed is a forested area subject to
timbering activities, anticipate sizeable logs and trees among the debris. Meandering
rivers also will tend to fell trees along the banks, carrying them toward downstream bridge
crossings. On the other hand, bridges immediately downstream from a pump station may
have little opportunity to encounter debris. Also, manmade canals tend to be stable
laterally and will fell many fewer trees than sinuous, moving natural rivers. In such low
debris cases, if a reduced vertical clearance is economically ideal, the hydraulic designer
should approach the District Drainage Engineer to reduce the debris drift clearance.

For new bridges, you should advocate for aligning the piers normal to the flow if there is
a possibility of debris being lodged between the pilings. The debris drift clearance is
shown on the Bridge Hydraulics Recommendation Sheet (BHRS).

5.1.4.2 Navigation
For crossings subject to small boat traffic, the minimum vertical navigation clearance is
set as six feet above the mean high water, normal high water, or control elevation.
Notably, other agencies may require different navigational clearances.

Chapter 5: Bridge Hydraulics 5-11


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

For tidally controlled or tidally influenced bridges, the BHR should document the tidal
datums for the bridge location. This includes not only the Mean High Water (MHW) for
use in navigational clearances, but also any other tidal datums available for the site. If
taken from a tidal bench mark, the BHR should document the bench mark ID as well as
the tidal epoch referenced.

Normal High Water is considered to be equivalent to the mean annual flood. The mean
annual flood is the average of the highest flood stage for each year. For gaged sites, you
can obtain this information from the U.S. Geological Survey (USGS). Statistically, the
mean annual flood is equivalent to the 2.33-year frequency interval (recurrence interval).
Therefore, if you use a synthetic hydrologic method to determine the Normal High Water,
use the 2.33-year event. In some cases, stain lines at the site indicating the normal flood
levels can be used to estimate the Normal High Water.

Obtaincontrol elevations from the regulating agency (Water Management Districts, water
control districts, U.S. Army Corps of Engineers, etc.).

5.1.4.3 Waves
Elevate coastal bridges one foot above the design wave crest, as required in the FDM
210. If the clearance is less than one foot, which often occurs near the bridge approaches,
you must design the bridge according to the Guide Specifications for Bridges Vulnerable
to Coastal Storms, a publication from the American Association of State Highway and
Transportation Officials (AASHTO).

5.1.5 Bridge Length Justification


It is seldom economically feasible or necessary to span the entire width of a stream at
flood stages. Where conditions permit, you can extend approach embankments onto the
flood plain to reduce costs, recognizing that in doing so the embankments will constrict
the flow of the stream during flood stages. Normally, this is an acceptable practice,
provided that the water surface profile and scour conditions are evaluated properly.

The BHR should demonstrate clearly that the proposed structure length and configuration
are justified for the crossing. Use historical records from the life of the bridge, along with
hydrologic and hydraulic calculations, to make recommendations. Using the same length
as an existing structure that may have been in place for many years is not justification to
use the same bridge length, given that the existing structure may not be hydraulically
appropriate and may not have experienced a significant flooding event.

The most effective way to justify the length of a proposed structure is with the analysis of
alternate structure lengths. Typical alternative bridge lengths that might be appropriate
include:

Chapter 5: Bridge Hydraulics 5-12


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

 Existing structure length


 Structure length that goes from bank to bank plus 20 feet to provide the minimum
maintenance berms
 Target velocity structure (for example, an average velocity through the bridge of 2
fps)
 Structure that spans the wetlands (the no-mitigation structure length)
 Concrete Box Culvert (CBC) structure
 Roadway geometrics structure length

As the analysis proceeds, the need to analyze another length may become apparent, and
that may turn out to be the proposed structure length.

5.1.6 Berms and Spill-Through Abutment Bridges


Normally, you would not place spill-through abutments in the main channel of a stream
or river for several reasons:

 Construction difficulties with placing fill and riprap below water


 Abutment slope stability during and after construction
 Increased exposure to scour
 Environmental concerns
 Stream stability or channel migration
 Maintenance

As stated in Section 4.9 of the Drainage Manual, you must determine the horizontal limit
of protection using the methods in HEC 23. However, a 10-foot width between the top of
the main channel and the toe of spill-through abutment slopes is considered the minimum
width necessary to address the above concerns. For stable banks, make the horizontal
10-foot measurement from the top edge of the main channel. The use of the minimum
berm width does not excuse the drainage engineer from conducting sufficient site analysis
to determine the existence of unusual conditions. If the natural channel banks are very
steep, unstable, and/or if the channel is very deep, or channel migration exists, additional
berm width may be necessary for proper stability. For these conditions, you should make
the horizontal 10-foot measurement from the point where an imaginary 1V:2H slope from
the bottom of the channel intersects the ground line in the floodplain.

In most situations, the structure that provides the minimum berm width often will be the
shortest bridge length considered as a design alternative.

The minimum abutment protection is stated in Section 4.9 of the Drainage Manual. The
standard rubble riprap was sized in accordance with HEC 23 for flow velocities (average)
not exceeding 9 fps, or wave heights not exceeding 3 feet. Determine the horizontal and

Chapter 5: Bridge Hydraulics 5-13


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

vertical extent using HEC 23. A minimum of 10 feet is recommended as a horizontal


extent if HEC 23 shows that a horizontal extent less than 10 feet is acceptable. Review
the limits of right of way to be sure the apron at the toe of the abutment slope can extend
out and along the entire length of the abutment toe, around the curved portions of the
abutment to the point of tangency with the plane of embankment slopes. If calculations
from HEC 23 show that the horizontal extent is outside the right-of-way limits, you can do
the following:

a. Recommend additional right of way.


b. Provide an apron at the toe of the abutment slope that extends an equal distance
out around the entire length of the abutment toe. In doing so, consider specifying
a greater rubble riprap thickness to account for reduced horizontal extent.

Chapter 5: Bridge Hydraulics 5-14


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.1-8: Limits of Rubble Riprap Protection

Figure 5.1-8 is a plan view that defines the limit of rubble riprap protection. Refer to the
FDOT Structures Detailing Manual for the recommended minimum distance.

Chapter 5: Bridge Hydraulics 5-15


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

In contrast, controlled canals in developed areas typically have very low velocities, no
stability problems, no overbank flow contracting into the bridge opening, and few
abutment maintenance problems. In such cases, the abutment slope usually drops
steeply from the abutment directly into the canal.

Use rubble with a specific gravity of 2.65 or other extra heavy revetment where large wave
attack is expected, typically in coastal applications. Avoid corrodible metal cabling or
baskets in coastal environments; even if coated, the coating may be marred and allow
corrosion. Follow the USACE Shore Protection Manual for design of coastal revetment.

Use bedding stone on all bank and shore rubble installations to guard against tearing of
the filter fabric during placement of the rubble. The bedding stone also helps dissipate
wave impacts on the revetment.

For revetment installations where wave attack is not expected to be significant, include
all options (e.g., fabric-formed concrete, standard rubble, or cabled interlocking block,
etc.) that are appropriate based on site conditions. All options shown to be inappropriate
for the site should be documented in the BHR. Write a technical specification based on
the use of the most desirable revetment material, with the option to substitute the other
allowable materials at no additional expense to the Department. This recommendation
will help in eliminating revetment Cost Savings Initiative Proposals (CSIPs) during
construction.

No matter what options are allowed, match the bedding (filter fabric and bedding stone)
to the abutment material. Some of the options are not self-healing, and a major failure
can occur if loss of the embankment material beneath the protection takes place.

5.1.7 Design Considerations for Dual Bridges


When two-lane roadways are upgraded to multi-lane divided highways, the existing bridge
on the existing roadway often has many years of remaining life. So a new dual bridge is
built next to the existing bridge. Years later, when the original bridge needs to be replaced,
the newer bridge still has years of remaining life. So a cycle of replacing one of the dual
bridges at a time is repeated. There is a tendency to keep the bridge ends aligned with
the bridge remaining in place. However, consider the potential for lateral migration of the
stream, and plan that the new bridge end locations should accommodate the stream.

Scour estimates must consider the combined effects of both bridges. Ideally, the
foundation of the new or replacement bridge will be the same type as the other foundation
and will be aligned with the other foundation. In such cases, the scour calculations will be
similar to that of a single bridge.

Chapter 5: Bridge Hydraulics 5-16


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

In some cases, it may not be reasonable to match and align the foundations of both
bridges because of such things as economics, geotechnical considerations, and channel
migration, etc. If the foundation designs are not the same, or are not aligned, or both, the
scour estimates must consider the combined obstruction of both foundations to the flow.
The techniques of HEC 18 do not specifically address this situation. If another approach
is not available, assume a single foundation configuration that accounts for the
obstruction of both foundations and use the techniques of HEC 18. You can develop a
conservative configuration by assuming each downstream pile group is moved upstream
(parallel to flow) a sufficient distance to bring it in line with the adjacent upstream pile
group. Figure 5.1-9 shows some configurations.

Figure 5.1-9: Configurations for Computing Scour of Dual Bridges

Chapter 5: Bridge Hydraulics 5-17


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.1.8 Design Considerations for Bridge Widenings


The new substructure or foundations under the widened portion of a bridge often are
different than the existing substructure in shape or depth. If a bridge has been through
the Statewide Bridge Scour Evaluation Process and, as a part of that process, has been
identified as "scour critical,” the existing foundation must accommodate the predicted
scour. If the existing foundation design cannot accommodate the predicted scour, the first
alternative is to reinforce the existing foundation so that it can. If it is not practical to
reinforce the existing foundation, the next alternative is to replace the existing structure
so that it can be removed from the scour critical list. These approaches are consistent
with the goal to remove all bridges from the scour critical list.

For minor widening (defined in Chapter 6 of the FDOT Structures Design Guidelines) of
bridges that have been through the Statewide Bridge Scour Evaluation Process and have
not been identified as scour critical, it is acceptable to leave the existing foundation
without modification. The foundation under the widened portion must be properly
designed to accommodate the predicted scour.

Widening existing bridges often will result in a minor violation of vertical clearances due
to the extension of the cross slope of the bridge deck. Consult the District Drainage
Engineer in documenting justification for deviating from criteria.

Structural Pier Protection Systems

Dolphins and fender systems are two structural systems designed to protect piers,
bents, and other bridge structural members from damage due to collision by marine
traffic. Dolphins are large structures with types ranging from simple pile clusters to
massive concrete structures that can either absorb or deflect a vessel collision.
Typically, they are located on both sides of the structure being protected, as shown in
Figure 5.1-10. Fender system types are less variable, consisting usually of pile-
supported wales, as shown in Figure 5.1-11. Fender systems typically wrap around the
protected piers and run along the main navigation channel.

Chapter 5: Bridge Hydraulics 5-18


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Dolphins

Figure 5.1-10: Dolphin Pier Protection at the Sunshine Skyway Bridge

Chapter 5: Bridge Hydraulics 5-19


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Fender System

Figure 5.1-11: Fender System at the Old Jewfish Creek Bridge

For design purposes, you can calculate scour around dolphins in the same manner as
bridge piers. Typically, dolphins are located sufficiently far from the piers so that you can
calculate local scour independently. However, check to ensure there is sufficient spacing
(greater than 10 effective diameters).

Scour at fender systems typically is taken as equal to that of the pier it is protecting. In
some cases, fender systems may “shield” bridge piers, reducing velocities and scour at
the pier. However, this shielding effect can vanish or be modified if the fender system is
lost due to collision or unforeseen scour problems, or if the flow attack angle is skewed
so that the pier is not in the hydraulic shadow of the fender system. Piers and fender
systems introduced into relatively narrow rivers may cause contraction scour between the
fender systems. This scour usually is greatest near the downstream end of the system.

Chapter 5: Bridge Hydraulics 5-20


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.2 RIVERINE ANALYSIS


A riverine analysis applies to inland streams and rivers. Flooding conditions for riverine
systems result from runoff from extreme rainfall events. Steady-state flow conditions
usually can be assumed.

5.2.1 Data Requirements


The data collected will vary depending on the site conditions and the data available. Two-
dimensional models require substantially more data than one-dimensional models.

5.2.1.1 Geometric Data


Follow these steps to collect geometric data for the analysis:

1. Determine the model domain. The geometric data must extend far enough
upstream, downstream, and laterally to provide an accurate representation of the
terrain within the domain. Refer to Section 5.2.4 for guidance.
2. Locate available geometric data within the model domain. You can use liberally
estimated boundaries of the domain when the cost of collecting existing data is
low.
3. Order survey for those portions of the model domain that do not have adequate
coverage from existing geometric data. Survey will be expensive, so estimate the
domain boundaries conservatively.

Existing Geometric Data


There are many potential sources of geometric data, and new sources of data continually
become known. The following is a list of potential sources:

 USGS
o Quadrangle maps
o A public source in both scanned and vector formats is the FDEP Land Boundary
Information System (LABINS) located at: http://www.labins.org/
o Digital Elevation Models (DEMs)
 DEMs are essentially x, y, z coordinate points on a 90-meter grid. They were
derived from the Quadrangle Maps. DEMs also are available at LABINS.
o LiDAR
 Coverage in Florida is not yet complete. Available data can be downloaded at:
https://lta.cr.usgs.gov/lidar_digitalelevation
 U.S. Army Corps of Engineers
o USACE performs hydrographic surveys on navigable waterways, which can provide
main channel information.
o Mobile District: http://navigation.sam.usace.army.mil/surveys/index.asp

Chapter 5: Bridge Hydraulics 5-21


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

o Jacksonville District:
http://www.saj.usace.army.mil/Missions/CivilWorks/Navigation/HydroSurveys.aspx
 Florida Department of Emergency Management
o Data for the Florida Coastal LiDAR project and links to other compatible data:
http://www.floridadisaster.org/gis/lidar/
 Water Management Districts
 Cities and counties
 Old plans and BHRs
 FEMA studies
o Refer to Section 5.2.1.1 for more information on how to determine if a detailed study
is available.

USGS Quadrangle Maps and DEMs are available for the entire state of Florida. They may
be useful for preliminary analysis and, in some circumstances, you may use them to fill in
gaps farther away from the site.

The remaining data sources usually will have a level of accuracy that was adequate for
hydraulic modeling at the time of collection. However, consider the age of the data. If the
terrain within the model domain has changed significantly, then you must find newer
existing data sources or you will need to order survey.

You may need data from different sources to cover the entire model terrain. Sometimes,
one source will have data within the overbank and floodplain areas, and a different source
will have hydrographic data within the channel. Be sure to convert all data to a common
datum and projection.

Ordering Survey Data


The FDOT Surveying Handbook (dated October 31, 2003) states that bridge survey and
channel survey requirements are project specific. You will need to provide site-specific
instructions to the surveyors so that they do not default to the previously used Location
Survey Manual.

Survey can be in either cross section or Digital Terrain Model (DTM) format for one-
dimensional models. Although you can use cross sections to develop two-dimensional
models, a DTM format is preferable. Discuss the survey format with the surveyor to
determine which format is most appropriate.

Always order survey in the immediate vicinity of the proposed bridge. The accuracy needs
in this area are greater than the accuracy needs of the hydraulic model, for two reasons:

1. Bridge and roadway construction plans need a higher degree of accuracy.

Chapter 5: Bridge Hydraulics 5-22


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

2. The approach roadway and bridge abutment, including abutment protection, must
fit within the right of way.

The typical roadway survey will be a DTM within the proposed right of way, and may
extend a minimal distance outside of the proposed right of way. Coordinate with the
roadway design engineer.

Determine the location of the approach and exit cross sections for the model and extend
survey information in the main channel to these locations. Additional survey information
in the adjacent floodplain and farther upstream and downstream of these extents will
depend upon the other available geometric data.

Provide a sketch to the surveyor on a topographic map or aerial showing the limits of the
DTM or the location, orientation, and length of cross sections. Also ask the surveyor for:

 Survey(s) of any adjacent utility crossings


 Elevations of stains on the existing pilings
 Any high water marks determined by the hydraulics engineer during the site visit
 Elevation of the water level on the day of the survey

When ordering survey, remember that most floodplains in Florida often have dense
vegetation. Surveying in these areas will be difficult. Not all cross sections need to be
surveyed at the actual location used in the hydraulic model. Surveyed cross sections can
be reasonably manipulated into model cross sections, so look for areas that would be
easier to survey, such as along power lines and open fields.

5.2.1.2 Geotechnical Data


Geotechnical information is required at bridge foundations to establish the bed
composition and its resistance to scour. Near surface bed materials in Florida range from
sand and silts to clays to rock. As will be discussed in Section 5.5, the composition of the
bed material dictates the procedure employed in the calculation of scour. For scour
studies, the required information is a characterization of the near surface bed material,
i.e., the layer over which scour will occur. The thickness of this layer will be a function of
the expected scour at the site.

For bridges with foundations in cohesionless sediments (sands and silts), include sieve
analyses in the geotechnical data collection to characterize the size of the bed sediments.
Obtain a sufficient number of samples to confidently characterize the sediment size, both
over the length of the bridge as well as over the thickness of the expected scour layer.
The parameter from the sieve analyses necessary for scour calculation is the median

Chapter 5: Bridge Hydraulics 5-23


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

grain size (D50). NRCS soil surveys can provide an estimated median grain size for
preliminary scour estimates.

For bridges with foundations in cohesive sediments (rock or clay), establish the bed
material’s scour resistance. For rock, the FHWA provides guidelines for scourability of
rock formations in HEC 18 (refer to Chapter 4).

Additionally, the Department has developed a Rock Scour Protocol, which you can find
at:
http://www.fdot.gov/roadway/Drainage/Bridgescour/Bridge-Rock-Scour-Analysis-Protocol-Jan2008.pdf.
The referenced protocol recommends obtaining core borings at each pier for testing at
the State Materials Office. It is your responsibility to follow the protocol procedure when
encountering soils of this type.

For smaller streams where a bridge culvert may be an appropriate hydraulic option,
consider obtaining a preliminary soil boring to determine if increased foundation costs for
the culvert need to be included in the alternatives cost comparisons.

5.2.1.3 Historical Data


Historical data provide important information for many aspects of the bridge hydraulics
and scour analysis. They provide numbers for calibration through gage measurements
and historical high water marks, data for calculation of long-term scour processes through
historical aerial photography and Bridge Inspection Reports, and characterization of the
hurricane vulnerability through the hurricane history.

Speak with local residents, business owners and employees, and local officials—
including fire and emergency services—to obtain anecdotal information about past floods.
This information can be very important in the absence of other historical data.

Gage Measurements
In bridge hydraulics analysis, you can use gage data in a number of ways:

 To determine the peak flow rates, although the Department usually relies upon
agencies, such as the USGS, to perform statistical analysis of the stream flow data.
Refer to Section 2.2 (Hydrology) for more information.
 To provide starting water surface elevations, or boundary conditions, for the model
if the gage is downstream of the bridge. Refer to Section 5.2.4.9 for more
information.
 To calibrate the model. Refer to Section 5.2.5.1 for more information.

Chapter 5: Bridge Hydraulics 5-24


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

If the gage is located at a distance from the bridge site, the gage flow rates may not be
the same as the bridge flow rates. However, the gage data still may be useful if the flow
rates can be adjusted. Refer to Section 4.5, Peak Flow Transposition in FHWA Highway
Hydrology, Hydraulic Design Series 2 (HDS-2), for more information.

USGS gage information can be found at this website: http://fl.water.usgs.gov/

Gage data also may be available from the Water Management Districts and other local
agencies.

Historical Aerial Photographs


Historical aerial photographs provide a means to determine the stream stability at a
highway crossing. Comparison of photographs over a number of years can reveal long-
term erosion or accretion trends of the shorelines and channel near the bridge crossing.
You also can use current aerial photographs as a base for figures in the Bridge Hydraulics
Report, showing such things as cross section locations and upstream and downstream
controls.

Recent and current aerial photographs can be found at many Internet sites. Be careful of
copyright infringements when using these aerials in the Bridge Hydraulics Report. For this
reason, it is probably best to obtain the photographs from government sites that give free
access.

Older aerial photographs can be obtained from the Aerial Photography Archive Collection
(APAC), maintained by the FDOT Surveying and Mapping Office. APAC archives aerials
dating back to the 1940s. Ordering information is available at the following link:

http://www.dot.state.fl.us/surveyingandmapping/aerialmain.shtm

The University of Florida also maintains a database of older aerial photographs:

http://ufdc.ufl.edu/aerials

Another useful site to obtain aerial photography is the FDEP Land Boundary Information
System (LABINS), which can be accessed at the following link:

http://www.labins.org/

Chapter 5: Bridge Hydraulics 5-25


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Existing Bridge Inspection Reports


The District Structures Maintenance Office is responsible for the inspection of each bridge
in the state at regular time intervals, including bridges owned by local agencies. The
reports will document any observed hydraulically related issues, such as scour or erosion
around the piers or abutments. Obtain Bridge Inspection Reports from the District
Structures Maintenance Office. Of particular interest will be the channel profiles that have
been collected at the site, which may show channel bottom fluctuations over time.

Channel profiles usually are created by taking soundings from the bridge deck. Soundings
are measurements taken using a weighted tape measure to keep the tape vertical. The
measurements are the distance from a consistent point on the bridge (usually the bridge
rail) to the stream bed. The measurements are made on both sides of the bridge at each
bridge pier and often at mid-span.

You may be able to find the Phase 1 Scour Evaluation Report for existing bridges. This
report will plot some of the bridge inspection profiles against the cross section from the
original construction, assuming that old plans or pile driving records were available to
obtain the original cross section. The example bridge shown in Figures 5.2-1 and 5.2-2
has a very wide excavated cross section beneath the bridge. This was a common bridge
design practice before dredge and fill permitting requirements brought the practice to an
end unless the required wetland impact was justified and mitigated. In the example, the
widened channel has filled back in and narrowed since the initial construction in 1963.

You can use the channel profiles to determine long-term bed changes at the bridge site.

Chapter 5: Bridge Hydraulics 5-26


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.2-1: Example Bridge Profile from a Bridge Inspection Report

Chapter 5: Bridge Hydraulics 5-27


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Chapter 5: Bridge Hydraulics 5-28


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.2-1: Example Bridge Profile from a Bridge Inspection Report (cont.)

Chapter 5: Bridge Hydraulics 5-29


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.2-2: Excerpt from Scour Evaluation Report

Chapter 5: Bridge Hydraulics 5-30


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.2-2: Excerpt from Scour Evaluation Report (continued)

Chapter 5: Bridge Hydraulics 5-31


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.2-2: Excerpt from Scour Evaluation Report (continued)

Chapter 5: Bridge Hydraulics 5-32


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Previous Studies
If the project replaces or widens an existing bridge, obtain the BHR or other hydraulic
calculations for the existing bridge, if possible. Other BHRs for bridges over the same
water body also may provide useful information.

If a detailed study was performed by FEMA, then obtain the Flood Insurance Study, the
NFIP Maps, and the original model (refer to Section 5.2.1.5).

Additional sources of existing studies can include the Water Management Districts, the
Florida Department of Environmental Regulation, county offices, and the U.S. Army Corps
of Engineers.

Maintenance Records
Contact the local district or local agency maintenance staff for bridge inspection reports,
historical overtopping, and/or maintenance issues at the bridge site.

5.2.1.4 Drainage Basin Information


You will need drainage basin information for the hydrologic analysis. The type of
information collected depends upon the hydrologic method used in the analysis. Refer to
Section 5.2.2 and Chapter 2 (Hydrology) for guidance on the hydrologic analysis and data
requirements.

Delineate the drainage basin boundaries on the Bridge Hydraulics Recommendation


Sheet. Federal, state, and local agencies—including the Water Management Districts—
often publish basin studies and delineate basin areas. Many of these are available online.
Verify the boundaries found on older maps.

You also should gather information on other structures on the river upstream and
downstream of the proposed bridge site, including the size and type of structure for
comparison with the proposed structure.

5.2.1.5 FEMA Maps


Obtain the FEMA Flood Insurance Rate Map and the Flood Insurance Study for the site.
You can order these maps or download them from the FEMA Map Service Center at the
following link:
http://msc.fema.gov/

Backup and supporting data for a detailed study, if the area has a detailed study, also can
be obtained from FEMA. At the time of this writing, this information cannot be ordered

Chapter 5: Bridge Hydraulics 5-33


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

through the website. A data request form must be completed and sent to FEMA. Call the
FEMA Map Service Center for ordering information.

https://www.fema.gov/media-library/assets/documents/7320

5.2.1.6 Upstream Controls


Upstream controls may influence the discharge at the crossing. Pump stations and dams
are two common controls. Salinity intrusion structures are another example. Contact the
agency exercising control over these structures to obtain information regarding
geometrics, intended mode of operation, flow rate data, and history, including structure
failures. It is important to consider the likelihood of upstream structure failures when
considering flow regimes. A dam break analysis may be appropriate.

5.2.1.7 Site Investigation


A field investigation is recommended for all new bridge construction. Data obtained during
a field investigation can aid in hydraulic model construction, identify problem erosion
areas, and characterize stream stability. Perform a field investigation during the early
stages of design. The following checklist (Neill, 1973) outlines some key items of basic
data to be collected (not all may apply to a particular site):

 Look for channel changes and new tributaries compared to the latest aerial
photographs or maps from the office data collection
 Look for evidence of scour in the area of the existing structure and check the
adequacy of existing abutment protection
 Check for recent repairs to the existing abutment protection (as compared with
the age of the bridge)
 Check for local evidence of overflow or breaching of the approaches
 Search the site for evidence of high flood levels, debris, or stains on the structure
that may indicate flood levels
 Search for local evidence of wave-induced erosion along the banks
 Note the velocity direction through the bridge and estimate the velocities (note
the date and time of these observations)
 Photograph the channel and adjacent areas
 Seek evidence of the main overflow routes and flood relief channels
 Search for hydraulic control points upstream and downstream of the structure
 Assess the roughness or flow capacity of the floodplain areas

Chapter 5: Bridge Hydraulics 5-34


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

 Describe and photograph the channel and overbank material in situ


 Seek evidence on largest size of stone moved by flood or waves
 Seek local evidence of channel shifting, bank and shore erosion, etc., and their
causes
 Seek local evidence of channel bed degradation or aggradation
 Seek evidence of unrecorded engineering works that would affect flows to the
bridge, such as dredging, straightening, flow diversions, etc.
 Observe the nearby land uses that might be affected by flood level changes

Consider visiting other structures across the stream or river upstream and downstream
of the proposed bridge site.

5.2.2 Hydrology
In most riverine analyses, you can assume steady-state conditions and perform the
hydraulic analysis using the peak discharge for each frequency analyzed. The peak
discharge may vary at different locations on the stream if there are tributaries within the
reach, but each discharge will be assumed to remain constant with respect to time.

Section 4.7.1 of the Drainage Manual gives criteria for selecting discharges used for
riverine analysis. Further guidance is given in Chapter 2 (Hydrology).

Generally, the length of the structure does not control the hydrology. That is, in general,
a longer structure will not significantly increase the discharge downstream. When
considering the inaccuracies associated with the hydrology, the effect of the structure
length and the resulting backwater (or reduction of backwater) usually will not significantly
affect the amount of water going downstream. However, if you or the regulatory agency
are significantly concerned about this effect, then you should conduct an analysis to verify
the concern. You can calculate the pre- and post-water surface profiles and route them
with an unsteady flow model.

5.2.3 Model Selection


Before selecting a specific model to use at a given bridge site, you must make two general
decisions to isolate groups of appropriate models.

The two basic decisions are:

1. One-dimensional or two-dimensional model?


2. Steady flow conditions or unsteady flow conditions?

Chapter 5: Bridge Hydraulics 5-35


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.2.3.1 One-Dimensional versus Two-Dimensional


The accuracy of a one-dimensional model depends upon the ability of the modeler to
visualize the flow patterns during the design events to properly locate the model cross
sections. Complicated flow patterns caused by site factors such as skewed approach
embankments, multiple openings, other nearby crossings, and the presence of bends,
meanders, and confluences within the reach, may indicate that a two-dimensional model
would be more appropriate.

5.2.3.2 Steady versus Unsteady Flow


Use an unsteady flow model for the following conditions:

 Mild stream slopes less than two feet per mile. If the slope is greater than five feet
per mile, steady flow can be used. For slopes between these values, consider the
cost and complexity of an unsteady model versus the cost importance of the
bridge.
 Situations with rapid changes in flow and stage. Models of dam breaks are the
primary example of this situation.
 Bifurcated streams (streams where the flow divides into one or more channels and
recombines downstream).

You can find more information on these situations in USACE Manual EM 1110-2-1416
(15 October 1993), River Hydraulics.

5.2.3.3 Commonly Used Programs


The most commonly used one-dimensional models are HEC-RAS and WSPRO. HEC-
RAS was developed by the U.S. Army Corps of Engineers Hydrologic Engineering Center
for a number of river hydraulic modeling applications, including the hydraulic design of
waterway bridges. WSPRO (Water Surface PROfile) is the acronym for the computer
program developed by FHWA specifically for the hydraulic design of waterway bridges.
Make sure you are using the latest version and document the version in the Bridge
Hydraulics Report.

HEC-RAS and WSPRO both are suitable to analyze one-dimensional, gradually varied,
steady flow in open channels, and you also can use them to analyze flow through bridges
and culverts, embankment overflow, and multiple-opening stream crossings. HEC-RAS
has the additional capability of analyzing unsteady flow.

The WSPRO program analyzes unconstricted valley sections using the standard step
method, and incorporates research for losses across a bridge constriction. HEC-RAS

Chapter 5: Bridge Hydraulics 5-36


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

allows the user to select the method used to analyze the bridge losses, including energy
(standard step), momentum, Yarnell, and WSPRO methods. Both programs allow you to
readily analyze alternate bridge openings. The output provides water surface elevations,
bridge losses, and velocities for both the constricted (with bridge) and the unconstricted
(with no bridge) condition. You can use this information to estimate the backwater effects
of the structure and provide input information for scour analysis.

The most commonly used two-dimensional models are FESWMS and RMA 2. The Finite
Element Surface Water Modeling System (FESWMS) was developed originally for FHWA
and the USGS. The FHWA has continued to maintain and sponsor development of
subsequent versions, which continue to incorporate features specifically designed for
modeling highway structures in complex hydraulic environments. As such, it includes
many features that other available two-dimensional models do not have, such as pressure
flow under bridge decks, flow resistance from bridge piers, local scour at bridge piers,
live-bed and clear-water contraction scour at bridges, bridge pier riprap sizing, flow over
roadway embankments, flow through culverts, flow through gate structures, and flow
through drop-inlet spillways. FESWMS can perform either steady-state or unsteady flow
modeling.

The Resource Management Associates software RMA 2 is a two-dimensional, unsteady,


depth-averaged, finite-element, hydrodynamic model. It computes water surface
elevations and depth-averaged horizontal velocity for subcritical, free-surface flow in two-
dimensional flow fields. The program contains the capability of solving both steady- and
unsteady-state (dynamic) problems. Model capabilities include: wetting and drying of
mesh elements, including Coriolis effects; applying wind stress; simulating five different
types of flow control structures; and applying a wide variety of boundary conditions.
Applications of the model include calculating: water levels and flow distribution around
islands; flow at bridges having one or more relief openings; in contracting and expanding
reaches; into and out of off-channel hydropower plants; at river junctions; and into and
out of pumping plant channels; circulation and transport in water bodies with wetlands;
and general water levels and flow patterns in rivers, reservoirs, and estuaries.

5.2.4 Model Setup


You will need the following data to perform the hydraulic and scour analysis for a bridge
crossing:

 Geometric data
 Flow data (upstream boundary)
 Loss coefficients
 Starting water surface elevations (downstream boundary)
 Geotechnical data (D50 soils information)

Chapter 5: Bridge Hydraulics 5-37


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.2.4.1 Defining the Model Domain


You will need upstream, downstream, and lateral study boundaries to define the limits of
data collection. The model must begin far enough downstream to assure accurate results
at the bridge, and far enough upstream to determine the impact of the bridge crossing on
upstream water surface elevations. The lateral extent should ensure that the model
includes the area of inundation for the greatest flood analyzed. Underestimating the
domain causes the water surface calculations to be less accurate or requires additional
survey at a higher cost than the inclusion in the initial survey. Overestimation results in
greater survey, data processing, and analysis cost.

Upstream
At a minimum, the upstream boundary should be set far enough upstream of the bridge
to encompass the point of maximum backwater caused by the bridge. If a point of concern
where the water surface elevation must be known is farther upstream, then the model
must be extended to that point. An example would be upstream houses or buildings
because the 100-year water surface elevation must be kept below their floor elevation.
Check with permitting agencies, including cities and counties, as some have limits on the
amount of backwater allowed at a given distance upstream.

The following equation can be used to determine how far upstream data collection and
analysis needs to be performed.

Lu = 10,000 * HD0.6 * HL0.5/S

where:
Lu = Upstream study length (along main channel), in feet for normal depth starting
conditions
HD = Average reach hydraulic depth (1-percent chance flow area divided by cross
section top width), in feet
S = Average reach slope, in feet per mile
HL = Head loss, ranging between 0.5 feet and 5.0 feet at the channel crossing
structure for the 1-percent chance flow

The values of HD and HL may not be known precisely since the model has not yet been
run to determine these values. They can be estimated from FEMA maps, USGS
Quadrangle Maps (or other topographic information).

Chapter 5: Bridge Hydraulics 5-38


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.2-3: Open Channel Depth Profiles

Downstream
Open channel hydraulics programs must have a starting water surface elevation specified
by the user at the downstream boundary of the model.

The programs allow for one or more of the following methods of specifying the starting
water surface elevation:

 Enter a water surface elevation at the downstream boundary.


 Enter a slope at the downstream boundary, which is used to calculate the normal
depth from Manning’s Equation.
 Assume critical depth at the downstream boundary.

The modeler must decide which method to use, and the decision will affect the distance
to the downstream boundary of the model.

Chapter 5: Bridge Hydraulics 5-39


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

For the storm frequency being modeled, if a point of known water surface elevation is
within a reasonable distance downstream, extend the model to that point. Refer to the
section below on convergence for guidance on determining if the point is within a
reasonable distance.

Gages are points with a known relationship between the discharge and the water surface
elevation. Lakes and sea level also can be points of known elevation. Other locations
where you can calculate the water surface elevation from the discharge can include weirs,
dams, and culverts if these locations are not significantly influenced by their tailwater.

When the downstream channel and overbank are nearly uniform, use the normal depth
assumption to determine the starting water surface elevation, both in cross section and
slope, for a long reach downstream. The length of uniform channel that will be adequate
will vary with the slope and properties of the channel. This reach should not be subject to
significant backwater from farther downstream. The following equation can be used to
determine how far downstream data collection and analysis needs to be performed.

Ldn = 8,000 * HD0.8/S

where:
Ldn = Downstream study length (along main channel), in feet for normal depth starting
conditions
HD = Average reach hydraulic depth (1-percent chance flow area divided by cross
section top width), in feet
S = Average reach slope, in feet per mile

Make some sound engineering judgment when determining the variables HD, S, and HL.
Guidelines are presented below:

a. Average reach hydraulic depth (HD): If limited existing data are available, an
estimate can be made using FEMA maps and quadrangle maps. Using the FEMA
map, outline on the Quadrangle Map the boundary of the 1-percent chance flow.
Select a representative location and plot a cross section using the Quadrangle
Map. Plotting several cross sections may improve the estimate. The area (A), top
width (TW), and, thus, the hydraulic depth (A/TW) for these cross sections are now
determined. Average these hydraulic depths to determine an average reach
hydraulic depth. Use survey data or other existing geometric data that are more
accurate than the Quadrangle Maps if available.
b. Average reach slope (S): Using the Quadrangle Maps, determine and average
the slope of the main channel, left overbank, and right overbank.
c. Head loss (HL): This term also is known as the "backwater.” Backwater is defined
as the difference in the water surface elevation between the constricted (bridge)

Chapter 5: Bridge Hydraulics 5-40


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

flow condition and the unconstricted (no bridge) flow condition at a point of interest
upstream of the structure crossing. Make an educated guess at the anticipated
head loss. For a new bridge, the allowable head loss would be a reasonable
estimate. In most cases, a maximum head loss of one foot would be expected for
Florida.

Lateral Extents
Extend the model laterally on both sides of the floodplain to an elevation that is above the
highest water surface elevation that will be modeled. Often, this water surface elevation
will be unknown until the model is complete. But you must collect data to complete the
model. So, you must estimate the water surface elevation and lateral extent for the data-
gathering effort. You can estimate the elevation or the lateral extent from FEMA maps
and other historical studies of the site. In some cases, it is appropriate to set up a
preliminary model based on limited data to estimate the water surface elevations.
Whichever method you use to estimate the lateral extent of the model, consider making
a conservative estimate to avoid additional data gathering at a later time, especially
survey data.

5.2.4.2 Roughness Coefficient Selection


You can use a number of references to select Manning's Roughness Coefficient within
the main channel and overbank areas of riverine waterways. Two recommended
references are:

1. Guide for Selecting Manning’s Roughness Coefficients for Natural Channels and
Flood Plains, USGS Water-Supply Paper 2339 (replaces Report Number FHWA-
TS-84-204), which can be accessed at the following link:
http://www.fhwa.dot.gov/bridge/wsp2339.pdf
2. Estimating Manning's Roughness Coefficients for Natural and Man-Made Streams
in Illinois, USGS and Illinois Department of Natural Resources.
http://il.water.usgs.gov/proj/nvalues/
Roughness values from previous models or studies can be useful. However, you should
verify these roughness values because conditions may have changed.

Roughness values can be varied within reasonable limits representative of the physical
conditions of the site to calibrate the hydraulic model.

Chapter 5: Bridge Hydraulics 5-41


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.2.4.3 Model Geometry


Model selection is discussed in Section 5.2.3. This section discusses the creation of one-
and two-dimensional models.

One-Dimensional Models
One-dimensional models use cross sections to define the geometry of the channel and
floodplain. There are several good references you can use as guidelines to locate and
subdivide the cross sections. One good source is Computation of Water-Surface Profiles
in Open Channels, by Jacob Davidian: USGS—Techniques of Water-Resources
Investigations Reports, Book 3, Chapter A15, 1984. This publication can be downloaded
from: http://pubs.usgs.gov/twri/twri3-a15/pdf/twri_3-A15_b.pdf.

Some of the guidelines presented below are from this reference.

a. Take cross sections where there is an appreciable change in slope.


b. Take cross sections where there is an appreciable change in cross sectional area
(i.e., minimum and maximum flow areas).
c. Space cross sections around abrupt changes in roughness to properly average the
friction loss between the sections. One method is to evenly space cross sections
on either side of the abrupt change. Refer to the spacing between XSEC1 and
XSEC2 and between XSEC3 and XSEC4 in Figure 5.2-4 as an example. Another
method is to locate a section at the abrupt change. Include the cross section twice,
separated by a short flow length (maybe 0.1 foot), and using the two different
roughness values as appropriate.

Chapter 5: Bridge Hydraulics 5-42


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.2-4: Example Cross Section Spacing

d. Take cross sections normal to the flood flow lines. In some cases, you may need
to “dog leg” cross sections. Figure 5.2-5 illustrates this procedure.
e. Place cross sections at closer intervals in reaches where the conveyance changes
greatly as a result of changes in width, depth, or roughness. The relation between
upstream conveyance, K1, and downstream conveyance, K2, should satisfy the
criterion: 0.7<(K1/K2)<1.4.
f. Avoid areas with dead flow, eddies, or flow reversals.
g. Extend cross section ends higher than the expected water surface elevation of the
largest flood that is to be considered in the sub-reach.
h. Place cross sections between sections that change radically in shape, even if the
two areas and the two conveyances are nearly the same.
i. Place cross sections at shorter intervals in reaches where the lateral distribution
of conveyance in a cross section changes radically from one end of the reach to
the other, even though the total areas, total conveyance, and cross sectional shape
do not change drastically. Increasing the number of subdivisions generally will
increase the value of alpha, and, therefore, increase the velocity head. Spacing
the cross sections closer together will help prevent drastic changes in the velocity
head.

Chapter 5: Bridge Hydraulics 5-43


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

j. Locate cross sections at or near control sections.


k. Locate cross sections at tributaries that contribute significantly to the main stem.
The cross sections should be placed such that the tributary enters the main stem
in the middle of the sub-reach.

Chapter 5: Bridge Hydraulics 5-44


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.2-5: “Dog Legging” Cross Section

Chapter 5: Bridge Hydraulics 5-45


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Subdivisions of cross sections should be done primarily for major breaks in cross-
sectional geometry. Major changes in the roughness coefficient also may call for more
subdivisions.

Figures 5.2-6 and 5.2-7 show guidelines on when to subdivide.

Figure 5.2-6: Subdivision Criteria of Tice (written communication, 1973)

Chapter 5: Bridge Hydraulics 5-46


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.2-7: Subdivision Criteria of Tice (written communication, 1973)

(A) Conveyance
Conveyance is a measure of the ability of a channel to transport flow. In streams of
irregular cross section, it is necessary to divide the water area into smaller but more or
less regular subsections, assigning an appropriate roughness coefficient to each, and
computing the discharge for each subsection separately. By rearranging the Manning’s
Equation, the following relationship is derived:

q 1.49 2/3
k= 1/2
= ar (Equation 1)
S n

Chapter 5: Bridge Hydraulics 5-47


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

where:
k= Channel subsection conveyance
q= Subsection discharge, in cubic feet per second
S = Channel bottom slope, in feet/feet
n= Manning's roughness coefficient
a= Subsection cross-sectional area, in square feet
r= Subsection hydraulic radius, in feet

Conveyance can, therefore, be expressed either in terms of flow factors or strictly


geometric factors. In bridge waterway computations, conveyance is used as a means of
approximating the distribution of flow in the natural river upstream from a bridge. Total
conveyance (K) is the summation of the individual conveyances comprising the particular
section. Example 5.2-1 illustrates a conveyance computation of a subdivided cross
section.

Example 5.2-1—Computing Conveyance

a. Compute the conveyance for the cross section shown above.

Solution:

Step 1: Compute the area, hydraulic radius, and conveyance for each of
the subareas:

Chapter 5: Bridge Hydraulics 5-48


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Subarea 1:

a1 = 10 ft. * 2 ft = 20 ft2
wp1 = 10 ft. + 2 ft. = 12 ft.
r1 = a1/wp1 = 20 ft2/12 ft. = 1.67 ft.

1.49 1.49
k1=
2/3
a1 r 1 = (20 ft .2 ) (1.67 ft. )2/3 = 419.5
n1 0.1
Subarea 2:

a2 = 40 ft. * 7 ft. = 280 ft2


wp2 = 40 ft. + 5 ft. + 5 ft. = 50 ft.
r2 = a2/wp2 = 280 ft2/50 ft. = 5.60 ft.

1.49 1.49
k2=
2/3
a2 r 2 = (280 ft .2 ) (5.60 ft. )2/3 = 32890.9
n2 0.04

Subarea 3:

a3 = 10 ft. * 2 ft. = 20 ft2


wp3 = 10 ft. + 2 ft. = 12 ft.
r3 = a3/wp3 = 20 ft2/12 ft. =1.67 ft.

1.49 1.49
k3 =
2/3
a3 r 3 = (20 ft .2 ) (1.67 ft. )2/3 = 419.5
n3 0.1

Total Conveyance (Ktotal) = k1 + k2 + k3


= 419.5 + 32,890.9 + 419.5
= 33,729.9

b. Assuming the total discharge for the water surface elevation of 107.0 feet in part
(a) is 4,000 cubic feet per second, determine the discharge distribution for each
subarea.

Solution:

Subarea 1:

Chapter 5: Bridge Hydraulics 5-49


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

k1 419.5
Q1 = * Qtotal = ( )* 4000 ft 3 /s = 49.8 ft 3 /s
k total 33729.9

Subarea 2:

k2 32890.9
Q2 = * Qtotal = ( )* 4000 ft 3 /s = 3900.5 ft 3 /s
k total 33729.9

Subarea 3:

k3 419.5
Q3 = * Qtotal = ( )* 4000 ft 3 /s = 49.8 ft 3 /s
k total 33729.9

(B) Velocity Head


The velocity head represents the kinetic energy of the fluid per unit volume and is
computed by:

 Q2
hv =
2g A2
where:

Q= Discharge at the section, in cubic feet per second


hV = Velocity head, in feet
= Kinetic correction factor for nonuniform velocity distribution
A= Total cross sectional flow area, in square feet

As the velocity distribution in a river varies from a maximum at the deeper portion of the
channel to essentially zero along banks, the average velocity head, computed as
(Q/A1)2/2g, does not a give a true measure of the kinetic energy of the flow. You can
obtain a weighted average value of the kinetic energy by multiplying the average velocity
head above by a kinetic energy coefficient (1) defined as:

 (qv 2)
1 = 2
Qv1

Chapter 5: Bridge Hydraulics 5-50


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

where:
1 = Kinetic energy coefficient, before the bridge
q= Discharge in a subsection, in cubic feet per second
v= Average velocity in same subsection, in feet per second
Q = Total river discharge, in cubic feet per second
v1 = Average velocity in river at Section 1, or Q/A1, in feet per second

Typical values of velocity coefficient, α, are shown in Table 5.2-1:

Table 5.2-1: Typical Values of Velocity Coefficient


Value of α
Channel Types Min. Avg. Max.

Regular Channels, Flumes, and Spillways 1.1 1.15 1.2

Natural Streams 1.15 1.30 1.5

River Valleys, Overflooded 1.5 1.75 2.0


Source: Chow, V.T., 1959, Open-Channel Hydraulics: New York, McGraw-Hill.

Additional guidelines on velocity coefficients can be found in the Techniques of Water-


Resource Investigations (TWRI) Reports of the United States Geological Survey.

In general, the more subdivisions in a cross section, the higher the alpha (α) value.

The energy equation for flow along a channel includes a term for the kinetic energy or
velocity head, V2/2g. Use the average velocity, V, for the entire cross section in the
equation. In reality the velocity is not a constant value. It is highest in the middle of the
channel near the water surface and lowest at the edges of the channel near the channel
bottom. Using the average velocity in the equation means that the sum of the differing
velocities in the cross section is being squared, (v1 + v2 + … + vn)2. However, to correctly
determine the kinetic energy, you should first square the differing velocities and then sum
them, v12 + v22 + … + vn2. Since the sum of the squares is greater than the square of the
sum, you will need to use the kinetic energy correction factor. This factor usually is
represented by the Greek letter alpha in the energy equation, and is, therefore, referred
to as alpha for short.

Alpha values are calculated and reported for each cross section in both HEC-RAS and
WSPRO. However, neither program provides warnings when alpha values are out of

Chapter 5: Bridge Hydraulics 5-51


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

range. Incorrect alpha values can cause significant errors. Check the alpha values to be
sure they are appropriate.

Alpha values typically should stay in the ranges shown in Table 5.2-1. In general, the
more subdivisions in a cross section, the larger alpha will become. Alpha values greater
than 3 should be checked. If adjacent cross sections have comparable values, or if the
changes are not sudden between cross sections, such values can be accepted. But if the
change is sudden, some attempt should be made to obtain uniformity. Consider the
following:

a. Resubdivide the cross section(s).


b. Place additional cross sections to provide a smoother transition of the alpha
values from one cross section to the next. Note that if the bridge routine in
WSPRO is used, additional cross sections cannot be placed between the exit
and approach sections.

Additional guidance is provided in the Techniques of Water-Resource Investigations


(TWRI) Reports.

The following examples illustrate the importance of proper subdivision, as well as the
effects of improper subdivision.

Example 5.2-2—Effects of Subdivision on a Panhandle Section


In Figure 5.2-8, the section given has a constant n value for the entire cross section. The
four calculations shown represent four methods of calculating total flow (conveyance),
depending on how the cross section is subdivided.

Figure 5.2-8: Effects of Subdivision on a Panhandle Section

Chapter 5: Bridge Hydraulics 5-52


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Given:

K = 1.49/n (AR2/3)
n is constant over cross section
Factor out 1.49 and compare AR2/3 = K feet

Note: K feet varies as to the number of sections selected as a function of R, or more


specifically Wp.

Method 1: Consider K1 feet as one section encompassing subareas “A,” “B,” and “C.”

2/3
2/3  (6x10) + (50x0.2) + (50x0.15) 
K 1 '= AR ; K 1 ' = [(6x10) + (50x0.2) + (50x0.15)]   = 57.3
 (0.1  50  50  5.8  10  6) 

Method 2: Consider K2 feet as two sections, “A” and “B” combined and “C.”

2/3
 2/3  (50x0.2) + (50x0.15) 
K 2 = [(6x10) (60/21.8) ] + [(50x0.2) + (50x0.15)]   = 117.8 + 5.5 = 123.3
 100.1

Method 3: Consider K3 feet as section “C” and ignore sections “A” and “B.”

2/3
  60 
K 3  (6  10)   117.8
 5.8  10  6 

Method 4: Consider K4 feet with “A,” “B,” and “C” treated as independent sections.


K 4  [(6  10)(60 / 21.8) 2 / 3 ]  [(50  0.2)(10 / 50) 2 / 3 ]  [(50  0.15)(7.5 / 50.1) 2 / 3 ]


K 4  117.8  3.4  2.1  123.3

Method 1 is incorrect. The problem is the method neglects the impact the hydraulic radii
of the shallow areas have on the overall flow calculation. This can be seen by looking at
method 3, which shows conveyance in just the main channel as being greater. Two
reasons why method 1 is incorrect are:

Chapter 5: Bridge Hydraulics 5-53


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

1. The total conveyance must be the sum of the conveyance of a channel’s


subsections.
2. Combining significantly different geometric sections of a cross section to simplify
a calculation is a misuse of the conveyance equation and will yield an incorrect
answer.

Method 2 is correct. It combines subareas of the channel cross section that have similar
hydraulic properties to yield a reasonable answer of total conveyance. If n values between
section “A” and “B” were significantly different, combining them to determine conveyance
might not provide the desired accuracy.

Method 3 is incorrect but exemplifies how easily you can underestimate total conveyance
by not considering the conveyance from the other subareas. Obviously, the total
conveyance cannot be less than that contained in one section.

Method 4 is correct. This may be considered overkill, but technically it is the most accurate
solution. If n values were significantly different between section “A” and “B,” this type of
subdivision for determining conveyance would be essential.

Example 5.2-3 Effects of Subdivision on a Trapezoidal Section


In Figure 5.2-9, a trapezoidal cross section having heavy brush and trees on the banks
has been subdivided near the bottom of each bank because of the abrupt change of
roughness there. A large percentage of the wetted perimeters (P) of the triangular
subareas (A1 and A3) and of the main channel (A2) have been eliminated. A smaller wetted
perimeter abnormally increases the hydraulic radius (R = A/P), and this, in turn, results in
a computed conveyance different from the conveyance determined for a section with a
complete wetted perimeter. In Figure 5.2-9, the total conveyance (KT) has been computed
to be 102,000 for the cross section. This would require a composite n value of 0.034. This
is less than the n values of 0.035 and 0.10 that describe the trapezoidal shape. The basic
shape should be left unsubdivided, and an effective value of n somewhat higher than
0.035 should be assigned to this cross section, to account for the additional drag imposed
by the larger roughness on the banks.

Chapter 5: Bridge Hydraulics 5-54


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.2-9: Effects of Subdivision on a Trapezoidal Section

Chapter 5: Bridge Hydraulics 5-55


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

(C) Friction Losses


Compute the friction loss as follows:

hf = L Sf

where:
L= Flow length, in feet
Sf = Average friction slope, in feet/feet

You can calculate the average friction slope using either the geometric mean slope
method, the average conveyance method, the average friction slope method, or the
harmonic mean friction slope method. WSPRO uses the geometric mean slope method
as the default option. The geometric mean slope is computed as:

2
[0.5 (Q1 + Q2)]
Sf =
K1 K 2

where:
Sf = Average friction slope, in feet/feet
Q1 = Discharge at Section 1, in cubic feet per second
Q2 = Discharge at Section 2, in cubic feet per second
K1 = Conveyance at Section 1
K2 = Conveyance at Section 2

(D) Expansion/Contraction Losses


Expansion Losses
Compute the expansion loss as follows:

he = k e (hv 2 - hv1)

where:
kc = Expansion loss coefficient
hV1 = Velocity Head in Section 1, in feet
hV2 = Velocity Head in Section 2, in feet

The expansion loss coefficient varies from 0.0 to 1.0 from ideal transitions to abrupt
transitions. HEC-RAS uses an expansion value of 0.3 as its default. WSPRO uses an
expansion value of 0.5 as its default. Brater and King’s Handbook of Hydraulics provides
additional guidance for selection of expansion coefficients.

Chapter 5: Bridge Hydraulics 5-56


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Contraction Losses
Compute the contraction loss as follows:

hc = k c (hv 2 - hv1)

where:
kc = Contraction loss coefficient
hV1 = Velocity Head in Section 1, in feet
hV2 = Velocity Head in Section 2, in feet

The contraction loss coefficient varies from 0.0 to 0.5 from ideal transitions to abrupt
transitions. HEC-RAS uses a contraction value of 0.1 as its default. WSPRO uses a
contraction value of 0.0 as its default. Brater and King’s Handbook of Hydraulics provides
additional guidance for selection of contraction coefficients.

(E) Step Backwater Computations


HEC-RAS and WSPRO computational procedure employs the Standard Step Method
for profile computations. The procedure used is similar to that described by Chow. The
standard step method is based on the principle of conservation of energy, i.e., the total
energy head at an upstream section must equal the total energy head at the
downstream section plus any energy losses that occur between the two sections.

Energy Equation

Write the energy equation between two adjacent cross sections as follows:

h1 + hv1 = h2 + hv 2 + hf + he + hc

where:
h1 = Water surface elevation in Section 1, in feet
hV1 = Velocity head in Section 1, in feet
h2 = Water surface elevation in Section 2, in feet
hV2 = Velocity head in Section 2, in feet
hf = Friction loss between Sections 1 and 2, in feet
he = Expansion loss between Sections 1 and 2, in feet
hc = Contraction loss between Sections 1 and 2, in feet

It is not possible to find a direct solution of the above equation when either h1 or h2 is
unknown, since the associated velocity head and the energy loss terms also are then
unknown. Therefore, an iterative procedure must be used to determine the unknown

Chapter 5: Bridge Hydraulics 5-57


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

elevation. The WSPRO model computes the difference in total energy between two
sections, H, as:

H = (h1 + hv1) - (h2 + hv 2 + hf + he + hc )

Use successive estimates of unknown elevations to compute the unknown velocity head
and the energy loss terms until the equation yields an absolute value of ∆H that is within
an acceptable tolerance. Generally, a tolerance between 0.01 and 0.05 is sufficient to
obtain satisfactory results. Slightly higher results may be satisfactory for some higher-
velocity situations. However, if a tolerance value exceeding 0.1 is required to obtain a
satisfactory solution, then there would be reason to suspect data inadequacies (example:
insufficient cross sections).

Computational Procedure
Given: Discharge Q and WSE at one cross section and the fact that the flow is
subcritical. We want to compute the WSE at the next upstream cross section.

Step 1: Calculate all the geometric and hydraulic properties of the downstream most
station using the known flows and WSE at that location.

Step 2: Estimate water surface elevation at the next upstream station.

Step 3: Calculate hydraulic properties that correspond to estimated water surface


elevation.

Step 4: Determine energy losses that correspond to estimated water surface elevation.

Step 5: Calculate water surface elevation using energy equation and energy losses
determined in Step 4.

Step 6: Compare estimated and computed water surface elevations.

Step 7: If the computed and estimated elevations do not agree within some
predetermined limit of error, try another value and start the procedure again
beginning with Step 2.

Example 5.2-4 illustrates a step backwater computation. Descriptions of conveyance,


velocity head, friction loss computations, and expansion and contraction losses are
provided after the example.

Chapter 5: Bridge Hydraulics 5-58


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Example 5.2-4: Standard Step Backwater Computation

Chapter 5: Bridge Hydraulics 5-59


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Example 5.2-4: Standard Step Backwater Computation (continued)

Chapter 5: Bridge Hydraulics 5-60


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.2.4.3.2 Two-Dimensional Models


Recommendations for developing model geometry for two-dimensional models will
depend upon the model employed. Two-dimensional models employ either finite element
or finite difference computation schemes. Finite difference models represent the model
domain with a regular grid of ground elevations. Figure 5.2-10: 5.2-10 displays examples
of the different types of grids employed in finite difference modeling. Finite element
methods represent the model domain with a network of triangular and quadrilateral
elements that can vary widely in both size and orientation. Figure and Figure 5.2-12
display examples of finite difference and finite element model meshes.

a) b) c)
Figure 5.2-10: Example of (a) Cartesian, (b) Rectilinear, and (c) Curvilinear Grids

Chapter 5: Bridge Hydraulics 5-61


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.2-11: Example of a Finite Difference Model Mesh


After defining the model domain, the next step in the model geometry development is to
specify element locations, sizes, and orientation. In other words, specify the resolution of
the model. Finite element models typically will incorporate increased resolution at the
project location, along bathymetric features that influence flow through the waterway
(shoals, point bars, etc.), and around physical structures in the flow field (causeways,
embankments, weirs, etc.) and less resolution with increased distance from the location
of interest. Additionally, higher resolution often is incorporated in areas of rapidly
changing bathymetry or topography. Examples include at channel banks, head cuts, drop
structures, seawalls, and bridge abutments. This varying resolution allows for optimization
of computation speed. An example of varying resolution is illustrated in Figure with the
increased resolution at the inlet and along the navigation channel and decreased
resolution in the deeper areas offshore. Mesh generation typically takes place via a
Graphical User Interface (GUI). One example is SMS (Surface Water Modeling System),
available through Aquaveo, which provides a number of mesh generation and editing
tools as well as pre- and post-processors for a wide variety of hydraulic and wave models.
Model resolution oftentimes is one of the model parameters that is modified to achieve
both model stability and model calibration.

Chapter 5: Bridge Hydraulics 5-62


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.2-12: Example of a Finite Element Model Mesh

Resolution specification for finite difference models is more challenging than with finite
element models. For models that can employ curvilinear or rectilinear grids, resolution
can be increased in a few select locations. By nature of the grids, however, this resolution
propagates in both ordinal directions from the area of interest through the remainder of
the grid. For Cartesian grids, the resolution of a grid is uniform throughout the domain.
Thus, the resolution at the bridge location will dictate the resolution for the remaining
domain. For large domains requiring fine resolution at the bridge location, a common
technique is to employ a nested grid scheme.

After specifying the model resolution, the final step in preparing the model geometry
involves specifying the elevations at the model element nodes. Again, this is typically
performed with automated mesh generation programs that interpolate a survey data set
onto the prepared grid or mesh. This step can sometimes lead to interpolation errors
depending upon the relative resolution of the survey data and the model grid/mesh as
well as the quality of the TIN (triangular irregular network) representing the survey data.

Chapter 5: Bridge Hydraulics 5-63


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Careful examination of how well the grid/mesh represents the elevations of the model
domain is an important part of the model calibration process.

5.2.4.4 Boundary Conditions


Upstream Flow
For a riverine analysis, give the flow at the upstream boundary. For a steady-state
analysis, specify the peak discharge for each frequency at the upstream boundary. For
an unsteady flow analysis, specify a flow hydrograph at the upstream boundary.

Downstream Stage
Specify the stage at the downstream cross section. Known water surface elevations are
the first choice. These can be lake levels, sea levels, or control sections such as a gage,
studies (e.g., FEMA), or critical depth sections.

You can use normal depth in many cases when the stream channel is nearly uniform for
a fairly long reach. You can use HEC-RAS or WSPRO to compute the normal depth by
providing an energy slope equal to the channel slope. This method also is known as
“slope conveyance.” Determine the channel slope using a USGS Quadrangle Map.
Determine the slope below the last downstream cross section where contour lines cross
the stream channel. You can use other estimates of energy slope; however, the resulting
water surface elevation would not be “normal depth.”

When there is no gage information available and when normal depth flow (slope
conveyance) cannot be assumed at the bridge site, you should use “convergence.”

Convergence
Water surface profiles will converge to a single profile if given enough distance to
converge. The distance depends on the channel and overbank properties and the slope
of the river. Estimate the distance as the downstream study length described in Section
5.2.4.1.

Determine convergence as follows:

a. Make trial-and-error calculations assuming a range of water surface elevations.


This assumed range of water surface elevations should bracket your best guess
of the water surface elevation at the farthest downstream cross section. Typically,
this is done using an estimate of the friction slope and calculating normal depth.
b. Using the estimate of water surface elevation at the farthest downstream cross
section, develop four water surface profiles for the design discharge based on a
range of potential water surface elevations. Two of the bracketed elevations should
represent the range between which the water surface should be, and the other two

Chapter 5: Bridge Hydraulics 5-64


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

should represent the range outside of which the water surface is unlikely to be.
Refer to Figure 5.2-13.
c. The computed profiles will converge toward the true profile. The profiles should
converge within an acceptable tolerance by the first section of interest in the reach
(see Figure 5.2-13). If the profiles do not adequately converge, then you should
obtain additional geometric data downstream.

Figure 5.2-13: Convergence Profiles

5.2.4.5 Bridge Model


Flow Characteristics at Bridges
Figure 5.2-14 illustrates the manner in which flow contracts in passing through the
channel constriction. The flow bounded by each adjacent pair of streamlines is the same
(1,000 cubic feet per second). Note that the channel constriction appears to produce
practically no alteration in the shape of the streamlines near the center of the channel. A
very marked change occurs near the abutments, however, since the momentum of the
flow from both sides (or floodplains) must force the advancing central portion of the stream
over to gain entry to the constriction. Upon leaving the constriction, the flow gradually
expands (5 to 6 degrees per side) until normal conditions in the stream are re-established.

Constriction of the flow causes a loss of energy, with the greater portion occurring in the
re-expansion downstream. This loss of energy is reflected in a rise in the water surface
and in the energy line upstream of the bridge. This is best illustrated by a profile along the
center of the stream, as shown in Figure 5.2-15 (Part A). The dashed line labeled "normal
water surface" represents the normal stage of the stream for a given discharge before

Chapter 5: Bridge Hydraulics 5-65


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

constricting the channel. The solid line labeled "actual water surface" represents the
nature of the water surface after constriction of the channel. Note that the water surface
starts out above normal stage at Section 1, passes through normal stage close to Section
2, reaches minimum depth in the vicinity of Section 3, and then returns to normal stage a
considerable distance downstream, at Section 4. Determination of the rise in water
surface at Section 1 is denoted by the symbol h1* and referred to as the bridge backwater.

Chapter 5: Bridge Hydraulics 5-66


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Reference: USDOT, FHWA HDS-1 (1978)


Figure 5.2-14: Flow Lines for Typical Bridge Crossing

Chapter 5: Bridge Hydraulics 5-67


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.2-15: Normal Crossings: Spill-through Abutments

Chapter 5: Bridge Hydraulics 5-68


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Roughness
The roughness around and under the bridge can be significantly different than the
roughness upstream and downstream due to rubble riprap protection and clearing of trees
and underbrush. The main channel roughness often is the same through the bridge from
upstream to downstream. The most common reason that the roughness will change is if
there is a significant extent of rubble riprap protecting the piers or channel banks.

Many Florida floodplains are heavily vegetated. Many riverine bridges span a significant
length across the floodplain. The area beneath the bridge often is cleared of the trees and
underbrush, and is maintained that way. This will reduce the roughness. However, rubble
protection of the abutment will increase the roughness. The guidelines for subdivision
(refer to Section 5.2.4.3) usually would recommend against subdividing at the toe of the
abutment, so a weighted roughness should be determined.

Be careful to model abrupt changes in roughness appropriately to properly account for


the friction loss between the cross sections. The Standard Step Method uses an average
of the conveyance for each cross section to calculate the friction loss between the cross
sections, which essentially averages the roughness values of the two sections. A good
method of modeling abrupt roughness changes is to include two cross sections closely
spaced at the change location. However, some of the bridge routines of the various
models will not allow the extra cross section.

Nodes and elements in two-dimensional models can be placed such that abrupt
roughness changes do not bisect elements.

Bridge Routine
Refer to HEC-RAS documentation for cross section location information. However, if you
are using the WSPRO bridge routine when modeling in HEC-RAS, don’t follow the
documentation; instead, use the following recommendations.

The bridge routine in WSPRO uses the Standard Step Backwater Method, only with more
complexity. The bridge hydraulics is based on the reach from the exit section to the
approach section, as defined in the WSPRO Manual. Although the manual specifies "one
bridge length,” this does not mean the exit section must be exactly one bridge length
downstream from the full-valley section or that the approach section must be exactly one
bridge length (plus roadway width) upstream from the Full-Valley section. The locations
of these sections can vary as follows.

Exit Section:
The exit section can be located no less than, but as much as 10 percent greater than,
one bridge length from the full-valley section. See Figure 5.2-16.

Chapter 5: Bridge Hydraulics 5-69


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.2-16: Location of Exit Section

Approach Section:
The approach section can be located as much as 15 percent less than or greater than
one bridge length plus the roadway width from the upstream face of the bridge. See Figure
5.2-17.

Figure 5.2-17: Location of Approach Section

If, for some reason, it is impossible to follow the cross section requirements, you may
need to analyze the site without using the bridge routine.

Chapter 5: Bridge Hydraulics 5-70


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Piers
You can model single-row pile-bent bridges without modeling the piles and the hydraulic
results will be the same as if they were included. However, regulatory agencies may want
to see the piles included in the model. As the blockage becomes greater for more complex
piers, the hydraulic results will change.

5.2.5 Simulations

5.2.5.1 Calibration
Calibration involves changing the value of coefficients until the model results match
observed field conditions for one or more known events. When the model has been
calibrated to known events, then you can model an unknown event, such as the design
frequency event, with more confidence.

Observed field data for a flood event can include:

 Water surface elevations


 Discharge measurements
 Velocity measurements
Obtain data from multiple flood events, if available. The closer the magnitudes of the
observed events are to the magnitude of the design events, the more certain the results
will be.

Generally, the most reliable source of information is gage data. Most gages used in
riverine situations measure the water surface elevation. Figure 5.2-18 shows a simple
staff gage that you must observe and record manually. More complex gaging stations will
record stages automatically and either store the records for later download or transmit the
data using telemetry.

Chapter 5: Bridge Hydraulics 5-71


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.2-18: Staff Gage on the Suwannee River

You can determine discharges indirectly from the water surface elevations. Traditionally,
you would use a velocity meter to take measurements at intervals across the stream and
then determine the discharge, as shown in Figure 5.2-19. When you have determined the
discharge at enough different water surface elevations, you can establish a stage versus
discharge relationship for the gage.

Chapter 5: Bridge Hydraulics 5-72


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.2-19: Discharge Determination with a Velocity Meter


(from USGS Streamgaging Fact Sheet 2005-3131, March 2007)

More recently, discharges have been measured on some larger rivers with an Acoustic
Doppler Current Profiler mounted on a boat (see Figure 5.2-20).

Figure 5.2-20: Discharge Determination with an Acoustic Doppler Current Profiler


(from USGS Streamgaging Fact Sheet 2005-3131, March 2007)

The primary benefit of a gage is to establish the discharge for an observed flood. If a gage
is located within the model reach, then the gage also can supply stage and velocity
information at one point in the model.

Chapter 5: Bridge Hydraulics 5-73


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

If gage data are unavailable, consider sending survey out to measure:

 High water marks associated with known floods (Figure 5.2-21)


 Local resident or official high water permanent markers/signs (Figure 5.2-22)
 Ordinary high water marks (stain lines on existing bridge pilings or vegetative
indicators)

Occasionally, the Department and agencies such as USGS, FEMA, DEM, or the Water
Management Districts may have surveyed or collected high water marks following a flood.
Contacting them is an avenue to pursue.

Figure 5.2-21: Examples of High Water Marks after a Flood

If a gage is not available to determine the discharge of the known event, then estimating
the discharge associated with the various high water marks will be difficult or impossible.
Obtaining rain gage information for the flood and estimating the runoff from the rainfall is
an option, assuming data from a suitable rain gage are available. Otherwise, the high
water marks can only be compared to the computed design frequency profiles from the
model to check the magnitudes for reasonableness.

Chapter 5: Bridge Hydraulics 5-74


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.2-22: Local Resident indicating Flood Level on the Caloosahatchee River near
LaBelle in 1913

After you obtain available gage data and/or high water mark elevations, the next step is
to develop the hydraulic model for the existing site conditions. In some situations, this
might entail creating multiple existing-condition models if the site conditions have
changed since some of the calibration floods. Develop the model using standard guidance
for the coefficients used in the model. Then compare the initial model results to the high
water marks and adjust the coefficients. The common coefficients to adjust are:

 Manning’s Roughness Coefficient


 Bridge loss coefficients (depending on the bridge routine used)
 Expansion and Contraction Coefficients

Manning’s roughness coefficient is the basic adjustment tool for unobstructed reaches.
Considerable uncertainty exists when estimating roughness values. Estimates by
experienced hydraulics engineers often vary by ± 20 percent (from USACE EM 1110-2-
1416). If you hold the channel roughness constant and vary the overbank roughness, you
should be well served.

Chapter 5: Bridge Hydraulics 5-75


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Also, remember that Manning’s roughness varies with depth, which can affect calibration
as follows:

 As the depth over the roughness elements increases, n decreases.


 If the flow encounters a new roughness element as the flow depth increases, n
will increase. For example, if tree branches are higher than a certain depth in the
floodplain, the roughness will increase when the flow reaches the tree branches.

Do not adjust the calibration coefficients outside of their normal ranges. If the calibration
attempts are not acceptable, re-examine the model. Common model parameters to
review if calibration is a problem include:

 Ineffective flow areas


 Starting conditions downstream
 Cross section locations
 Cross section subdivisions
 Accuracy of survey data or other geometric data
 Datums of geometric data
 Flow lengths
 Warning messages

Note that calibration problems can be caused by different issues. Use your best judgment
in the calibration process. There is no universally accepted procedure or criterion for
calibration.

Calibrating unsteady flow models is more difficult than calibrating steady flow models.

Adjust to steady flow conditions first, if possible. Unsteady flow models need to be
calibrated over a wider range of flows than steady-state models. Storage in the system is
an important parameter for unsteady flow, and essentially can be used as an adjustment
parameter. For more detail on techniques for unsteady flow calibration, refer to USACE
Manual EM 1110-2-1416, River Hydraulics.

Two-dimensional models have eddy viscosity, or turbulent loss coefficient, that becomes
another calibration parameter. This term in essence replaces expansion and contraction
losses in a one-dimensional model. However, there is not an established correlation
between the two losses. The best way to calibrate eddy viscosity is with measured
velocities. Remember that the two-dimensional velocity is depth-averaged, so you must

Chapter 5: Bridge Hydraulics 5-76


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

convert the measured velocity to a depth-averaged velocity for comparison. Set the value
high first, and then lower it until you obtain the ideal velocity distribution. The general
order of calibration for two-dimensional models would be to calibrate roughness values
to observed water surface elevations, and then adjust eddy viscosity to observed
velocities.

When using both velocities and stages for calibration, check for internal consistency of
the observed data. The velocity times the area for the stage should be approximately
equal to the discharge.

5.2.5.2 Existing Conditions


Model the existing conditions to compare with the results from the proposed structure and
to calibrate the model to observed flood data. If the existing condition has a bridge at the
site, then consider also modeling the natural conditions at the site prior to construction of
the existing bridge.

5.2.5.3 Design Considerations


Review Project Development and Environment (PD&E) documents for commitments
made during the NEPA process. During PD&E, a Location Hydraulics Study should look
at alternate locations for the plan view of the roadway crossing of the stream or river.
Identify adverse hydraulic conditions in the Location Hydraulics Study for consideration
when planning the roadway crossing. The final location will not depend solely on hydraulic
aspects, but consider them during the initial planning of the roadway. The location and
alignment of the highway can either magnify or eliminate hydraulic problems at the
crossing. By the time the Bridge Hydraulics Report is prepared, the location and alignment
of the road should be set; however, minor changes to the alignment still may be possible.

Usually, you will evaluate and select the length of the bridge and the location of the
abutments in the Bridge Hydraulics Report. Traditionally, at least three lengths are
analyzed. One is the minimum hydraulic structure, the bridge that creates no more than
one foot of backwater and does not violate other allowable water surface conditions.
Another bridge length examined is the bridge that spans all wetlands. Other potential
bridge lengths to investigate include:

 The length of the existing bridge


 For dual bridges, the length of the existing dual bridge that will be left in place
 Breaks in fill height if bridging is less expensive than roadway fill
 Minimum bridge length based on setbacks from the channel banks

Chapter 5: Bridge Hydraulics 5-77


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Other considerations when designing and modeling the proposed conditions are:

 Place the bridge in a crest vertical curve, if possible. Allowing the approach
roadways to overtop more frequently than the bridge will provide relief for the
bridge, and reduce the possibility of damage to the structure. If a portion of the
roadway is damaged, it usually can be repaired more easily than the bridge.
 Try to center the bridge over the main channel of the flow. At a minimum, set the
toe of the abutments 10 feet back from the top of the channel bank.
 Consider skewing the abutments and intermediate bents to align with the flood flow
direction to reduce scour potential.

5.3 TIDAL ANALYSIS


A qualified coastal engineer should perform hydraulic and scour analyses of tidal and
tidally influenced bridges. Section 5.1.1 defines the requirements and credentials of
coastal engineers qualified to perform tidal analyses for the Department.

5.3.1 Data Requirements


Evaluation and design of tidally influenced bridges requires a preliminary, systematic data
collection effort to determine the hydraulic conditions at the structure, calculate the scour,
and develop the wave climate at the structure. This information includes details of the
bridge geometry, the bed composition and elevations, and historical measurements and
studies.

5.3.1.1 Survey Data


You will need survey data to perform several aspects of a bridge hydraulics and scour
analysis. Survey data not only provide the elevation data to construct hydraulic and wave
models, but also provide needed sediment characteristics for scour calculations. The
requirements for a tidal analysis are the same as those for riverine analyses with one
exception: typically, the size of the modeling domain for tidal studies is substantially larger
than for riverine studies. Since new survey acquisition of the required data over the entire
domain is rarely cost-effective, you can supplement survey data acquired around the
bridge with publicly available data. Several sources exist for supplemental data, including
the following examples:

 Bathymetric and topographic data from the National Geophysical Data Center
(http://www.ngdc.noaa.gov/mgg/bathymetry/relief.html, Example: Figure 5.3-1
 Digital Elevation Models from the FDEP Land Boundary Information System
website (http://www.labins.org/mapping_data/dem/dem.cfm)

Chapter 5: Bridge Hydraulics 5-78


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

 Coastal LiDAR data from NOAA’s Coastal Services Center


(http://coast.noaa.gov/digitalcoast/data/coastallidar)

Be careful when combining data from several sources. There can be wide ranges in
accuracy due to differing measurement techniques and survey dates. Pay close attention
to conversion between different horizontal and vertical coordinate systems. Examine
boundaries between survey data sets for inconsistencies and corrections.

The accuracy and density of survey data become more important near the site of interest.
This is especially true of bathymetry for wave modeling when you expect depth limitation
to govern wave conditions.

Figure 5.3-1: NOAA National Geophysical Data Center Website

Chapter 5: Bridge Hydraulics 5-79


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.3.1.2 Geotechnical Data


To calculate scour at bridge foundations, you will need geotechnical information to
establish the bed composition and its resistance to scour. Data requirements for tidal
bridges are the same as those for riverine bridges. Refer to Section 5.2.1.2 for a
discussion of geotechnical data requirements.

5.3.1.3 Historical Information


Historical information provides data for calibration through gage measurements and
historical high water marks, data for calculation of long-term scour processes through
historical aerial photography and Bridge Inspection Reports, and data for characterization
of the hurricane vulnerability through the hurricane history.

Tidal Bench Marks


Tidal datums are vertical elevations that describe the tidal fluctuation at a particular
location. Several tidal datums are in common use, including mean high water (MHW),
which is the base elevation for structure heights, bridge clearances, etc., and mean low
water (MLW), which is the officially designated navigational chart datum for the United
States and its territories. To be accessible when needed, these datums are referenced to
fixed points known as bench marks. NOAA maintains numerous tidal bench marks
throughout the state of Florida that are available from the Center for Operational
Oceanographic Products and Services (CO-OPS) website
(http://www.co-ops.nos.noaa.gov/). The Florida Department of Environmental Protection
(FDEP) is an additional source for this information. The FDEP website LABINS (Land
Boundary Information System) contains a water boundary data map interface that lists
not only the MLW and MHW at the NOAA bench mark locations, but also these datums
at interpolated locations along interior tidal waterways. The LABINS website information
(http://www.labins.org/survey_data/water/water.cfm) is recommended for locations
where NOAA tidal bench marks are either unavailable or display a wide range of vertical
variation around the project location.

Several other tidal datums are available, and you should document them for each tidally
controlled or influenced project.

The east coast of Florida experiences semi-diurnal tides and the panhandle experiences
diurnal tides. The coastline from the tip of the peninsula to Apalachicola experiences
mixed tides—tides characterized by a conspicuous diurnal inequality in the higher high
and lower high waters and/or higher low and lower low waters. Figure 5.3-2 and Table
5.3-1 display an example of tidal bench mark information and gage data (with tidal
datums) for Key West, Florida.

Chapter 5: Bridge Hydraulics 5-80


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Table 5.3-1: Elevations of Tidal Datums in ft-NAVD88 for NOAA Tidal Bench Mark
#8724580 (Key West, FL) for the 1983-2001 Tidal Epoch

MEAN HIGHER HIGH WATER (MHHW) +0.05


MEAN HIGH WATER (MHW) -0.24
MEAN TIDE LEVEL (MTL) -0.88
MEAN SEA LEVEL (MSL) -0.87
MEAN LOW WATER (MLW) -1.52
MEAN LOWER LOW WATER (MLLW) -1.76

1
Spring Tides

0.5

0 MHHW
MHW
Elevation (ft-NAVD)

-0.5
MSL
-1 MTL

-1.5 MLW
MLLW
-2
Neap Tides
-2.5

-3
6/26 7/16 8/5 8/25 9/14 10/4
Date

Figure 5.3-2: Measured Tides at Key West and Tidal Datums

Chapter 5: Bridge Hydraulics 5-81


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Gage Measurements
Gage measurements provide information both for model calibration and model boundary
conditions. Several sources of gage data are available to the public. The types of gage
measurements typically employed in tidal analyses include:

 Streamflow and river stage gages—for establishing inland boundary conditions


and calibration
 Tide gages—for oceanward boundary conditions and calibration of tidal circulation
 Wave gages—for calibration of wave models

Data sources of streamflow and river stage records are the same as those discussed for
riverine analyses.

You also can employ tide gage data for development of model boundary conditions, as
well as for model calibration. Tide gages record stages at a fixed location in tidally
influenced areas. NOAA maintains gages throughout the state. You can access both
recent and historic data online at http://www.co-ops.nos.noaa.gov/. In Florida, the site
provides data at 29 active stations (Figure 5.3-3) and historic data at 722 locations.

Chapter 5: Bridge Hydraulics 5-82


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.3-3: Location of Florida’s Active Tide Stations Maintained by NOAA


(Source: http://www.co-ops.nos.noaa.gov/gmap3/)

Used to calibrate data for wave models, wave gage data typically is much more rare than
either streamflow or stage records. The National Data Buoy Center (NDBC)—a part of
the National Weather Service (NWS)—designs, develops, operates, and maintains a
network of data-collecting buoys and coastal stations. Several of these stations measure
wave parameters, including significant wave height, swell height, swell period, wind wave
height, wind wave period, swell wave direction, wind wave direction, wave steepness,
and average wave period. The NDBC website (http://www.ndbc.noaa.gov/) provides both

Chapter 5: Bridge Hydraulics 5-83


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

recent and historical observations at several locations around Florida (Figure 5.3-4).
Figure 5.3-5 provides an example of these data as time series of significant wave heights.
Sources of wave gage data for interior waters (such as bays, estuaries, intracoastal
waterways, etc.) are much harder to locate. Possible sources may include previous
studies and academic institutions.

Figure 5.3-4: Locations of NDBC Stations around Florida


(Source: http://www.ndbc.noaa.gov/)

Chapter 5: Bridge Hydraulics 5-84


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Significant Wave Height (m) 9


8
7
6
5
4
3
2
1
0
8/16/2005 8/26/2005 9/5/2005
Date

10
Significant Wave Height (m)

0
8/16/2005 9/5/2005
Date

Figure 5.3-5: Example of Wave Gage Data at NDBC Station 42039 during Passage
of Hurricane Katrina

Chapter 5: Bridge Hydraulics 5-85


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Historical High Water Marks


The historical hurricane high water marks provide additional calibration data sets for the
storm surge numerical model during specific hurricane events. FEMA typically performs
post-storm damage assessments. Although the survey accuracy has significantly
increased in recent years, be cautious when using these data. Coastal high water marks
typically are designated as one of three basic types:

 Surge—represents the rise in the normal water level


 Wave height—represents the coastal high water mark elevation due to more direct
wave action
 Wave run-up—represents the height of water rise above the stillwater level due to
water running up from a breaking wave

You often can find high water marks near each other and they can vary widely in elevation.
Surge-only high water marks occur only where the structure is at a location sheltered from
waves. As waves propagate inland during a surge, the high water conditions on structures
and land can vary widely. When the crest of the wave rides on the surge, this creates
coastal wave height flooding. Thus, differences will occur between high water marks
measured on the interior and exterior walls of a structure. Finally, wave run-up high water
marks include the effects of waves breaking on sloping surfaces. After a wave breaks on
a beach or sloping surface, a portion of the remaining energy will propel a bore that will
run up the face of the slope. The vertical distance the bore travels above the still water
level is termed the wave run-up. Wave run-up often pushes debris to its maximum limit,
where it is left as a wrack line (a line of debris illustrating the extent of the wave run-up).

Hurricane History
The hurricane history of the project location characterizes the hurricane frequency at the
project, as well as the historical impacts to the site location. Including this information in
the Bridge Hydraulics Report elevates the importance of examining hurricane surge and
wave impacts, providing a qualitative examination of the frequency of hurricane influences
at the bridge site. Additionally, it can provide a tool for comparing the selected calibration
hurricane to the overall activity for the area. The BHR should include the historical
hurricane paths, historical storm year, and category, as well as discussion of significant
storms to impact the area. An example of the hurricane paths and listing of the historical
hurricanes is displayed in Figure 5.3-6 and Table 5.3-2 (from
https://coast.noaa.gov/hurricanes/).

Chapter 5: Bridge Hydraulics 5-86


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.3-6: Hurricane and Tropical Storm Tracks Passing within 50 Nautical
Miles (nmi) of Miami (Source: NHC)

Chapter 5: Bridge Hydraulics 5-87


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Table 5.3-2: Hurricanes Passing within 50 nmi of Miami


Wind Speed Pressure
Year Month Day Storm Name Category
(kts) (mb)
1865 10 23 NOTNAMED 90 0 H2
1870 10 10 NOTNAMED 90 0 H2
1878 10 21 NOTNAMED 70 0 H1
1885 8 24 NOTNAMED 70 0 H1
1888 8 16 NOTNAMED 110 0 H3
1891 8 24 NOTNAMED 75 0 H1
1903 9 11 NOTNAMED 75 976 H1
1904 10 17 NOTNAMED 70 0 H1
1906 10 18 NOTNAMED 105 953 H3
1909 10 11 NOTNAMED 100 957 H3
1924 10 21 NOTNAMED 70 0 H1
1926 9 18 NOTNAMED 120 0 H4
1926 10 21 NOTNAMED 95 0 H2
1935 9 28 NOTNAMED 100 0 H3
1935 11 4 NOTNAMED 65 973 H1
1941 10 6 NOTNAMED 105 0 H3
1945 9 15 NOTNAMED 120 0 H4
1947 9 17 NOTNAMED 135 947 H4
1947 10 12 NOTNAMED 75 0 H1
1948 9 22 NOTNAMED 100 0 H3
1948 10 5 NOTNAMED 110 975 H3
1950 10 18 KING 95 0 H2
1964 8 27 CLEO 90 968 H2
1964 10 14 ISBELL 110 968 H3
1965 9 8 BETSY 110 952 H3
1966 10 4 INEZ 75 984 H1
1979 9 3 DAVID 85 973 H2
1987 10 12 FLOYD 65 993 H1
1992 8 24 ANDREW 130 937 H4
1999 10 16 IRENE 65 986 H1
2005 10 24 WILMA 110 953 H3

Chapter 5: Bridge Hydraulics 5-88


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Historical Aerial Photographs


Historical aerial photographs aid in evaluating the channel stability at a bridge crossing.
Comparison of photographs over a number of years can reveal long-term erosion or
accretion trends of the shorelines and channel near the bridge crossing. An example of
this is provided in Figure 5.3-7 and Figure 5.3-8. From the figures, changes in shoreline
location occur south of the east abutment as well as to the spit south of the inlet. Section
5.5.1.1 further discusses calculation of long-term trends. Sources of historical aerial
photography are the same as those discussed in Section 5.2.1.3.

Figure 5.3-7: Heckscher Drive (SR-A1A) near Ft. George Inlet in 1969

Chapter 5: Bridge Hydraulics 5-89


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.3-8: Heckscher Drive (SR-A1A) near Ft. George Inlet in 2000

Existing Bridge Inspection Reports


Existing Bridge Inspection Reports often provide sources of recent and historical cross
section measurements, as well as identify areas of hydraulic/scour-related damage or
repairs. Refer to Section 5.2.1.3 for additional discussion on obtaining and using these
reports in hydraulic analyses.

Wave Information Studies


Another source of coastal wave hindcast data is the Wave Information Studies (WIS),
developed and maintained by the U.S. Army Corps of Engineers (USACE) Coastal and
Hydraulic Laboratory. The WIS project produced an online database of hindcast,
nearshore wave conditions along the U.S. coasts. The hindcast data provide a source of
decades-long wave data that can provide boundary conditions or calibration data for
nearshore wave modeling. The data include hourly wave parameters of significant wave
height, peak period, mean period, mean wave direction, and wind speed and direction
(Figure 5.3-9). The database includes both nearshore and offshore gages along both
Florida’s Atlantic Ocean and Gulf of Mexico shorelines. The data are available via the
following link: http://wis.usace.army.mil/

Chapter 5: Bridge Hydraulics 5-90


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.3-9: Example of Available WIS Data from:


http://wis.usace.army.mil/hindcasts.html

Chapter 5: Bridge Hydraulics 5-91


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Previous Studies
Previously performed studies of a waterway can provide additional sources of data. Refer
to Section 5.2.1.3 for sources and discussion of previous studies.

5.3.1.4 FEMA Maps


FEMA Flood Insurance Rate Maps (FIRM) are the official maps of communities that
display the floodplains—specifically, special hazard areas and risk premium zones—as
delineated by FEMA. They are located at https://msc.fema.gov/portal. These maps
display areas that fall within the 100-year flood boundary. Information pertinent to bridge
hydraulics analysis includes whether the bridge resides in a FEMA floodway (see Section
5.1.2). Additionally, the map’s 100-year elevations can provide a check for modeling
results for the area. It is not unusual for the FEMA-listed elevations to differ significantly
from hurricane storm surge modeling results developed at an individual site. Many of
FEMA’s older coastal studies were performed via application of either the TTSURGE or
FEMA SURGE two-dimensional models, models driven with atmospheric (wind and
pressure) boundary conditions. A Joint Probability Method analysis of the model results
determined the return periods of surge elevations. The last time the FEMA SURGE model
was used in a new or updated flood insurance study to revise the FIRMs occurred in the
late 1980s. Thus, you can attribute deviation in 100-year flood elevations from the
published FEMA values to differences in the numerical models, boundary conditions,
inclusion of wave setup, as well as in the post-simulation analysis. More recently, FEMA
has begun to perform coastal restudies of locations throughout Florida, employing more
up-to-date modeling and statistical analyses. As the new maps become available, they
will replace older currently available maps.

5.3.1.5 Inland Controls


Data collection for inland controls follows the same recommendations as for the upstream
controls of riverine analysis (see Section 5.2.1.6).

5.3.1.6 Site Investigation


You should plan to do a field investigation for all new bridge construction. Refer to Section
5.2.1.7 for a detailed list outlining key items you should collect during site investigations.
In addition to this list, data collection at tidal bridges also should include the following:

 Look for evidence of wave scarping in bridge approaches.


 Note directions of largest fetches.
 Look for evidence of wave overtopping of seawalls and bulkheads.
 Note scattering of rubble riprap at toes of revetments, seawalls, and bulkheads
by waves.

Chapter 5: Bridge Hydraulics 5-92


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.3.2 Hydrology (Hurricane Rainfall)


During hurricane events associated with heavy rainfall, you can experience significant
surface runoff from land. For coastal areas, even though the storm surge is the larger
concern, surface runoff may increase or decrease the surge effects depending on the
phasing between the two (Douglass and Krolak, 2008).

The USACE reference, Engineering and Design Storm Surge Analysis EM 1110-2-1412
(1986), provides a methodology for estimating rainfall associated with landfalling
hurricanes. The methodology applies to the area within 25 miles of the coast. It provides
graphs of point rainfall depth for a given frequency and a given distance from the left or
right of the storm track. The rainfall varies uniformly along the coast for any given storm.
Also, the rainfall depths are uniform along any line parallel to the storm track extending
across the 25-mile-wide zone. The reference provides point rainfall graphs (Figure 5.3-
10) for selected frequency levels at either 6-hour or 12-hour intervals before landfall and
after landfall. The reference provides techniques for estimating rainfall associated with
hurricanes traveling at high, moderate, and slow speeds by multiplying the rainfall from
the graphs by a ratio coefficient that is a function of area.

Alternatively, as a rule of thumb, you may assume a steady 10-year discharge over the
duration of the surge. This is likely to be conservative in light of a recent examination of
hurricane rainfall in North Carolina that suggests that a two-year rainfall well represented
historical storms in that state (OEA, 2011). Bridges over streams with short times of
concentration (< four hours) are more likely to have coincidence between the storm surge
passage and high runoff values. Historical review of the timing and magnitude of runoff
at gaged locations near the project site can provide additional insight into the appropriate
return period flow rates for boundary conditions. At a minimum, you should perform a
sensitivity study to characterize the influence of the runoff magnitude on the flow
properties at a subject bridge during a surge event.

Chapter 5: Bridge Hydraulics 5-93


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.3-10: Rainfall for Selected Frequency Levels for Six Hours before
Landfall (Source: USACE 1986)

5.3.3 Model Selection


If you perform hydraulic studies, you must weigh several factors when selecting a
modeling approach, including:

 Types of models (e.g., one-dimensional vs. two-, or three-dimensional models;


finite-element vs. finite-difference models)
 Site conditions (e.g., embankment skew, multiple openings, etc.)
 Data availability (e.g., survey data, design flows/stages, etc.)
 Familiarity with the model
 Schedule and budget

Chapter 5: Bridge Hydraulics 5-94


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Weigh all the factors mentioned above and select the appropriate model for the
application. NCHRP Web-Only Document 106: Criteria for Selecting Hydraulic Models
(Gosselin et al., 2006) provides a decision analysis tool and guidelines for selecting the
most appropriate numerical model for analyzing bridge openings in riverine and tidal
systems. The decision tool takes the form of a decision matrix that incorporates all the
factors that influence model selection, including site conditions, design elements,
available resources, and project constraints. The utility of the decision tool is that it
presents a formal procedure to apply for the selection of the appropriate model rather
than an intuitive process.

Figure 5.3-11 presents an example where an engineer is selecting between one- and two-
dimensional models. The figure shows the scoring and weighting of different aspects of
the project, with the final selection of the one-dimensional model based largely on
advantages in scheduling. The selection procedure provides an easy-to-understand and
defensible method for presentation to non-technical readers or policy makers. Also,
through its application, it clearly identifies which features of the project are most important
in the model selection for a specific application.

Figure 5.3-11: Example of Model Selection Worksheet


from NCHRP Web-Only Document 106

Chapter 5: Bridge Hydraulics 5-95


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

For tidal analyses, in general, one-dimensional modeling works well for waterways with
well-defined channels in areas that are not subject to lateral overtopping. An example
would include rivers or canals that discharge directly to the open coast (e.g., Suwannee
River, Florida Barge Canal). More complex waterways and flow circulation will require
two-dimensional modeling. Examples requiring two-dimensional flow modeling include:

 Multiple interconnected channels


 Influence of multiple inlets
 Overtopping of barrier islands
 Bridges over tidal inlets
 Bridges over causeway islands
 Bridges through island chains

For wave models, there is not currently a similar selection procedure available. Selecting
the appropriate model is left to your experience and discretion after carefully weighing the
required design criteria and model features. Confirm your final model selection with the
District Drainage Engineer.

5.3.3.1 Storm Surge Model


Developing design hydraulic parameters at a bridge location requires the model to
simulate storm surge propagation from an open coast to the bridge site. This necessitates
application of an unsteady-state model. The following partial list includes several
commonly employed one-dimensional and two-dimensional models for simulating
hurricane storm surge:

 Advanced Circulation Model (ADCIRC) 2DDI


 TUFLOW
 DELFT3D
 FESWMS 2DH
 HEC-RAS 3.1.1 and up
 MIKE 11 HD v.2009 SP4
 MIKE 21 (HD/NHD)
 TABS RMA2
 UNET 4.0

Chapter 5: Bridge Hydraulics 5-96


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.3.3.2 Wave Model


You can use either numerical models or deterministic methods in developing design wave
climate parameters. The USACE references Coastal Engineering Manual (2002) and
Shore Protection Manual (1984) both provide empirical equations and methodologies for
calculating wave parameters over open water fetches. The following partial list includes
several commonly employed tools and models for simulating hurricane-generated waves:

 ACES
 MIKE 21 Flexible Mesh Spectral Wave Model
 MIKE 21 Nearshore Spectral Wave Model (NSW)
 RCPWAVE
 Simulating Waves Nearshore (SWAN)
 Steady-State Spectral Wave (STWAVE)

5.3.3.3 Model Coupling


Model coupling refers to the interaction between the wave model and the surge model
when simulating hurricanes. With no coupling, the surge and wave models run
independently. Since the wave model requires a water surface elevation for input, this
can lead to under-prediction if the surge is not taken into account. Figure 5.3-12, taken
from Sheppard et al., Design Hurricane Storm Surge Pilot Study, FDOT Contract No. BD
545 #42 (2006), displays wave simulation modeling of Hurricane Katrina at a location
offshore of Mississippi. In the figure, the “Without SS” curve is the wave height simulated
without the storm surge as an input boundary condition. The “With SS” curve includes
storm surge as an input into the wave model. Including storm surge produces a four-
meter increase in the predicted significant wave height.

Chapter 5: Bridge Hydraulics 5-97


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.3-12: Wave Height Simulation during Hurricane Katrina


with No Coupling (Without SS Curve) and
with One-Way Coupling (With SS Curve)
(Source: Sheppard et al., 2006)

With one-way coupling, input results (water elevations and currents) from the surge model
into the wave model. This leads to more accurate prediction of the wave climate. With
two-way coupling, transmit results from each model between the models at regular
intervals. The wave model receives the simulated surge elevations and currents as an
input, and the surge model receives the wave radiation stresses (a source term in the
momentum equations that gives rise to wave setup) as an input. In general, two-way
coupling provides the most accurate predictions.

5.3.4 Model Setup


Model setup involves development of the model inputs for the hydraulic or wave model.
It includes defining the model domain, assigning friction (roughness), creating the model
geometry, and developing boundary conditions.

5.3.4.1 Defining the Model Domain


The model domain is the spatial coverage of the model upstream of and oceanward of
the bridge. The limits of the model extents are different for storm surge modeling than for
riverine flood modeling. The model domain oceanward should extend to the point where

Chapter 5: Bridge Hydraulics 5-98


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

boundary conditions can be well described. For storm surge studies, this is generally the
open coast. Application of storm surge hydrograph boundary conditions, developed for
the open coast, at upland locations (e.g., at river entrances on estuaries or bays) will
result in overly conservative estimates of both surge elevation and flow rate at the bridge
location. If the model involves wind and pressure boundary conditions rather than a
hydrograph, the model should extend far enough offshore to accurately describe the
coastal effects (wind and wave setup) that contribute to the storm surge.

At a bridge, the accuracy of the surge hydrograph will be a function of the model resolution
between the open coast and the bridge location. Definition of the major tidal waterways
between the ocean and the bridge is recommended. Often, this includes extending the
model not only from the closest tidal inlet to the bridge, but also to nearby inlets as well.
This is particularly true for bridges located on or near intracoastal waterways.

Flow through the bridge is a function of the storage upstream (inland) of the bridge. The
model domain should extend far enough upstream and upland to accurately describe the
flow prism during the surge event. Underestimating the storage area upstream of a bridge
will result in underestimation of flow and scour at the site.

Definition of wave model extents will depend on the purpose of the wave model. If the
modeling results will provide wave radiation stresses for the surge model, then the wave
model should include similar offshore and lateral extents as the surge model as well as
the interior waters. If the purpose of the wave model is only to provide local wave
conditions at the site, then the model should extend from the bridge to the shoreline in all
directions so that the fetch (distance that the wind blows over a water body) is adequately
described in all directions.

5.3.4.2 Roughness Selection


Specification of the roughness parameters for tidal analyses follows the same procedures
as for riverine conditions (Section 5.2.4.2). Some surge models can include different
bottom stress parameterizations. For example, ADCIRC provides options for linear and
quadratic bottom friction assignment in addition to a Manning’s n formulation. Refer to the
individual model documentation for roughness specification other than Manning’s
coefficient. Most wave models also include options for bottom friction. For example, the
SWAN model includes frictional dissipation via the methodologies of JONSWAP, Collins,
and Madsen. Again, refer to the software documentation for recommended values of
friction parameters.

Roughness values through developed areas, inundated during the surge, are especially
difficult to predict. The density of buildings is a key influence on roughness in these areas.
Calibration data are helpful in targeting the proper roughness value.

Chapter 5: Bridge Hydraulics 5-99


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.3.4.3 Model Geometry


Model geometry refers to the spatial resolution incorporated into the model to describe
the waterway bathymetry and overbank topography. For one-dimensional models, this
refers to not only the cross section locations, but also the number of points across the
cross section. For two-dimensional models, this refers to the nodes and elements that
comprise either the finite element mesh or the finite difference grid.

One-Dimensional Models
Specification of one-dimensional model geometry for tidal analyses follows the same
recommendations as for riverine analyses (Section 5.2.4.3.1). In general, the only
difference is the size of the model domain, which is discussed in Section 5.3.4.1.

Two-Dimensional Models
Specification of two-dimensional model geometry for tidal analyses follows the same
recommendations as for riverine analyses (Section 5.2.4.3.2). Again, the only difference
is the size of the model domain, discussed in Section 4.4.1, which can extend into the
offshore area. Adequate resolution should be incorporated into the model to resolve tidal
inlet and offshore features (such as flood and ebb shoals, or coastal structures) that affect
the flow properties of the inlets.

5.3.4.4 Boundary Conditions


Boundary conditions for tidal analyses depend upon the types of simulations, the models
employed, and site-specific properties. One-dimensional modeling of coastal bridges
during surge events typically involves specification of an upstream flow boundary
condition and an oceanward stage boundary condition where the stage is an open coast
hurricane hydrograph. Two-dimensional surge modeling has more options for boundary
conditions. These can include:

 Specifying the stage and flow similar to the one-dimensional model


 Specifying the same boundary conditions as above, with an additional wind
boundary condition specified over the entire model domain
 Specifying tidal constituent boundary conditions on the offshore, upstream flow,
and meteorological forcing (wind and pressure) at each node

This section describes several of the possible model boundary conditions for coastal
bridge hydraulics analyses.

Chapter 5: Bridge Hydraulics 5-100


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Upstream Flow Boundary Conditions


Specification of upstream flow boundary conditions follows the same recommendations
as those for riverine flow boundary conditions (Section 5.2.4.1), with some exceptions. In
tidal analyses in Florida, inland boundaries typically are located far from the bridge
locations. This is done to accurately describe the storage inland of the bridge, which is a
significant factor in determining flow through the bridge. In general, bridges over low-
elevation, wide floodplains inland will experience more flow during a surge than bridges
over high-elevation, narrow floodplains inland. This is because low-elevation, wide
floodplains have substantial storage compared to high-elevation, narrow floodplains. The
greater storage makes the floodplain less responsive to incoming flood flow of a storm
surge, so the stage of low-elevation, wide floodplains rises more slowly than for
floodplains with less storage. This creates a greater difference in water surface across
the bridge, which increases the flow rate through the bridge.

The hydrology for the boundary condition should be developed for the bridge location
rather than at the location where the boundary condition is applied. Hurricane hydrology
is discussed in Section 5.3.2.

Storm Surge Hydrographs


A frequent type of coastal bridge hydraulics analysis involves application of an open coast
storm surge hydrograph as the oceanward boundary condition. Fortunately, in Florida,
several agencies have developed coastal surge elevations associated with several return
period intervals. In a study for the Department, Sheppard and Miller (2003) reviewed the
literature to determine what information was available regarding 50-, 100-, and 500-year
return interval open coast storm surge peak elevations and time history hydrographs.
Based on information from the literature review, the study developed recommendations
for selecting ocean boundary conditions for modeling inland storm surge propagation in
Florida’s coastal waters. From their findings, Sheppard and Miller recommended that the
Department employ the storm surge heights for 50-, 100- and 500-year return interval
hurricane storm surges developed by the FDEP. This recommendation was made on the
basis that FDEP had included all of the major surge generation mechanisms
(astronomical tides, wind setup, wave setup, etc.) in their analyses and that they had
compared their results with near coast water marks in buildings where possible. One
shortcoming of the FDEP values was that only the counties with sandy beaches (25 of
the 34 coastal counties) in Florida were analyzed by FDEP. To address this problem,
Sheppard and Miller developed surge elevations by interpolating values from the
surrounding counties using FEMA and NOAA results as guides. Figure 5.3-13 presents
the locations of the FDEP-developed elevations, as well as the locations of the
interpolated elevations (in italics).

Chapter 5: Bridge Hydraulics 5-101


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.3-13: Storm Surge Peak Elevation and Hydrograph Locations

The above guidance and supporting report are available at the following website:

http://www.dot.state.fl.us/rddesign/Drainage/FCHC.shtm

Hurricane-Generated Winds
For bridges located near the ends of bays and estuaries, wind setup can be a major
contributor to the surge elevation. Figure 5.3-14 illustrates the effects that local wind setup
can have on surge elevations. It displays results of a hindcast of the 1852 Unnamed
Hurricane in Tampa, Florida, at the Courtney Campbell Bridge near the northern end of
Old Tampa Bay. Hindcasts were performed with meteorological (spatially and temporally
varying wind and pressure fields) boundary conditions and tidal constituent forcing on the
offshore boundary. The line labeled Surge and Wind includes the “real” hindcast. For the

Chapter 5: Bridge Hydraulics 5-102


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

simulation represented by line labeled Surge Only, the wind speeds in the boundary
condition file were set to zero only at inland locations. Thus, this line represents the case
where surge at the bridge is only created from propagation of the surge hydrograph
inland.

3.5

Surge and Wind


3.0
Surge Only
2.5
Water Surface Elevation (m-MSL)

2.0

1.5

1.0

0.5

0.0

-0.5

-1.0
10 30 50 70 90 110
Simulation Time (hrs)

Figure 5.3-14: Surge Elevations at the Courtney Campbell Bridge Location during
the 1852 Unnamed Hurricane both with and without Local Wind Effects

Another example of how bridge location affects the importance of wind setup is seen in
the hindcast of Hurricane Ivan in 2004 that made landfall near Pensacola, Florida. Figure
5.3-15 displays the calculated storm surge elevation time series at the Interstate 10 (I-10)
Bridge over Escambia Bay (red line) and at the Pensacola Bay Bridge (blue line). Located
near the back of Escambia Bay, the I-10 Bridge experienced a significantly higher storm
surge than did the Pensacola Bay Bridge even though the Pensacola Bay Bridge was
located nearer to the inlet. This is directly attributable to the wind setup that occurred near
the back of Escambia Bay.

Chapter 5: Bridge Hydraulics 5-103


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

3.5

3
Water Surface Elevation (m-MSL)

2.5

1.5

0.5

0
Pensacola Bay Bridge
-0.5
I-10 Bridge

-1
0 20 40 60 80 100 120 140
Simulation Time (hr)
Figure 5.3-15: Hindcasted Surge Elevations at the I-10 over Escambia Bay Bridge
and Pensacola Bay Bridge during Hurricane Ivan 2004.

Figure 5.3-15 shows that hurricane winds can play a major role in describing surge
propagation. The reference AASHTO Guide Specifications for Bridges Vulnerable to
Coastal Storms (AASHTO 2008) provides a methodology for determining peak design
wind speeds for a number of mean recurrence intervals. It references ASCE Standard 7-
05 as the source for determining design wind speeds throughout the country. The
AASHTO Specification also states that if design coastal storm wind speeds exist at a site,
then these values should be used.

In Florida, Dr. Michel Ochi at the University of Florida (Ochi, 2004) presents a
methodology for predicting the hurricane landfall wind speeds along the Florida coast. He
examined tropical cyclones (including hurricanes) that landed on or passed nearby the
Florida coast from the NOAA hurricane database HURDAT. He divided the Florida coast
into 15 districts (Figure 5.3-16), and developed expected extreme values for different
return periods. Table 5.3-3 gives the expected maximum sustained (1-min average) wind
speed for landfalling hurricanes calculated from Ochi’s methodology.

Chapter 5: Bridge Hydraulics 5-104


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.3-16: Locations of Coastline Division Employed in Wind Speed Analysis


by Ochi (2004) (Source: Ochi, 2004)

Chapter 5: Bridge Hydraulics 5-105


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Table 5.3-3: Example of Extreme Landfall Wind Speeds for Florida using the Ochi
Methodology
Most Probable Maximum
Sustained Wind Speed (mph)
50- 100- 500-
District* year year year
K 130.9 141.4 162.3
1 110.5 120.5 140.5
2 107.0 116.6 135.7
3 97.5 107.5 127.5
4 82.9 88.8 100.3
5 104.0 115.3 138.4
6 89.7 101.3 125.1
7 96.8 112.4 144.9
8 127.1 137.9 159.4
9 136.5 148.0 171.2
10 140.2 147.7 162.8
11 104.0 112.0 127.6
* Districts 12-14 did not have enough storm impacts to generate a confident statistical analysis.

Hurricane Hindcasts
Hurricane hindcasts simulate the wave and surge climate associated with a unique
historical hurricane (Section 5.3.5). These types of simulations are performed primarily
with two-dimensional models. Boundary conditions typically take the form of temporally
and spatially variable wind and pressure fields (meteorological boundary conditions)
applied over the entire model domain. Additional boundary conditions include an offshore
stage boundary condition equal to the daily tidal fluctuation at the condition locations. This
can take the form of either specified tidal elevation time series (e.g., tidal hydrographs) or
be a feature of the model as selected tidal constituents (e.g., ADCIRC). The best source
for tidal hydrographs is NOAA’s Center for Operational Oceanographic Products and
Services (http://www.co-ops.nos.noaa.gov/) for real-time and measured tidal gage data,
as well as tidal prediction.

Hurricane wind and pressure fields can be developed in a number of ways. They range
from simple analytic models (e.g., Holland, 1980) to three-dimensional modeling. Several
agencies—including FEMA, NOAA, and USACE—have performed hindcasts of specific
storms. These hindcasts are available sometimes upon request. Additionally, several
commercially available sources of hindcast data also exist.

Chapter 5: Bridge Hydraulics 5-106


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.3.4.5 Bridge
When constructing a model to simulate hurricane surge propagation and wave climate,
you will need an accurate representation of the bridge and its influence on the
hydrodynamic processes. In general, the same techniques employed for riverine analyses
also apply to the analysis of coastal bridges during storm surges.

Roughness
Roughness specification at bridge cross sections for tidal analyses follows the same
recommendations as for riverine analysis (Section 5.2.4.2).

Bridge Routine
Selection of the appropriate bridge routines for tidal analyses follows the same
recommendations as for riverine analysis (Section 5.2.4.5).

Piers
Incorporating the effects of bridge piers into the hydraulic model for analysis of coastal
bridges follows the same procedure as for riverine bridges (Section 5.2.4.5). For two-
dimensional modeling, typically, you would not model piers directly because their
planform areas are significantly smaller than the areas of elements that resolve the bridge
openings. However, there are several options for including the effects of bridge piers.
Several models incorporate the loss effects into the hydraulic computation routines. An
example is FST2DH (part of FESWMS). FST2DH contains an automatic routine that
accounts for the effect of piers or piles on flow by increasing the bed friction coefficient
within elements that contain them (Froelich 2002). ADCIRC also contains routines for
incorporating the effects of bridge piers through a loss term in the momentum equations
due to the pier drag (http://adcirc.org/home/documentation/special-features/).

Gosselin et al. (2006) examined the effects of resolving bridge piers through element
elimination in cases where the pier width was a large percentage (5 percent to 35 percent)
of the overall bridge cross section top width. The piers were incorporated by deleting
elements within the mesh occupied by the piers. The authors compared results of the
two-dimensional modeling with one-dimensional modeling results for the same geometry
and flow conditions. The results compared well at the bridge cross section, but compared
poorly downstream of the piers. The authors concluded that whereas the one-dimensional
model incorporates the frictional losses from the piers through an increase in the wetted
perimeter, by modeling the piers through element deletion, the two-dimensional model
does not account for frictional losses if using a slip boundary condition along the model
edges. Rather, you can attribute losses from the piers to the momentum losses
associated with the creation of the secondary flows around the piers and in the wake
region.

Chapter 5: Bridge Hydraulics 5-107


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Regarding wave models, most publicly available software does not include effects of
bridge piers on wave propagation.

5.3.5 Simulations
Following construction of the surge and wave model domains, development of the
boundary conditions, and specification of the input model parameters, you can begin
running the model simulations. This section describes the model simulations typically
performed as part of the hydraulic analysis of a coastal bridge.

5.3.5.1 Model Calibration


Before performing design simulations, you should calibrate the surge and wave model
properly. Typically, you evaluate model performance through calibration and verification
both qualitatively and quantitatively, involving both graphical comparisons and statistical
tests. For surge models, calibration should include both tidal propagation simulations and
historical storm events. For wave models, calibration is achieved by comparing tidal
simulations for a period of record to either measured data collected at specific locations
or to widely available NOAA predictions at several locations. FEMA (2007) recommends
that your calibration results for amplitude variation throughout the domain and phase
variation be within 10 percent. In general, you typically do not perform flow rate or velocity
calibration because of lack of reliable data. Flow calibration is more difficult to achieve
than for water surface elevation data. However, if these data are available, acceptable
limits for calibration should be more generous than those for tidal amplitude, yet still
provide reasonable representation of the flow. FEMA also indicates that failure to achieve
calibration may be indicative of inadequate grid resolution, especially at inlets and other
critical points. Zevenbergen et al. (2005) provides a thorough description of model
troubleshooting, including suggestions for addressing model execution failures, numerical
instability, and calibration problems. These suggestions are contained in Table 5.3-4:

Chapter 5: Bridge Hydraulics 5-108


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Table 5.3-4: Suggestions for Model Calibration (Source: Zevenbergen et al. (2005))
If a model fails to execute, The causes of numerical Model calibration will be
check: instability are: affected by:
• Program output • Computational time • Appropriate model
error messages step too long extents
• Missing input data • Lack of geometric • Accuracy of model
• Incorrect input data refinement bathymetry
• Missing input files • Wetting and drying • Correct datum
• Inconsistent input problems conversions for
data • Weir flow bathymetry
• Correct datum
conversions for tide
gages
• Inclusion of wind
effects
 Inclusion of
appropriate
upstream inflow

Calibration to known storm events is significantly more complex than tidal calibration.
Ideally, the calibration would include accurate measurements of both the model inputs
(surge hydrograph or wind and pressure fields), as well as accurate surge measurements
at locations throughout the model domain (gage measurements or high water marks).
This is seldom the case. In fact, high water marks provide one of the more difficult data
sources to calibrate to since they often contain effects of local wave climate and can vary
significantly in close proximity to each other. If reliable information is available, calibration
to a known storm event is ideal. Comparison of model results with gage data or high water
marks helps identify problems with domain extents, model resolution, grid resolution, or
friction assignment.

Calibration of wave models also is difficult because calibration data are rarely available.
If you can acquire the data, then the calibration process should involve qualitative and
quantitative comparisons of measured and simulated wave height, period, and direction.
However, if measurements are unavailable, then the coastal engineer should
demonstrate that the wave model simulations provide reasonable results, were performed
employing accepted standards for input parameters, and incorporate an appropriate level
of conservatism.

Chapter 5: Bridge Hydraulics 5-109


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.3.5.2 Storm Surge Simulations


Storm surge simulations should include, at a minimum, the design and check events for
scour and the design frequency event for the bridge as specified in Section 5.1.3 (e.g.,
the 50-year for mainline interstate, high use, or essential bridges). Results from the
simulations include time series of water surface elevation, velocity, and flow rate. Extract
simulation results not only at the bridge cross section, but at locations upstream of the
bridge piers (for local pier scour calculation). The length of the bridge dictates the number
of locations. For shorter bridges, extracting conditions at the location of the maximum
velocity will be sufficient. For longer bridges, there will be greater variation in velocity
magnitude and direction. Thus, you should extract results at a greater number of locations
to resolve the variation. Extract flow rates and water depths upstream of the bridge
constriction for contraction scour calculations.

Figure 5.3-17 displays an example of water surface elevation and velocity time series
during the 100-year return period hurricane through Wiggins Pass near Naples, Florida.
The figure is typical of storm surge propagation through coastal waters. A peak in velocity
magnitude precedes the peak in water surface elevation as the surge propagates inland.
A second peak in velocity magnitude occurs as the surge recedes. The magnitude, phase,
and duration of the velocity magnitude peaks are a function of the shape of the surge
hydrograph and the response of the interior waterways.

Chapter 5: Bridge Hydraulics 5-110


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

18
Velocity (ft/s)
16 Water Surface Elevation (ft-NAVD88)
Water Surface Elevation (ft-NAVD88)

14
Velocity Magnitude (ft/s),

12

10

-2
0 5 10 15 20 25 30 35 40
Simulation Time (hr)

Figure 5.3-17: Example of Water Surface Elevation and Velocity Time Series
during the 100-year Return Period Hurricane through Wiggins Pass near Naples,
FL

5.3.5.3 Design Considerations


Typically, coastal bridges are not located in FEMA floodways and are not examined for
their effects on backwater. As the designer, you would select the bridge location and
profile for reasons related to right of way, environmental impacts, navigation, corrosion,
etc., rather than for bridge hydraulics (backwater impacts). Review the recommendations
contained in Section 5.2.5.3 for riverine studies to determine whether they apply for a
particular coastal bridge location. Situations that do require comparison of existing and
proposed conditions include: major modifications to the bridge profile or to the floodplain
(e.g., causeway islands), bridge replacements that transition from spill-through to wing-
wall abutments, etc.

An additional design consideration involves vessel collision. The LRFD specifications


require using the “average current velocity across the waterway.” Determining this

Chapter 5: Bridge Hydraulics 5-111


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

velocity for tidal flows requires a separate simulation of the spring tidal flows. The average
current velocity should correspond to the peak velocity occurring over this simulation.

5.3.5.4 Wave Simulations


Wave parameters are necessary both for calculation of wave forces on bridge
superstructures and for design of abutment protection. According to AASHTO (2008),
calculate wave forces (discussed in Section 5.3.5.6) from 100-year return period wave
conditions only. Similarly, design abutment protection to resist the 100-year wave
conditions. The wave model should simulate, at a minimum, the 100-year return period
hurricane-generated wave conditions at the site.

Time-dependent (unsteady) wave modeling gives more accurate design wave conditions
at the bridge location. As an alternative, steady-state modeling of the wave conditions
during the peak storm surge provides sufficient, though conservative, design conditions.
Inputs to the wave modeling will include design wind speeds, water surface elevations,
bathymetry/topography, and wind direction. If the wind direction is unknown, the wave
modeling should include, at a minimum, steady-state simulations of the wind field along
the direction of the longest fetches (Figure 5.3-18).

Wave models typically provide the significant wave height and the peak period. The
significant wave height is a statistical parameter representing the average of the highest
one-third of the waves in a wave spectrum. The peak period is the wave period
corresponding to the maximum of the wave energy spectrum. For design of bridge
superstructures, AASHTO recommends employing the maximum wave height rather than
the significant wave height. The AASHTO equation for converting between the two is Hmax
= 1.80 Hsignificant.

Chapter 5: Bridge Hydraulics 5-112


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.3-18: Example of Significant Wave Height Contours from Wave Modeling

5.3.6 Wave Forces on Bridge Superstructures


Bridge design must consider wave forces on bridge superstructures to prevent the type
of damage experienced at the I-10 bridge over Escambia Bay during Hurricane Ivan in
2004 (Figure 5.3-19 ). Section 4.9.5 of the Drainage Manual and Section 2.5 of the
Structures Design Guidelines address wave forces on bridge superstructures. The
bulletin provides guidance on applying the specifications in the AASHTO Guide
Specifications for Bridges Vulnerable to Coastal Storms to Department bridges. For
bridges spanning waters subject to coastal storms, it states that the superstructure low

Chapter 5: Bridge Hydraulics 5-113


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

chord must have a minimum one-foot vertical clearance above the 100-year design wave
crest elevation. If this clearance cannot be met, the bridge superstructure should be
raised as high as feasible and the bridge superstructure designed to resist storm wave
forces. For these bridges, the design strategy depends on the importance/criticalness of
the bridge when considering the consequences of bridge damage caused by wave forces.
If you judge a bridge to be extremely critical, you would design it to resist wave forces.
Bridges that you might judge to be non-critical do not require evaluation for wave forces.

Figure 5.3-19: Damage to the I-10 Bridge over Escambia Bay during Hurricane
Ivan (2004)

Figure 5.3-20 defines the parameters involved in estimating wave forces and moments
on bridge superstructures from the AASHTO Specifications. The interaction between the
wave and bridge superstructure produces vertical (uplift) forces, horizontal forces, and
over-turning moments. Computing design surge/wave-induced forces and moments on
bridge superstructures requires knowledge of the meteorological and oceanographic
(met/ocean) design conditions and the proper force and moment equations. The AASHTO
Specifications provide methods to determine both.

Chapter 5: Bridge Hydraulics 5-114


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.3-20: Definition Sketch for Wave Forces

The AASHTO Specifications provide a series of parametric equations for calculating the
wave forces. There are two sets of equations—one corresponds to the time of the
maximum vertical force and one corresponding to the time of the maximum horizontal
force. For example, for the maximum vertical force, the vertical force is the maximum
value experienced by the structure during passage of the design wave and the horizontal
force and moment are the values at the time of maximum vertical force.

Chapter 5: Bridge Hydraulics 5-115


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.4 MANMADE CONTROLLED CANALS


Manmade controlled canals have the following typical characteristics:

 They will have some type of downstream control structure, such as salt water
intrusion barriers, flood control weir, and/or pumps that will regulate the discharge.
 They will not normally flood out of bank, even in a 100-year storm.
 They have low design velocities—typically 1 fps to 3 fps—and often are subject to
aggradation requiring periodic dredging to maintain the needed cross section
 Their abutments typically do not encroach into the cross section of the canal;
therefore, there will be no contraction of flow and little backwater caused by the
bridge.
 Even if there are piles in the flow of the canal, the design discharge will not create
substantial scour around the piles because the velocity is low and the pile size
typically is small.
 Usually, the canal owner can provide the hydraulic design discharge and stage.

Given the typically innocuous hydraulic and scour conditions at controlled canal bridges,
you will find that the prudent level of effort required to perform the bridge hydraulics
analysis is considerably less than for the typical bridge. In fact, you can abbreviate the
traditional Bridge Hydraulics Report. Use the following outline for topics that should be
included for controlled canals:

5.4.1 Introduction
 Bridge Location Map
 Waterway owner (LWDD, SFWMD, CBDD, etc.)
 Description of waterway: manmade, straight, controlled canal, etc.
 Use of canal: navigation, recreation, flood protection, irrigation, etc.
 Other unusual details

5.4.2 Watershed Description & Flow


 Basin map from Water Management District or permitting agency
 Any available information on drainage area: maps, acreage, control structures, etc.
 Design discharge and stage information from owner: usually 10- or 25-year (Note:
If design frequency information is less than frequency requirements in the
Drainage Manual for hydraulic or scour design, consult the District Drainage

Chapter 5: Bridge Hydraulics 5-116


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Engineer. Also, if the design discharge and stage are not available, then a full
bridge hydraulics analysis is needed.)
 Testimony from Bridge Inspection records: aggradation/degradation, condition of
revetment, debris problems, etc.

5.4.3 Channel Excavation, Clearance, and Other Owner Requirements


 Required canal typical section from owner
 Lateral limits of channel excavation—usually 10 feet beyond bridge drip edge
 Any other pertinent information from owner: sacrificial pile, bank overtopping,
vertical and horizontal clearance requirements, etc.

5.4.4 Scour Estimation


 General scour—usually none due to lack of natural meander and tendency toward
aggradation
 Contraction scour—none if no overbank flow, unless pile blockage is > 10 percent
of the waterway width
 Typically, pier scour on controlled canals is less than five feet; with no additional
general or contraction scour, the CSU equations may be used

5.4.5 Abutment Protection


 Refer to Minimum Abutment Protection in Section 4.9.1 of the Drainage Manual
 Boat wakes and wave impact may dictate more robust abutment protection than
would be needed to protect for the flood flow velocities; consider this and document
as needed
 Owner may have specific requirements for abutment protection

5.4.6 Bridge Deck Drainage


Refer to Section 3.9 of the Drainage Manual, and Appendix H and Section 5.6 of this
document.

5.4.7 Appendix
 Correspondence with owner regarding canal design parameters and requirements
 Pictures from Bridge Inspection Reports, if significant
 Evidence of field review

Chapter 5: Bridge Hydraulics 5-117


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.5 BRIDGE SCOUR


Lowering the streambed at bridge piers is referred to as bridge sediment scour or simply
bridge scour. Bridge scour is one of the most frequent causes of bridge failure in the
United States and a major factor that contributes to the total construction and
maintenance costs of bridges in the United States. Under-predicting design scour depths
can result in costly bridge failures and possibly in the loss of lives; while over-predicting
can result in significant construction cost increases. For these reasons, proper prediction
of the amount of scour anticipated at a bridge crossing during design conditions is
essential. Policy on scour estimates can be found in the Section 4.9.2 of the Drainage
Manual.

For new bridge design, bridge widenings, and evaluation of existing structures, develop
scour elevation estimates for each pier/bent for the following conditions:

1. Worst-case scour condition (long-term channel processes, contraction scour and


local scour) up through the design flood event (Scour Design Flood Event)
2. Worst-case scour condition (long-term channel processes, contraction scour and
local scour) up through the check flood event (Scour Check Flood Event)
3. Long-term scour for structures required to meet the extreme-event vessel collision
load; “long-term scour” refers to either everyday scour for live-bed conditions or
the 100-year total scour for clear-water conditions; refer to Section 5.5.2 for further
discussion

Include the components discussed in the following sections in your scour estimates.

5.5.1 Scour Components


For engineering purposes, sediment scour at bridge sites is divided into three categories:

1. Long-term channel processes (channel migration and aggradation/degradation)


2. Contraction scour
3. Local scour

5.5.1.1 Long-Term Channel Processes


Scour associated with long-term channel processes is the change in bed elevation
associated with naturally occurring or manmade movement of the reach over which the
bridge is located. These bed changes are characterized both as horizontal changes
(channel migration) and as vertical changes (aggradation/degradation).

Chapter 5: Bridge Hydraulics 5-118


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Changes upstream and downstream affect stability at the bridge crossing. Natural and
manmade disturbances may change sediment load and flow dynamics, resulting in
adverse changes in the stream channel at the bridge crossing. These changes may
include channel bank migration, aggradation, or degradation of the channel bed. During
aggradation or degradation of a channel, the channel bed and thalweg tend to accrete or
erode.

Channel stability, as characterized by channel migration and aggradation/degradation of


the channel bed, is an important consideration in evaluating the potential scour at a bridge
for two reasons. First, because aggradation and degradation influence the channel’s
hydraulic properties and, second, because bank migration, thalweg shifting, and
degradation may cause foundation undermining regardless of whether the bridge
experiences the design event.

Channel Migration
Lateral channel migration is an important factor to consider when deciding on a bridge’s
location. Factors affecting lateral channel migration include stream geomorphology,
bridge crossing location, flood characteristics, characteristics of the bed and bank
material, and wash load (Richardson and Davis, 2001).

There are techniques to address channel migration in the FHWA document HEC 20
(Legasse et al., 2001). These techniques generally include critical examination/
comparison of historical measurements/records combined with field observations to
forecast future trends. Sources of historical records include bridge inspection records,
historical maps, historical aerial photography, and historical surveys. In general, at
bridges where the waterway exhibits a history of meandering, the hydraulics engineer
should consider assuming that the elevation of the thalweg could occur at any point within
the bridge cross section, including along the floodplain. If this conservative approach is
excessively costly, it may be more cost-effective to mitigate potential future meander by
river training or armoring.

Chapter 6 of HEC 20 (Legasse et al., 2001) provides procedures for predicting and
evaluating lateral channel migration through aerial photograph analysis. See Section
5.2.1.3 for sources of aerial photographs.

A special case of migration found in coastal zones is inlet migration. Inlets either migrate
along the coast or remain fixed in one location. This is due to a complex interaction
between the tidal prism (volume of water transported through the inlet during tides), open
coast wave energy, and sediment supply. Although many of Florida’s inlets are improved
through jetty construction and bank stabilization, several inlets are not—particularly along
the southwest coast. New bridge construction and evaluation of existing structures over
unimproved inlets should include a thorough investigation of the historical behavior of the
inlet (through examination of historical aerial photographs and charts) to discern the

Chapter 5: Bridge Hydraulics 5-119


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

migration trends to incorporate into the foundation design/evaluation, as well as design/


evaluation of the abutment protection. Types of inlet behavior can include:

 Updrift migration
 Downdrift migration
 Fluctuations in inlet width and depth
 Spit growth and breaching (resulting in oscillation of inlet location)

A coastal engineer should perform the analysis of coastal hydraulics for the design and
evaluation of bridges over tidal inlets. References and aids in design/evaluation include
the USACE’s EM 1110-2-1810 Engineering and Design—Coastal Geology (1995) and
EM 1110-2-1100 Coastal Engineering Manual (2006).

Aggradation/Degradation
Aggradation and degradation relate to the overall vertical stability of the bed. Long-term
aggradation and degradation refers to the change in the bed elevation over time over an
entire reach of the water body. Aggradation refers to the deposition of sediments eroded
from the channel or watershed upstream of the bridge resulting in a gradual rise in bed
elevation. Degradation refers to the gradual lowering of the bed elevation due to a deficit
in sediment supply from upstream.

Given the potential influence of changes in the watershed on stability at a bridge location,
you must not only evaluate the current stability of the stream and watershed, but also the
potential future changes in the river system (within reason). Examples of this include
incorporation of watershed management plans or known planned projects (bridge/culvert
replacements, dams, planned dredging, etc.) into evaluation of the vertical stability at the
bridge location. As such, it is important that you perform the necessary data collection
(including contacting local agencies) to become aware of future projects/plans and
incorporate them appropriately into the analysis.

For information on aggradation/degradation in riverine environments, refer to FHWA’s


HEC 18 and HEC 20. For more information, refer to the U.S. Army Corps of Engineers’
Coastal Engineering Manual.

For existing bridge locations, the most common evaluation of a channel’s vertical stability
is through examination of Bridge Inspection Reports. The reports (available upon request
from the individual Districts) typically contain recent and historical inspection survey
information. These surveys (typically lead-line surveys at each pier location on both sides
of the bridge) are an excellent source of data on long-term aggradation or degradation
trends. Additionally, inspection reports from bridges crossing streams in the same area
or region also can provide information on the behavior of the overall waterway if

Chapter 5: Bridge Hydraulics 5-120


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

information at a new location is unavailable. For new alignments, a review of historical


aerial photography is another method of channel stability analysis.

Estimate long-term vertical stability trends over the lifetime (for new projects) or remaining
lifetime (for evaluations of existing bridge or widening projects) of the subject bridge. If
the result is degradation, add the estimate at the end of the project life to the total scour.
If the result is aggradation, then document the estimate in the BHR. However, do not
include this estimate in the estimate of total scour. Rather, the existing ground elevation
should serve as the starting elevation for contraction and local scour.

As with channel migration, inlet stability is a special case of vertical stability. Examining
long-term trends through available historical information provides indicators of the inlet
behavior over time. Additionally, inlet stability analyses can provide information on the
evolutionary trends at the subject project. A qualified coastal engineer should perform
these analyses. The references USACE’s EM 1110-2-1810 Engineering and Design—
Coastal Geology (1995) and EM 1110-2-1100 Coastal Engineering Manual (2006)
provide additional resources.

5.5.1.2 Contraction Scour


Contraction scour occurs when a channel’s cross section is reduced by natural or
manmade features. Possible constrictions include the construction of long causeways to
reduce bridge lengths (and costs), the placement of large (relative to the channel cross
section) piers in the channel, the encroachment of abutments, and the presence of
headlands (examples in Figure 5.5-1 and Figure 5.5-2). For design flow conditions that
have long durations—such as those created by stormwater runoff in rivers and streams
in relatively flat country—contraction scour can reach near equilibrium depths. Equilibrium
conditions exist when the sediment leaving and entering a section of a stream is equal.
Laursen’s contraction scour prediction equations were developed for these conditions. A
summary of Laursen’s equations is presented below. For more information and
discussion, refer to HEC 18.

Chann Chann
Piers
el el

Roadwa
Figure 5.5-1: Examples of Contractions at Bridge Crossings

Chapter 5: Bridge Hydraulics 5-121


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.5-2: Example of Manmade Causeway Islands Creating a Channel


Contraction

Chapter 5: Bridge Hydraulics 5-122


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Steady, Uniform Flows


Laursen’s contraction scour equations (Laursen, 1960), or rather a modified version of
the equations recommended by HEC 18, were developed for steady uniform flow
situations. This methodology provides the estimation of contraction scour for most bridge
locations. However, predictions using these equations tend to be conservative, since the
rate of erosion decreases significantly with increased contraction scour depth. Laursen
developed different equations for clear-water and live-bed scour flow regimes. If the
estimates of contraction scour via these equations are deemed too conservative (through
application of engineering judgment), you may pursue alternative analyses, including
sediment transport modeling. In these situations, consult the District Drainage Engineer
regarding the need to perform such an analysis.

A brief summary of the HEC 18 equations are presented below. Refer to HEC 18 for more
information.

Live-Bed Contraction Scour Equation


The live-bed scour equation assumes that the upstream flow velocities are greater than
the sediment-critical velocity, Vc. The contraction scour in the section, ys, is calculated
from the equation below:
6
K1
y 2  Q2  7  W1 
   
y1  Q1   W2 

ys = y2 - yo = average contraction scour

where:
y1 = Average depth in the upstream channel, ft
y2 = Average depth in the contracted section after scour, ft
y0 = Average depth in the contracted section before scour, ft
Q1 = Discharge in the upstream channel transporting sediment, ft3/sec
Q2 = Discharge in the contracted channel, ft3/sec
W1 = Bottom width of the main upstream channel that is transporting bed material, ft
W2 = Bottom width of the main channel in the contracted section less pier widths, ft
K1 = Exponent listed in Table 5.5-1

Table 5.5-1: Determination of Exponent, K1


V*/ω K1 Mode of bed material transport
<0.50 0.59 Mostly contact bed material discharge
0.50 to 2.0 0.64 Some suspended bed material discharge
>2.0 0.69 Mostly suspended bed material discharge

Chapter 5: Bridge Hydraulics 5-123


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

where:
V* = (τo/ρ)0.5, Shear velocity in the upstream section, ft/sec
ω= Fall velocity of bed material based on the D50, ft/sec (Figure 5.5-3)
g= Acceleration of gravity, 32.17 ft/sec2 (9.81 m/s2)
τo = Shear stress on the bed, lbf /ft2 (Pa (N/m2))
ρ= Density of water, 1.94 slugs/ft3 (1,000 kg/m3)

Figure 5.5-3: Fall Velocity of Sediment Particles with Diameter Ds and Specific
Gravity of 2.65 (Source: HEC 18, 2001)

HEC 18 provides guidance for selecting upstream cross section locations, as well as the
widths at the bridge and upstream cross sections. Notably, separate contraction scour
calculations should be performed for the channel and left and right overbank areas
(assuming they extend through the bridge). For cross sections that include multiple
openings (including causeway bridges), upstream width selection involves delineating the
flow patterns upstream of the bridge to properly identify the division of the flow from the
upstream sections to the bridge.

As stated previously, application of this methodology may result in overly conservative


estimates. See the subsection “Unsteady, Complex Flows” in this section for an
alternative methodology for calculating contraction scour.

Chapter 5: Bridge Hydraulics 5-124


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Clear-Water Contraction Scour Equation


The clear-water scour equation assumes that the upstream flow velocities are less than
the sediment-critical velocity. The contraction scour in the section, ys, is calculated from
the equation below:
3
 2 
7
K Q
y 2 =  2u 
 2

 Dm W 
3

ys = y2 - yo = average contraction scour

where:
y2 = Average equilibrium depth in the contracted section after contraction scour, ft
Q = Discharge through the bridge or on the set-back overbank area at the bridge
associated with the width W, ft3/sec
Dm = Diameter of the smallest non-transportable particle in the bed material (1.25 D50)
in the contracted section, ft
D50 = Median diameter of bed material, ft
W = Bottom width of the contracted section less pier widths, ft
yo = Average existing depth in the contracted section, ft
Ku = 0.0077 (English units) or 0.025 (SI units)

For a more detailed discussion of these equations, the reader is referred to HEC 18.

Unsteady, Complex Flows


Application of Laursen’s modified contraction scour equations at locations that experience
design flows that are either unsteady or exhibit a complex flow field sometimes results in
overly conservative estimates of contraction scour. These situations include cases where:
(1) the flow boundaries are complex, (2) the flows are unsteady (and/or reversing), and
(3) the duration of the design flow event is short, etc. In these situations, an alternative to
employing Laursen’s modified equations is to perform two-dimensional flow and sediment
transport modeling to estimate contraction scour depths (e.g., the USACE’s RMA2
hydraulics model and SED2D sediment transport model). In these situations, consult the
District Drainage Engineer regarding the need to perform sediment transport modeling.

5.5.1.3 Local Scour (Pier and Abutment)


You can divide local scour into pier and abutment scour. The main mechanisms of local
scour are: (1) increased mean flow velocities and pressure gradients in the vicinity of the
structure; (2) the creation of secondary flows in the form of vortices; and (3) increased
turbulence in the local flow field. Two kinds of vortices may occur: (1) wake vortices
downstream of the points of flow separation on the structure, and (2) horizontal vortices

Chapter 5: Bridge Hydraulics 5-125


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

at the bed and free surface due to stagnation pressure variations along the face of the
structure and flow separation at the edge of the scour hole.

You can divide local scour into two different scour regimes that depend on the flow and
sediment conditions upstream of the structure. Clear-water scour refers to the local scour
that takes place under the conditions where sediment is not in motion on a flat bed
upstream of the structure. If sediment upstream of the structure is in motion, then the local
scour is called live-bed scour.

For work in Florida, calculation of local pier scour must involve application of the Sheppard
Pier Scour Equations detailed in the FDOT Bridge Scour Manual (Sheppard, 2005) rather
than the CSU Pier Scour Equation when the total scour (long-term channel conditions,
contraction scour, and pier scour) is greater than five feet. The Florida Complex Pier
Scour Procedure is described in HEC 18, Fifth Edition. The Florida Complex Pier Scour
Calculator and Procedure can be downloaded at:
http://www.fdot.gov/roadway/Drainage/Bridge-Scour-Policy-Guidance.shtm .

A brief overview of Sheppard’s Pier Scour Equation and the Florida Complex Pier Scour
Procedure are presented below. Refer to the FDOT Bridge Scour Manual for detailed
guidelines and examples.

Sheppard’s Pier Scour Equations


Sheppard’s Pier Scour Equations target three dimensionless hydraulic and sediment
transport parameter groups to predict scour at simple piers. You can apply the equation
to both riverine and tidal flows and to sediment sizes typical within the continental U.S.
The equations give good results for both narrow and wide piers. The FDOT Bridge Scour
Manual includes a detailed discussion. The pier scour equations are summarized below:

In the clear-water scour range:


( 0 .4  V  1 .0 )
Vc

ys
= 2.5 f1 f2 f3
D*

In the live-bed scour range:


( 1.0 < V  Vlp )
Vc Vc

Chapter 5: Bridge Hydraulics 5-126


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

  V   Vlp V  
  -1   -  
ys   Vc   Vc Vc  
= f 2.2 + 2.5 f
D*
1
  Vlp  3
 Vlp  
 V -1   V -1  
  c   c  

and in the live-bed scour range above five feet:


V Vlp
( > )
Vc Vc

ys
= 2.2 f1 ,
D*

where:
 y 0 0.4 
f1  tanh  *   ,
 D  
   V   
2

f2  1-1.2 ln     ,
   Vc   
  D*  
   
  D50   , and
f3   1.2 -0.13 
 0.4  D  +10.6  D 
* *

     
D
 50  D
 50 
V1 = 5Vc
V2 = 0.6 g y 0
 V for V1 > V2
Vl p = live bed peak velocity =  1
 V2 for V2 > V1

where:
ys = Equilibrium scour depth, ft
D* = Effective diameter of the pier, ft
yo = Water depth adjusted for general scour, aggradation/degradation, and contraction
scour, ft
V = Mean depth-averaged velocity, ft/sec
Vc = Critical depth-averaged velocity, ft/sec
Vlp = Depth-averaged velocity at the live-bed peak scour depth, ft/sec
D50 = Median sediment diameter, ft

Chapter 5: Bridge Hydraulics 5-127


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Methodology for determining depth-averaged critical velocity and depth-averaged live-


bed peak velocity are found in the FDOT Bridge Scour Manual.

Florida Complex Pier Procedure


Most large bridge piers are complex in shape and consist of several clearly definable
components. While these shapes are sensible and cost effective from a structural
standpoint, they present a challenge for those responsible for estimating design sediment
scour depths at these structures. The Complex Pier Methodology applies to any bridge
piers different from a single circular pile. They can be composed of up to three
components referred to here as the column, pile cap, and pile group, as shown below in
Figure 5.5-4.

Figure 5.5-4: Complex Pier Components

The methodology is based on the assumption that a complex pier can be represented (for
the purposes of scour depth estimation) by a single circular pile with an “effective
diameter” denoted by D*. The magnitude of the effective diameter is such that the scour
depth at this circular pile is the same as that at the complex pier for the same sediment
and flow conditions. The problem of computing equilibrium scour depth at the complex
pier is, therefore, reduced to one of determining the value of D* for that pier and applying
Sheppard’s Pier Scour Equation to the circular pile for the sediment and flow conditions
of interest. The methodology to determine the total D* for the complex structure can be
approximated by the sum of the effective diameters of the components making up the
structure, that is:

D*  D*col  D*pc  D*pg

Chapter 5: Bridge Hydraulics 5-128


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

where:
D* = Effective diameter of the complex structure
D*col = Effective diameter of the column
D*pc = Effective diameter of the pile cap
D*pg = Effective diameter of the pile group

The procedure for computing local scour depth for complex piers is further divided into
three cases, as illustrated in Figure 5.5-5 below:

 Case 1 complex pier with pile cap above the sediment bed
 Case 2 complex pier with pile cap partially buried
 Case 3 complex pier with pile cap completely buried

Figure 5.5-5: Three Cases of Local Scour Depth for Complex Pier Computations

Refer to the FDOT Bridge Scour Manual for a more detailed discussion on the procedure
and the application of the equations.

HEC 18 also provides equations for calculating local scour at abutments. However, as
stated in the Drainage Manual, abutment scour estimates are not required when the
design provides the minimum abutment protection. Where you have significantly wide
floodplains with high-velocity flow around abutment, consider analyzing abutment spatial
requirements using HEC 23.

Chapter 5: Bridge Hydraulics 5-129


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.5.1.4 Scour Considerations for Waves


Waves are an important factor that you must address when designing bridges exposed
to long fetches. This is particularly true at bridge abutments and approach roadways.
Figure 5.5-6 displays an example of the damage waves can cause during a hurricane
event. The photograph shows the east approach to the I-10 Westbound Bridge over
Escambia Bay after Hurricane Ivan. During the storm, waves breaking on the shoreline
removed the undersized protection and eroded the fill at the approach slab, eventually
undermining it. Proper design of abutment protection to withstand wave impact will be
discussed in Section 5.5.4.

Many bridges in coastal environments incorporate seawalls into the design of abutment
protection. Scour at vertical walls occurs when waves either break on or near the wall or
reflect off the wall, thus increasing the shear stress at the bottom of the wall. This is known
as toe scour. Toe scour decreases the effective embedment of the wall and can threaten
the stability of the structure. Current USACE guidance (CEM, 2001) indicates that, as a
rule of thumb, the depth of scour experienced in front of a vertical wall structure is on the
same order of magnitude as the incident maximum wave height. Methodologies for
designing toe scour protection are presented in Section 5.5.4.

Figure 5.5-6: East Approach to the I-10 WB Bridge over Escambia Bay
After Hurricane Ivan (2004)

Regarding the impacts of waves on scour at bridge piers, laboratory modeling indicates
that vertical piles subject to both waves and currents experience an increase in the
effective shear stress at the bed. Additionally, there is an increase in the amount of

Chapter 5: Bridge Hydraulics 5-130


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

suspended sediment and, thus, the sediment transport in the vicinity of the pile as
compared with the transport associated with currents or waves alone. No current
analytical methods are available for design purposes. However, some sediment transport
models (e.g., SED2D) include methodologies for calculating the shear stress due to
combined waves and currents.

5.5.2 Scour Considerations for Ship Impact


Piers designed to resist ship impact include in their load combinations estimates of “long-
term scour.” This long-term scour is different from the long-term channel conditions
discussed in the previous section. The previous information referred to the lateral or
vertical long-term processes that occur at a bridge crossing over the lifetime of the bridge.
Rather, the scour incorporated into design for ship impact is the scour that may be present
at a pier when the impact occurs. For sites where everyday (normal daily) flows are in the
clear-water regime—i.e., below the critical value for incipient motion of the bed
sediments—this scour is the total 100-year scour for the structure. The reasoning is that
if a design event occurs during the lifetime of the bridge, the daily flows are not sufficient
to fill in the hole. For bridges where flows are in the live-bed regime, the "long-term scour"
is the normal, everyday scour at the piers combined with the degradation and channel
migration anticipated during the life of the structure. The reasoning here is that if the
structure experiences a design event, the flows are sufficient to refill the scour hole
following such an event.

For bridge replacements, parallel bridges, major widenings, etc., Bridge Inspection
Reports and the design survey should be the primary basis for determining normal
everyday scour. If the proposed piers are the same as the existing piers, the normal,
everyday scour elevation should be reflected in the inspection reports and the design
survey (Figures 5.5-7 and 5.5-8). Slight differences in scour will likely exist between
inspection reports and between the reports and the design survey. In these cases, an
average scour elevation will be a reasonable estimate of normal, everyday scour. If there
is a large difference, an extreme storm event may have occurred just before the inspection
or survey. Investigate this and address it on a case-by-case basis.

Chapter 5: Bridge Hydraulics 5-131


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Upstream Bed Cross Section

0
Distance from Top of Rail, (ft)

-5

-10

1992
-15 1994
2010

-20

-25
1 2 3 4
Bent

Figure 5.5-7: Example of Normal, Everyday Scour Holes


from Bridge Inspection Data

Chapter 5: Bridge Hydraulics 5-132


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.5-8: Example of Normal, Everyday Scour Holes from Survey Data

For structures in which the proposed piers will be a different size or shape than the
existing or for new bridges/new alignments where there are no historical records
available, base estimates of the normal everyday scour on hydraulic modeling results of
expected daily flows. For riverine bridges, this should correspond to flows equal to the
normal high water. For tidal flows, everyday flows correspond to the maximum flows
experienced during spring tides.

5.5.3 Florida Rock/Clay Scour Procedure


The Florida Rock/Clay Scour Procedure was developed to address the scour resistance
of cemented strata, rock, and clay. The procedure was originally developed for cohesive
bed materials considered “scourable” according to FHWA guidelines. Refer to HEC 18,
Fifth Edition, Chapter 4 for an explanation of rock characteristics that relate to strength
and scour potential. Consult the District Drainage Engineer and the District Geotechnical
Engineer before initiating the Rock/Clay Scour Procedure.

Chapter 5: Bridge Hydraulics 5-133


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

The test methods establish the shear stress response of soils and the procedure
integrates that response over the lifetime of expected flows at the bridge site. The
procedure involves establishing the shear stress response of a site-specific sample using
the RETA (Rotating Erosion Test Apparatus) and SERF (Sediment Erosion Recirculating
Flume) devices, shown below in Figures 5.5-9 and 5.5-10, respectively, and then
integrating that response over the flows expected in the life of the bridge to predict
contraction or local scour at the bridge.

Figure 5.5-9: Rotating Erosion Test Apparatus (RETA, above)

Chapter 5: Bridge Hydraulics 5-134


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.5-10: Sediment Erosion Recirculating Flume (SERF)

The procedure includes an appropriate amount of conservatism by incorporating the


following assumptions: (1) the shear stress does not decrease within a local scour hole,
(2) the bridge experiences an extremely aggressive bridge flow history over the bridge
lifetime, (3) there is no refill of the predicted scour, and (4) only the more conservative of
the RETA and SERF results of all cores tested for a particular bridge characterize the
erosion properties of the bed. Districts should contact the State Drainage Engineer if
scour-resistant soils are expected to be encountered in bridge design or the evaluation of
existing bridge scour. The following link contains the FDOT Bridge Rock Scour Analysis
Protocol and describes initiation of the process:
http://www.fdot.gov/roadway/Drainage/Fla-Rockclay-Proc.shtm

5.5.3.1 Pressure Scour


See HEC 18 for detailed information on pressure scour.

5.5.3.2 Debris Scour


See HEC 18 for detailed information on debris scour.

Chapter 5: Bridge Hydraulics 5-135


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.5.4 Scour Countermeasures


Scour countermeasures are defined as a measure intended to prevent, delay, or reduce
the severity of scour problems. For this discussion, they address the class of armoring
countermeasures (as defined by HEC 23, Legasse et al., 2009) to resist the erosive forces
caused by a hydraulic condition. This section addresses scour countermeasures at both
abutments and interior bents.

5.5.4.1 Abutment Protection


Proper bridge design includes abutment protection to resist the hydrodynamic forces
experienced during design events. The Drainage Manual specifies the following minimum
protection requirements:

Spill-Through Abutments
Where flow velocities do not exceed 9 fps, and/or wave heights do not exceed 3 feet,
minimum protection consists of one of the following protection methods placed on a
1V:2H or gentler slope:

 Rubble riprap (Bank and Shore), bedding stone, and filter fabric—Rubble riprap
(Bank and Shore) is defined in the FDOT Standard Specifications for Road and
Bridge Construction, Section 530
 Articulated concrete block (cabled and anchored)—Articulating concrete block also
is defined in Section 530
 Grout-filled mattress (articulating with cabling throughout the mattress)

You must create site-specific designs when using articulated concrete block or grout-filled
mattress abutment protection. As of May 2016, the Department does not have standard
specifications for grout-filled mattresses. You will need to prepare a technical specification
if grout-filled mattresses are proposed for a project. The FDOT Structures Detailing
Manual provides typical details for standard revetment protection of abutments and extent
of coverage. Determine the horizontal limits of protection using HEC 23. Provide a
minimum distance of 10 feet if HEC 23 calculations show less than 10 feet. Notably,
neither grouted sand-cement bag abutment protection nor slope paving is considered
adequate protection for bridges spanning waterways. Slope paving can develop cracks
or upheaved slabs where loss of fill can occur. Grouted sand-cement bags often fail when
cracks form around the individual bags and sediment is lost through cracks or displaced
elements (Figure 5.5-11). Additionally, these systems are prone to failure due to
undermining (erosion at the toe of the protection) or flanking (erosion at the edges of the
protection) when the edges of the protection are not sufficiently buried.

Chapter 5: Bridge Hydraulics 5-136


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.5-11: Damage to Sand-Cement Grouted Riprap Abutment Protection

Determine the horizontal and vertical extents, regardless of protection type, using the
design guidelines contained in HEC 23. If the results from the HEC 23 calculations show
that a horizontal extent less than 10 feet is acceptable, you should still provide a minimum
of 10 feet. Review the limits of right of way to ensure the minimum apron width at the toe
of the abutment slope both beneath and around the bridge abutments along the entire
length of the protection. If calculations from HEC 23 result in a horizontal extent outside
the right of way limits, do the following:

Chapter 5: Bridge Hydraulics 5-137


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

a. Recommend additional right of way.


b. Provide an apron at the toe of the abutment slope that extends an equal distance
out around the entire length of the abutment toe. In doing so, consider specifying
a greater rubble riprap thickness to account for reduced horizontal extent (Figure
5.5-12).

Make additional considerations regarding extents in coastal areas subject to wave attack.
Prolonged exposure to hurricane-generated waves on unprotected approaches may lead
to damage to the approach slabs (Figure 5.5-6) as well as the approach roadways.
Consider extending the limits of protection to include the approach spans in wave-
vulnerable areas.

Figure 5.5-12: Example of Increased Toe Thickness to Offset


Decrease in Toe Width

When bridges are to be widened, you may not be able to simply recommend using
standard rubble riprap, as defined in Section 4.9 of the Drainage Manual. Constructability
issues may arise at existing bridges where the low chord elevations may prevent uniform
riprap placement due to height constrictions. If this case arises, you can do the following:

a. Rather than simply employing the minimum FDOT Bank and Shore Rubble Riprap,
size the rubble according to the design average velocities determined at the
abutment using HEC 23. This may result in smaller armor stone sizes, thus
enabling easier placement.
b. Provide an alternate material in the plans that should be approved prior to
installation.

Chapter 5: Bridge Hydraulics 5-138


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Bulkhead/Vertical Wall Abutments


You must protect abutments by sheet piling with rubble toe protection below the bulkhead,
and with revetment protection above the bulkhead when appropriate. Design the size and
extent of the protection for the individual site conditions.

Allow abutment protection to extend beyond the bridge along embankments that may be
vulnerable during a hurricane surge. You need to consider wave attack above the peak
design surge elevation and wave-induced toe scour at the foot of bulkheads. In such
cases, consult a qualified coastal engineer to determine the size and coverage of the toe
scour protection. The choice of cabling material for interlocking block or concrete
mattresses must consider the corrosiveness of the waterway. Avoid using steel cabling
in salt or brackish waters (stainless steel is permissible).

Rubble riprap abutment protection is the preferred protection type for new bridges. Rubble
riprap has several advantages (HEC 11), including:

 The riprap blanket is flexible and is not impaired or weakened by minor


movement of the bank caused by settlement or other minor adjustments.
 Local damage or loss can be repaired by placement of more rock.
 Construction is not complicated.
 Vegetation often will grow through the rocks, adding aesthetic and structural
value to the bank material and restoring natural roughness.
 Riprap is recoverable and may be stockpiled for future use.

A drawback to rubble riprap is that it can be more sensitive than some other bank-
protection schemes to local economic factors. For example, transport costs can
significantly affect the construction costs. For an illustration of bridge abutment slope
protection adjacent to streams, refer to the FDOT Structures Detailing Manual at the
following link:
http://www.fdot.gov/structures/structuresmanual/currentrelease/vol2sdm.pdf

Where velocities do not exceed 9 fps and waves do not exceed three feet on a 1V:2H
slope, protection should consist of a 2.5-foot-thick armor layer comprised of FDOT
Standard Bank and Shore Rubble Riprap over a one-foot thick layer of bedding stone
over filter fabric. Size the filter fabric appropriately to prevent loss of the fill sediments.
The purpose of the bedding stone is to ensure consistent contact between the filter fabric
and the soil; and to prevent the armor stone from damaging the filter fabric during
construction; and to inhibit movement during design events. Ensure the riprap has a well-
graded distribution to promote interlocking between the individual units, which improves
performance of the protection. For riverine applications, compare these minimums to the

Chapter 5: Bridge Hydraulics 5-139


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

guidance presented in HEC 23 (Design Guideline No. 14) to ensure proper design. A
notable feature of the slope protection cross-sections, illustrated in the FDOT Structures
Detailing Manual’s link above, is the sand cement bags located between the revetment
and the abutment. This detail was added to the Standard following field inspection
observations that the protection/abutment interface often was a point of failure. Shifting
of the stones during a minor event would cause a gap to open at the top of the slope,
allowing erosion to take place. This addition ensures that the filter fabric remains in
contact with the abutment so that any settlement will not produce a gap between the
structure and the stones.

For locations subject to wave impacts with wave heights greater than three feet, you must
also design the revetment to resist hurricane-generated waves. Design of abutment
protection should follow the same procedures and methodologies as design of rubble
riprap protection that serves as shore protection. The U.S. Army Corps of Engineers
provides guidance in the references (USACE, 2006, and USACE, 1995). USACE
Engineering Manual 1110-2-1614 (USACE 1995), in particular, provides multiple
methodologies for properly sizing armor stone as well as designing the revetment extents,
toe geometry, bedding stone, and armor layer distribution.

Often, this analysis will result in an armor stone size greater than that provided by the
FDOT Standard Bank and Shore Rubble Riprap. When this occurs, use the more
conservative (larger stone size) design. For these designs, develop a modified special
provision for the non-standard rubble riprap. The provision must specify the new riprap
distribution developed employing the techniques located in USACE (1995) or a similar
procedure. Develop a well-graded distribution to the armor stone to ensure optimal
performance. Additionally, for large armor stone, it may become necessary to include
additional intermediate stone layers into the design to prevent loss of bedding stone
between gaps in the armor stone. The USACE (1995) reference presents guidelines for
design of granular filter layers as a function of the armor stone size.

For toe scour protection, the USACE (1995) reference provides guidance on sizing stones
and designing the apron width. Toe apron width will depend on both geotechnical and
hydraulic factors. For a sheet-pile wall, you must protect the passive earth pressure zone.
The minimum width from a hydraulic perspective should be at least twice the incident
wave height for sheet-pile walls and equal to the incident wave height for gravity walls.
Additionally, the apron should be at least 40 percent of the depth at the structure.
Compare this apron width to that required by geotechnical factors and adjust it
appropriately. Regarding size of the armor stone, the reference provides a method
developed by Brebner and Donnelly. USACE (2006) also provides guidance for toe scour
protection in front of vertical wall structures in Section VI-5-6 of the Coastal Engineering
Manual.

Chapter 5: Bridge Hydraulics 5-140


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

For revetment installations where you don’t expect significant wave attack, include all
options that are appropriate based on site conditions (e.g., fabric-formed concrete,
standard rubble, cabled interlocking block, etc.; see Figure 5.5-13 through Figure 5.5-15).
HEC 23 provides guidance for design of these protection systems, as follows:

 Design Guideline 8—Articulating Concrete Block Systems


 Design Guideline 9—Grout-Filled Mattresses
 Design Guideline 14 – Rock Riprap at Bridge Abutments

Figure 5.5-13: Example of Rubble Riprap Abutment Protection

Chapter 5: Bridge Hydraulics 5-141


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.5-14: Example of Articulating Concrete Block Abutment Protection

Chapter 5: Bridge Hydraulics 5-142


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.5-15: Example of Grout-Filled Mattress Abutment Protection

Document options shown to be appropriate for the site in the BHR. You may write a
technical specification based on the use of the most ideal revetment material, with the
option to substitute the other allowable materials at no additional expense to the
Department. This recommendation would help to eliminate revetment CSIPs (Cost
Savings Initiative Proposals) during construction. No matter what options are allowed,
match the bedding (filter fabric and bedding stone) to the abutment material. Some of the
options are not self-healing (i.e., not rubble riprap), and a major failure can occur if loss
of the embankment material beneath the protection takes place.

As a final note, coastal bridges often incorporate seawalls into the abutment protection
design. The caps of these structures often have a low elevation (below the design surge
elevation) to tie into neighboring structures. Address the design of these structures as
containing elements of both spill-through and vertical wall abutments. The area in front of
the seawall should include a toe scour apron designed in the same manner as for vertical
wall abutments. Design areas between the seawall and the abutment using the same
procedures as spill-through abutments. These designs should ensure encapsulation of
the fill behind the seawall (Figure 5.5-16) to prevent loss of fill and potential failure of the
anchoring system (Figure 5.5-17).

Chapter 5: Bridge Hydraulics 5-143


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.5-16: Example of Abutment Protection Design Including a Seawall

Chapter 5: Bridge Hydraulics 5-144


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.5-17: Seawall Failure Following Hurricane Frances (2004)

Chapter 5: Bridge Hydraulics 5-145


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.5.4.2 Scour Protection at Existing Piers


For bridges evaluated as scour critical and where monitoring is not an option, one of the
countermeasures you should consider is a bed armoring countermeasure around the
critical pier. As with abutment protection, pier scour protection can take many forms.
Examples of these include rubble riprap, articulating concrete block, grout-filled
mattresses, gabion/marine mattresses, and partially grouted riprap. HEC 23 provides
design guidance for these protection systems in the following design guidelines (located
in Volume 2 of the reference):

 Design Guideline 8—Articulating Concrete Block Systems at Bridge Piers


 Design Guideline 9—Grout-Filled Mattresses at Bridge Piers
 Design Guideline 10—Gabion Mattresses at Bridge Piers
 Design Guideline 11—Rock Riprap at Bridge Piers
 Design Guideline 12—Partially Grouted Riprap at Bridge Piers

The guidelines provide:

 Procedures for selecting safety factors


 Methodologies for sizing the material
 Recommendations for designing coverage extents, filter requirements, and
installation guidelines

You will see several similarities between the procedures. All guidelines recommend
ensuring that the top of the protection remain level with the bed of the approach.
Suggestions for achieving this include placing sand-filled geotextile containers within the
scour hole to raise the bed elevation and serve as a filter for the overlaying protection.
The guidelines all also recommend that the horizontal extent of the protection extend a
distance equal to twice the effective diameter of the pier in all directions. For the non-
riprap options, the guidelines recommend that the protection slope away from the pier
with the edges of the protection buried below the maximum scour depth for the overall
cross section (i.e., depth of contraction scour and long-term degradation). A common
failure point of the non-riprap protection schemes is at the edges of the protection if the
mattress becomes undermined. Thus, it is important to incorporate trenching of the edges
and use of anchoring systems (if appropriate) into the protection design. Another common
failure point is at the pier/protection interface. The guidelines suggest grouting this
interface to prevent loss of fill for both the articulating concrete block and gabion
protection systems. You should review disadvantages and advantages of each system,
including construction feasibility and cost.

Chapter 5: Bridge Hydraulics 5-146


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.6 DECK DRAINAGE


To drain the deck of a bridge, there are three options, in order of preference:

1. Rely on the longitudinal grade of the bridge to convey the deck runoff to the end
of the bridge.
2. Use freely discharging scuppers or inlets to drain the deck runoff to the area
directly below the bridge. These sometimes are referred to as open systems.
3. Collect the discharge from the scuppers or inlets in a pipe system. The pipe
system can discharge down a pier or at the ends of the bridge. These systems
sometimes are referred to as closed systems.

Spread criteria will control the need to eliminate option 1 and use either option 2 or 3. The
inability to discharge to the area below the bridge will control the need to eliminate option
2 and use option 3.

5.6.1 Bridge End Drainage


If the profile grade of the roadway is sloping off of the bridge, roadway inlets collect runoff
from the bridge, often immediately beyond the bridge approach slab. Inlets typically are
not placed in the approach slab so that runoff does not seep between the concrete
approach slab and the roadway inlet. If spread issues mandate that you place an inlet in
the approach slab, obtain concurrence from the District Drainage Engineer and
coordinate with the District Structures Design Engineer.

For rural roadways, shoulder gutter is typically used to convey the bridge flow to a
shoulder gutter inlet (See Standard Plans, Index 425-040, Gutter Inlet Type S). This inlet,
including its 5-foot-long gutter transition, is usually located about 35 feet from the end of
the approach slab to provide space for the guardrail’s Approach Transition Connection to
Rigid Barrier, including its curb transition to shoulder gutter (See Standard Plans, Index
536-001, Guardrail). Additionally, check the spread at the shoulder gutter inlet for the 10-
year flow to ensure that runoff does not overtop the shoulder, causing erosion of the
embankment (refer to Chapter 6 and Appendix H for more information).

If the profile grade is sloping onto the bridge for rural roadways, then the calculations for
the deck drainage may need to include roadway runoff flowing onto the bridge. The
shoulder gutter transition directs the rainwater from the bridge into the inlet (refer to Figure
5.6-1). For standard cross slopes of 0.02 ft/ft for bridge shoulders and 0.06 ft/ft for
roadway shoulders, with a 10-foot wide shoulder, the longitudinal slope of the gutter due
to the transition is 2.1 percent. For this situation, the roadway grade would need to be
greater than 2.1 percent for roadway runoff to flow onto the bridge. Appendix H shows

Chapter 5: Bridge Hydraulics 5-147


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

how this slope was determined, and the same method can be used to calculate the slope
for other situations.

Figure 5.6-1: Shoulder Gutter Transition at Bridge End

For urban locations, if there is not a barrier wall between the sidewalk and the travel lanes,
or if there is no sidewalk, a curb inlet can be placed at the end of the approach slab.

The Drainage Manual does not require bridge sidewalk runoff to be collected on the
bridge. Scuppers or drains are not necessary to control the runoff on the bridge sidewalk
unless the runoff becomes great enough to overwhelm the collection system at the end
of the bridge. Scuppers used to drain the sidewalk must be ADA compliant.

In handling runoff from the sidewalk at the end of the bridge, the best option is to transition
the sidewalk slope toward the roadway immediately downstream of the bridge. The flow
then can be picked up in the first curb inlet or barrier wall inlet off of the bridge.

5.6.2 No Scuppers or Inlets (Option 1)


If possible, you should allow stormwater to flow to the end of the bridge and collect in the
roadway drainage system. To determine if this option is feasible, check the spread:

 Where the barrier wall or curb ends at the edge of the approach slab
 At the first inlet off of the bridge

Chapter 5: Bridge Hydraulics 5-148


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Calculate spread based on the Gutter Flow Equation in Section 6.3.2 of this document.
Spread criteria are given in Chapter 3.9 of the Drainage Manual. If the spread exceeds
the allowable spread criteria, then you may need scuppers or inlets on the bridge to
reduce the spread. If the spread exceeds the criteria, consider adjusting the profile grade
to reduce the spread before adding scuppers or inlets on the bridge. Reduce spread by:

 Steepening the longitudinal slope of the bridge at the bridge ends


 Including a profile crest in the middle of the bridge rather than using a profile that
slopes to only one end of the bridge

After determining grades that would eliminate the need for scuppers or inlets, talk with
the roadway designer to determine the feasibility of adjusting the profile grade.

Example 5.6-1
A bridge for a two-lane rural roadway has the following characteristics:

 200-foot length
 30-foot approach slabs
 A longitudinal slope of 0.3 percent
 Shoulder gutter inlets located 30 feet from the uphill approach slab
 The bridge typical section has two 12-foot travel lanes, 10-foot outside shoulders,
1.5-foot barriers, 0.02 ft/ft cross slopes, and is crowned in the middle.

Chapter 5: Bridge Hydraulics 5-149


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Solution:

Determine the drainage area to the end of the downhill approach slab. On the uphill end
of the bridge, the shoulder gutter transition will cause the runoff from the area between
the shoulder gutter inlet and the end of the approach slab to flow back to the shoulder
gutter inlet. Therefore, the drainage area contributing to the downhill side will include the
bridge deck and the approach slabs:

Area = (12+10+1.5) (30+200+30) / 43560 = 0.14 acres

Conversion from square ft. to acres


Approach slab length
Bridge length
Approach slab length
Width of barrier wall
Width of shoulder
Width of travel lane

Chapter 5: Bridge Hydraulics 5-150


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

The flow is:

Q = CiA = 0.95 (4) (0.14) = 0.53 cfs

where:
C = Rational runoff coefficient
i= Rainfall intensity, inches per hour
(4 in/hr, refer to Chapter 6 for explanation)
A = Drainage area, acres

Solving the gutter flow equation for spread:

3 3
 Qn  8
 (0.53)(0.016)  8
Spread   5 / 3 1/ 2 
 5/3 1/ 2 
 7.1 ft.
 0.56 S X S   0.56(0.02) (0.003) 

Since the spread at the end of the downhill approach slab is less than 10 feet, with 10
feet being the width of the shoulder, scuppers are not necessary.

You also can check the spread at the shoulder gutter inlet on the downhill side of the
bridge. There will be an additional drainage area from the end of the approach slab that
needs to be added to the drainage on the bridge. The drainage area to the shoulder gutter
inlet is:

Area = 0.14 + (((12+8+3.5+4) (30)) / 43560) = 0.16 acres

Conversion from square ft. to acres


Distance to inlet
Width behind shoulder gutter
Shoulder gutter
Width of shoulder
Width of travel lane
Drainage area from bridge

Assuming that the bridge is in Zone 1 for the IDF curves, the flow to the inlet is:

Q = CiA = 0.95 (7.0) (0.16) = 1.06 cfs

Chapter 5: Bridge Hydraulics 5-151


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

where:
i= The 10-year, 10-minute rainfall intensity = 7.0 inches per hour
(Refer to Chapter 6 for explanation)

Note that this value is slightly conservative. The one-foot unpaved strip behind the
guardrail was assumed to be paved in this calculation.

The allowable conveyance in the shoulder gutter is K = 28 cfs. Refer to Section 6.3.2.3 of
this document for further explanation of this value. The allowable flow at the shoulder
gutter inlet is:

Q = K S1/2 = (28) (0.003)1/2 = 1.53 cfs

Since the gutter flow just uphill of the shoulder gutter inlet is less than the allowable flow,
the deck drainage design is acceptable.

5.6.3 Scuppers (Option 2)


Scuppers typically are formed by tying PVC pipe into place prior to pouring the concrete
for the bridge deck (Figure 5.6-2). The deck runoff will flow into the scuppers, through the
deck, and then freefall to the ground or water surface below the bridge.

Figure 5.6-2: Standard FDOT Scupper Detail

Avoid placing scuppers over certain areas due to the direct discharge. These areas
include:

Chapter 5: Bridge Hydraulics 5-152


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

 Driving lanes, railroad tracks, and sidewalks


 Major navigation channels
 Bridge bents
 Erodible soil, unless the free discharge is at least 25 feet above the soil
 Environmentally sensitive water bodies as negotiated with permitting agencies
 Wildlife shelves, unless the bottom of the bridges is 25 feet or more above the shelf

As stated in Section 4.9.4 of the Drainage Manual, the standard scupper drain is four
inches in diameter and spaced on 10-foot centers, unless spread calculations indicate
closer spacing is required. Typically, the 10-foot spacing will provide adequate drainage
for most bridges. You can evaluate the intercepted flow for four-inch bridge scuppers on
a grade using the capacity curves in Figure 5.6-3 and Figure 5.6-4. The curves were
derived from laboratory studies performed at the University of South Florida (Anderson,
1973).

Grated scuppers or inlets, as shown in Figure 5.6-5, are more uncommon, especially as
free-draining scuppers. Although grated inlets can be used with open systems, they are
normally used with closed systems. You might use this type of grated scupper, or perhaps
one with a smaller grate, to drain a bridge sidewalk or if you expect significant bicycle or
pedestrian traffic on the shoulder. The four-inch ungrated scuppers will not meet ADA
requirements.

Chapter 5: Bridge Hydraulics 5-153


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.6-3: Intercepted Flow for 4-inch Bridge Scuppers


Cross Slope = 0.03 ft/ft

Figure 5.6-4 Intercepted Flow for 4-inch Bridge Scuppers


Cross Slope = 0.02 ft/ft

Chapter 5: Bridge Hydraulics 5-154


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.6-5: Grated Free-Draining Scupper

The Department does not have standard grated scuppers or inlets; therefore, it does not
have capacity charts as with other standard Department inlets. Section 6.3.1.5 provides
references to documents that you can use to derive inlet capacities. Manufacturers may
publish capacity charts for their inlets. Keep in mind that the pipe opening at the bottom
of the inlet may control the capacity rather than the inlet opening.

The length, width, and depth of the grated inlet will be limited by the reinforcement in the
deck of the bridge. You will need to coordinate the dimensions and locations of the inlets
with the structural engineer. Use standard prefabricated inlets whenever possible. Refer
to Section 7.4 for more information on grated scuppers.

Example 5.6-2
A bridge deck grated scupper is located where the shoulder width is 10 feet and the cross
slope is 0.02 ft/ft. The longitudinal grade of the bridge is 1.5%. The dimensions of the
grated scupper as defined in Figure 5.6-5 are:

W = 5 feet
L = 1 foot
D = 7 inches
Outlet Pipe Diameter = 8 inches

The flow along the barrier wall at the scupper is 1.65 cfs. Determine the intercepted flow.

Chapter 5: Bridge Hydraulics 5-155


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Solution:

The spread in the gutter prior to the inlet is:

3 3
 Qn  8
 (1.65)(0.016)  8
Spread   5 / 3 1/ 2 
 5/3 1/ 2 
 8.06 ft.
 0.56 S X S   0.56(0.02) (0.015) 

Calculate the intercepted flow using the method presented in FHWA Hydraulic
Engineering Circular No. 12, Drainage of Highway Pavements, March 1984 (HEC 12).

The flow directly over the grate is called the frontal flow. The frontal flow can be
determined using Equation 7 from HEC 12:

8 8
Q  W 3
 5  3
E 0  W  1  1    1  1    0.924
Q  T   8.06 

where:
E0 = Ratio of flow in width, W, to the total flow, Q
QW = Flow in width, W, less than T, in cfs
Q = Total flow, in cfs
W = Width of flow, W, in feet
T= Total width of flow (also called the spread), in feet

The frontal flow, QW = E0Q = 0.924 (1.65) = 1.52 cfs

The inlet will intercept all of the frontal flow unless the velocity is great enough to cause
the flow to skip over the grate. This velocity is called the splash-over velocity. Use Chart
7 of HEC 12 to determine the splash-over velocity. Figures 8 through 13 of HEC 12 show
the dimensions of the grates in Chart 7. If the grate dimensions do not match one of the
grates shown on Chart 7, then the reticuline grate usually will provide a conservative
assumption for the splash-over velocity.

Determine the velocity in the gutter:

S X T 2 0.028.062
Flow Area    0.650 ft.
2 2

Gutter Velocity  Q  1.65  2.53 fps


A 0.65

Chapter 5: Bridge Hydraulics 5-156


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

The splash-over velocity is estimated conservatively as 2.4 fps from Chart 7, HEC 12.
Using Equation 9 from HEC 12, the flow in width, W, that is intercepted can be
determined:

RF  1  0.09(V  V0 )  1  0.09(2.53  2.4)  0.988

where:
RF = Ratio of the frontal flow intercepted to the total frontal flow
V = Velocity of flow in the gutter, in fps
V0 = Splash-over velocity, in fps

The intercepted frontal flow is:

RF * QW = 0.988(1.52) = 1.50 cfs

The gutter flow that does not flow directly over the grate is called the side flow, QS. You
can determine the side flow by subtracting the frontal flow from the total gutter flow.

QS = Q – QW = 1.65 - 1.52 = 0.13 cfs

Momentum can carry the side flow past the inlet before all of the flow can turn into the
side of the inlet. The amount of flow that turns into the inlet and is intercepted can be
calculated using Equation 10 from HEC 12:

 0.15V 1.8   0.152.531.8 


RS  1 1  1
2.3 
1    0.0245
S L  0 .021 2.3 
 X   

RS is the ratio of the side flow intercepted to the total side flow. The intercepted side flow
is: RS * QS = 0.0245(0.13) = 0.00 cfs. Therefore, the total flow intercepted, which is the
sum of the frontal and side flows intercepted, is conservatively estimated as 1.50 cfs.

You also must check the capacity of the outlet pipe in the bottom of the scupper inlet. You
can do this using the orifice equation.

Q  CA2gh 
1
2

Chapter 5: Bridge Hydraulics 5-157


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

where:
C = Orifice coefficient = 0.6
A = Area of the orifice opening, in square feet
g= Gravitational force (32.17 ft/sec2)
h= Head on the orifice opening, in feet

Assuming that the orifice will not impact the intercepted flow unless the head is equal to
the distance from the outlet pipe opening to the top of the grate, D, the outlet pipe capacity
is:

D 2  (8 / 12) 2
A   0.349 ft 2
4 4

Q  0.60.349 232.17 7 / 12  2  1.28cfs


1

This flow is less than the capacity of the grate, and, therefore, the outlet pipe controls the
interception capacity of the inlet. The actual capacity of the outlet pipe will be slightly
greater because the actual head on the pipe will be slightly greater than the top of the
grate. However, this value is a conservative estimate of the intercepted flow.

Example 5.6-3
Constant Grade
Scupper flow on bridges with a constant grade will reach an equilibrium state if the bridge
is long enough. The equilibrium state occurs when the runoff from the area between
scuppers is equal to the flow intercepted by the scuppers.

The spread at scuppers prior to reaching equilibrium will be less than the equilibrium
spread. Therefore, equilibrium spread is a conservative estimate for scuppers on a
constant grade.

Determine the equilibrium spread for standard scuppers on a bridge with the following
characteristics:

 One of dual bridges for a six-lane divided roadway


 The deck has a constant 0.02 ft/ft cross slope
 The typical section has three 12-foot travel lanes, a 10-foot outside shoulder, and
a 6-foot inside shoulder. The barrier walls on each side are 1.5 feet wide. The total
deck width is 55 feet.

Chapter 5: Bridge Hydraulics 5-158


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

 The longitudinal grade is a constant 0.2 percent. (Normally, the minimum gutter
grade of 0.3 percent also should be applied to a bridge with flow along its barrier
wall. However, older bridges with flatter slopes are sometimes widened rather than
replaced. Occasionally, even flat-grade bridges are widened.)

Solution:
Since clogging can be a problem for scuppers, it is common to assume that every other
scupper is clogged. This assumption doubles the length between functioning scuppers
from 10 feet to 20 feet. Using this assumption, the deck runoff generated between each
scupper is:

Q = CiA = (0.95)(4)[(55)(20)/43560)] = 0.096 cfs

If the bridge is long enough, the equilibrium flow intercepted by the last scupper also will
be equal to this flow rate. Using 0.096 cfs as the intercepted flow, you can use Figure 5.6-
4 to determine the bridge deck flow just upstream of a scupper. Entering the y-axis with
the equilibrium intercepted flow of 0.096, an equilibrium flow just upstream of the scupper
of 0.61 cfs is read from the x-axis.

The spread just upstream of the scupper is:

3 3
 Qn  8
 (0.61)(0.016)  8
Spread   5 / 3 1/ 2 
 5/3 1/ 2 
 8.1 ft.
 0.56 S X S   0.56(0.02) (0.002) 

This is the equilibrium spread. Since this value is less than 10 feet, the width of the
shoulder, the standard scuppers will be adequate for this bridge.

Usually, you will omit scuppers near the end of a bridge due to potential soil erosion near
the abutments. Add the runoff from this area and the approach slab to the bypass at the

Chapter 5: Bridge Hydraulics 5-159


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

last scupper and the combined Q used to check the spread at the end of the approach
slab.

Example 5.6-4
For this example, use the information for the bridge in Example 5.6-3, with the following
substitutions:

 Omit scuppers in the last 50 feet of the bridge.


 Use a 30-foot approach slab for the bridge.

Determine the spread at the end of the approach slab.

Solution:
If a bridge has scuppers continuously from the crest of the bridge, then a conservative
estimate of the bypass from the last scupper is the equilibrium bypass. From Example
5.6-3, the equilibrium bypass is:

0.61 cfs – 0.096 cfs = 0.51 cfs

Equilibrium bypass
Equilibrium scupper interception
Equilibrium flow just upstream of scupper

The runoff from the area between the last scupper and the end of the approach slab is:

Q = CiA = 0.95 (4) [(50 +30) 55 / 43560] = 0.38 cfs

Bridge width from Example 5.6-3

The total flow at the end of the approach slab can be conservatively estimated as:

QTotal = 0.51 + 0.38 = 0.89 cfs

The spread can be conservatively estimated as:

3 3
 Qn  8
 (0.89)(0.016)  8
Spread   5 / 3 1/ 2 
 5/3 1/ 2 
 9.3 ft.
 0.56 S X S   0.56(0.02) (0.002) 

Chapter 5: Bridge Hydraulics 5-160


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Since the spread is less than 10 feet, the scupper design is acceptable.

If this estimate exceeded the allowable spread, the bridge deck drainage design does not
necessarily need to be changed. The spread can be checked with a more accurate
approach that accounts for the flow at each scupper, as described in Section 5.6.4.

Example 5.6-5
Flat Grade
You can determine the capacity of a scupper on a bridge with 0-percent longitudinal grade
from the figure shown below:

Scupper Capacity in Sump Conditions

Using the bridge from Example 5.6-3, except with a 0-percent grade, determine if
standard scuppers are adequate.

Solution:
Assuming that every other scupper is clogged, each scupper would need to take the flow
from a strip of the bridge deck that is 20 feet wide. You determined the runoff from this
area in Example 5.6-3 to be 0.096 cfs. Entering the above figure with this discharge, the

Chapter 5: Bridge Hydraulics 5-161


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

scupper flow will be in the transitional range between weir and orifice flow. The flow
conditions are imprecise because of this transition. However, the depth of water above
the orifice can be conservatively estimated as 0.11 feet. The spread is:

Spread = depth / Sx = 0.11 / 0.02 = 5.5 feet

Since the spread is less than the width of the shoulder, which is 10 feet, standard
scuppers meet the criteria.

Vertical Curves
Vertical curves complicate the analysis of scupper interception and spacing. However,
you can check scuppers on crest curves at various locations by assuming the grade at
that location is a constant grade. This will be conservative for crest vertical curves, but
also can be overly conservative. Consider using a more detailed analysis procedure, as
described in Section 5.6.4, before using scupper spacing that deviates from the standard.

At the crest of a vertical curve, there is a point where the slope is zero, and—depending
on the length of the curve—there can be a significant portion where the slope is almost
flat. The flow depth in this area is not well represented by the gutter flow equation because
this equation is a normal depth equation. The flow at the crest will not be at normal depth
because it will be experiencing a drawdown due to the combination of steeper slopes and
scupper interception downhill. Checking the spread near the crest with the gutter flow
equation will be conservative. For slopes less than 0.002 ft/ft, check the spread with the
flat grade assumptions if the spread criteria is violated using the gutter flow equation. This
is true for both the equilibrium analysis of this section and the more detailed analysis of
Section 5.6.4.

Avoid sag vertical curves. If this is not possible, then use the more detailed analysis
procedure described in Section 5.6.4.

Chapter 5: Bridge Hydraulics 5-162


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Example 5.6-6
Use the bridge from Example 5.6-3, except with the following roadway profile information:

Length of V.C = 600’


101+40
Station

107+40
Station
+ 0.5%
Begin Bridge

- 0.75%
Begin Appr.

End Bridge

End Appr.
100+00

100+30

107+30

107+60
The ground beneath the bridge is less than 25 feet below the bottom of the bridge deck
for a distance of 50 feet from each bridge end. Determine the required deck drainage
features.

Solution:
Determine the location of the high point on the bridge:

XHIGH POINT = (g1 x L) / (g2 – g1)


= (0.005 x 600) / (0.0075 – 0.005)
= 240 feet

Therefore, the high point is located at Station 103+80. The drainage area at the edge of
the approach slab at Station 100+00 is:

Area = (55) (380) / 43560 = 0.48 acres

The flow is:

Q = CiA = 0.95 (4) (0.48) = 1.82 cfs

Chapter 5: Bridge Hydraulics 5-163


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

where:
C = Rational runoff coefficient
i= Rainfall intensity, inches per hour
(Refer to Chapter 6 for explanation to use 4 in/hr)
A = Drainage area, acres
Solving the gutter flow equation for spread:

3 3
 Qn  8
 (1.82)(0.016)  8
Spread   5 / 3 1/ 2 
 5/3 1/ 2 
 10.3 ft.
 0.56S X S   0.56(0.02) (0.005) 

The spread exceeds the allowable spread of 10 feet. Minor changes to the roadway and
bridge profile would reduce the spread to an acceptable amount, which is less than 10
feet. However, after discussions with the roadway and the bridge engineers, if you cannot
adjust the roadway grade, then consider using standard scuppers. For this example, we
will assume the roadway grade cannot be adjusted.

The drainage area and flow are the same at the other bridge end at Station 107+60. The
spread is:

3 3
 Qn  8
 (1.82)(0.016)  8
Spread   5 / 3 1/ 2 
 5/3 1/ 2 
 9.5 ft.
 0.56 S X S   0.56(0.02) (0.0075) 

Since this spread is less than 10 feet, scuppers are not needed from the high point of the
bridge at Station 103+80 to the bridge end at Station 107+30.

Omitting scuppers within 50 feet of the bridge end, place standard scuppers every 10 feet
starting at Station 100+80 and ending at Station 103+70. The next step is to determine if
this design meets spread criteria. The previous examples show this design will work:

 Example 5.6-5 shows that standard scuppers on this bridge will meet the spread
criteria for flat grades. Therefore, scuppers at the top of the vertical curve where
the longitudinal slope is less than 0.002 ft/ft will meet the spread criteria.
 Example 5.6-3 shows that standard scuppers on this bridge will meet the spread
criteria for grades equal to or greater than 0.002 ft/ft.
 Example 5.6-4 shows that the spread at the end of the approach slab also will
meet the spread criteria.

Chapter 5: Bridge Hydraulics 5-164


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Therefore, the deck drainage design for this bridge is standard scuppers starting at
Station 100+80 and ending at Station 103+70.

The evaluation above uses simplified, but conservative, assumptions of equilibrium flow.
If the design failed to meet criteria under the conservative assumptions, then you can
perform a more-detailed analysis to evaluate the design. The following will illustrate the
detailed analysis procedure and explain how a spreadsheet can be used to automate the
analysis.

Enter the values of the cells in Row 1 through Row 8 of the spreadsheet as shown; i.e.,
none of these cells have formulae.

Although the scupper spacing is 10 feet, the spacing was entered as 20 feet to
conservatively assume that every other scupper was clogged.

The vertical curve data are not entered in the same manner as listed on the profile sheets
in the Plans or in Geopak. For the formulation in this spreadsheet, the peak of the vertical
curve must be determined, and all distances referenced from the peak. The slopes must
be entered so that the calculated slopes always have a positive value. G1 should be the
slope at the uphill end, and G2 the slope at the downhill end.

Chapter 5: Bridge Hydraulics 5-165


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

The remaining rows will have formulae in some of the cells.

In Row 9, enter the following formulae in each column:

Column A: =A8+$B$1
Column B: =(A9-A8)*$B$4/43560
Column C: =G8+0.95*4*B9
Column D: =($E$3-$E$2)*A9/$E$4+$E$2
Column E: =(C9*$B$3/0.56/$B$2^(5/3)/D9^0.5)^(3/8)
Column F: =IF(D9<0.002,J9,(IF(D9>0.005,K9,(J9+(K9-J9)*(D9-0.002)/0.003))))
Column G: =C9-F9
Column J: =IF(C9>1,Chart!$B$15,PERCENTILE(Chart!$B$4:$B$15,
PERCENTRANK(Chart!$A$4:$A$15,C9,20)))
Column K: =IF(C9>1,Chart!$E$15,PERCENTILE(Chart!$E$4:$E$15,
PERCENTRANK(Chart!$D$4:$D$15,C9,20)))

Column A keeps track of the distance from the upstream end.

Column B determines the drainage area between the current scupper and the previous
scupper uphill. This spreadsheet assumes that the bridge has a constant width along the
length of the bridge being analyzed.

Column C determines the flow immediately upstream of the current scupper using the
Rational Equation. The rainfall intensity is assumed to be four inches per hour and the
Runoff Coefficient is assumed to be 0.95. The bypass from the previous scupper is
combined with the runoff from the area between the scuppers.

Chapter 5: Bridge Hydraulics 5-166


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Column D determines the slope of the profile grade at the current scupper.

Column E determines the spread using the gutter flow equation.

Column F determines the intercepted flow rate based on Figure 5.6-4. If the slope is less
than 0.002, the curve labeled “0.2%” is used. If the slope is greater than 0.005, the curve
labeled “0.5, 1, 2%” is used. If the slope is between 0.002 ft/ft and 0.005 ft/ft, a value is
interpolated between the two curves. Values for these two curves are determined in
Column J and Column K.

Column G determines the scupper bypass flow.

Column J and Column K read the flows for the two curves of Figure 5.6-4. In the
formulation of this spreadsheet, the curves are represented on another sheet named
“Chart.” The values for the chart are presented on the next page.

At the end of the vertical curve (or, in this case, at the Begin Vertical Curve Station, since
the flow is in the opposite direction of the stationing), the profile grade slope becomes a
constant value. The formula in Column D is changed to the constant of 0.005 ft/ft, as
shown below.

The last scupper is at Station 100+80, which is 300 feet from the crest. The final row, Row
24, checks the spread at the edge of the approach slab. Since the spread at each scupper
and at the edge of the approach slab is less than the shoulder width of 10 feet, the design
meets the spread criteria.

As noted above, a separate sheet named “Chart” is included to represent the two curves
in Figure 5.6-4. The values entered on “Chart” are shown below:

Chapter 5: Bridge Hydraulics 5-167


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.6.4 Closed Collection Systems (Option 3)


The third option is a closed system. You will need to use a closed system if:

 The spread criteria is exceeded without scuppers or inlets on the bridge


 The deck drainage cannot be allowed to freefall to the area below the bridge
 The roadway profile or shoulder width cannot be adjusted

Use grated inlets in closed systems to minimize debris in the piping system. Refer back
to Section 5.6.3 for guidance on determining the interception capacity of grated inlets.
You will need to coordinate the dimensions and locations of the inlets with the structural
designer. Analyze the above-deck design (i.e., size and location of the grated inlets) using
a more detailed procedure rather than the equilibrium assumptions from the previous
sections. Table 5.6-1 illustrates a typical procedure.

Chapter 5: Bridge Hydraulics 5-168


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Table 5.6-1: Typical Inlet Location Analysis

Inlet Location Drainage Area Discharge Spread Bypass

Station 1
Station 2
...

Station n

Station 1: The first inlet downhill of the crest


Drainage Area: The area between the inlet and the crest for the first inlet; for
subsequent inlets, the area uphill to the previous inlet
Discharge: The sum of the discharge from the drainage area plus the bypass
from the previous inlet
Spread: Calculated using the gutter flow equation or the flat area
assumptions
Bypass: Determined by the inlet or scupper capacity

The below-deck system will have a network of pipes to convey the discharge collected by
the inlets to an outlet location. There are two types of systems. One type discharges
downward at the piers or bents. You will find this type of system more commonly for
overpasses. Typically, you will locate the inlets near the pier, so there are few horizontal
segments of pipe and flow is not combined from multiple inlets. Therefore, the controlling
point hydraulically typically will be the entrance to the piping system at the inlet.

The other type of system discharges at the bridge ends. The system will require
longitudinal pipes along the bridge that will carry the combined flow of multiple inlets.
Design the below-deck piping system using a procedure similar to the procedure in
Chapter 6 of this document. The procedure may be modified to use the driver visibility-
limiting rainfall intensity of four inches per hour.

Beside the hydraulic capacity of the piping system, the layout of the system also should
consider:

 Minimum cleaning velocities—Three feet per second is recommended.


 Cleanout locations—The locations should consider both access to all segments of
the pipe system and access to the cleanout by maintenance personnel.

Chapter 5: Bridge Hydraulics 5-169


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

 Underdeck closed drainage system—Design the system to minimize sharp bends,


corner joints, junctions, etc. These features occasionally reduce the hydraulic
capacity of the system but, more importantly, they provide opportunities for debris
to snag and collect. Use Y-connections and bends for collector pipes and
downspouts to help prevent clogging in mid-system.
 UV resistance—Pipes should be UV resistant. If they are not, then locate pipes to
prevent UV exposure. Tucking the pipe system behind the bridge beams will
prevent UV exposure.

You can find optional material for bridge collection pipes in Chapter 22 of the Structures
Detailing Manual. No matter what type of pipe is used, give attention to the design of a
hanger system, which the bridge design engineer should design, or you can design in
coordination with the bridge design engineer. If the collection system is connected to a
roadway structure, you may need to call for a resilient connector. For proper design, it is
critical that you coordinate with the structures engineer.

5.7 BRIDGE HYDRAULICS REPORT FORMAT AND


DOCUMENTATION
Section 4.11.2 of the Drainage Manual lists the minimum information that you must
include in the BHR. The minimum requirements are broken down for:

 Bridge and bridge culvert widening


 Bridge culverts
 Category 1 and 2 bridges

The introduction to Section 4.11.2 has a concise set of rules to guide production of all
sections in the BHR. Reviewing this brief paragraph before compiling the documentation
can help focus the BHR. Additional general guidance to follow while preparing the BHR
is:

 Present the BHR in clear and concise language, without redundant information or
unsubstantiated comments.
 Make sure graphics address the technical aspects of the project with the public’s
point of view in mind.
 Use a consistent report format, as well as consistent units with alternative units
presented where appropriate.

Chapter 5: Bridge Hydraulics 5-170


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.7.1 Bridge Hydraulics Report Preparation


Although the level of detail will vary depending on the type of work (i.e., bridge widening,
bridge replacement, or a new bridge crossing), the complexity of the hydrology and
hydraulics of the site, and the regulatory requirements, the following general chapter
outline is sufficient for most reports:

 Executive Summary
 Introduction
 FEMA/Regulatory Requirements
 Hydrology
 Hydraulics
 Scour
 Deck Drainage
 Appendices

The required documentation can be organized into this suggested outline.

5.7.1.1 Executive Summary


The Executive Summary should be a concise statement of findings. Describe the existing
and proposed bridges. Include a summary of all design recommendations for the
proposed bridge crossing (Items 1-10 for Category 1 and 2 bridges from Section 4.11.2.4
of the Drainage Manual).

The objective of the Executive Summary is to provide the findings in an opening statement
so that when the reviewer assesses the report in the future, the reviewer would
immediately understand the reasons for choosing the particular bridge. Include a brief
conclusion recounting why you selected the proposed bridge length. The discussion
should include other bridge considerations that were pertinent or had an important
influence on this project. (For bridge widening, this discussion is not necessary.) The
important influences might include the following:

 Costs
 Maintenance of traffic
 Roadway geometrics that affect bridge length
 Hydrology

Chapter 5: Bridge Hydraulics 5-171


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

 Hydraulics
 Scour
 Stream geomorphology
 Constructability
 Environmental concerns
 Wildlife shelf requirements
 Other unique concerns particular to the site

Include a discussion of any variations from policies in the Drainage Manual, FDM, or
Structures Manual.

5.7.1.2 Introduction
The introduction should describe the location of the bridge briefly, including the name of
the water body being crossed. Giving the latitude and longitude and/or the township,
range, and section will enhance the location description. Include a figure showing a
location map.

Describe the waterway and floodplain at the proposed crossing. Describe the existing
crossing, if any, including the bridge, relief bridges, and roadway embankment within the
floodplain. The description of bridges should include only details that affect the hydraulics:

 Bridge length
 Span lengths
 Foundation type and sizes
 Low member elevations
 Deck and beam heights
 Other details that affect the hydraulics

Also, describe the purpose of the project (widening, replacement, etc.).

Describe the land use in the area potentially affected by backwater from the crossing.
Discuss any nearby buildings or other structures that potentially will control the allowable
backwater from the crossing.

State the date of the site visit, and include photographs as figures.

Chapter 5: Bridge Hydraulics 5-172


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Describe any pertinent information from the Bridge Inspection Report (BIR), and consider
including a copy of the report in an appendix. Discuss any information obtained from
contact with Department Maintenance.

State the associated datums for each data source and provide datum conversions needed
to convert elevations between differing datums.

5.7.1.3 Floodplain Requirements


Discuss requirements of FEMA and other regulatory agencies (Section 5.1.2) that may
influence the design of the crossing. Consider including an appendix with the
correspondence, meeting minutes, phone notes, etc. from coordination efforts with the
agencies. If the original FEMA model was obtained, include a copy in the appendix.

5.7.1.4 Hydrology
Discuss the methods used to determine and check the flow rates applied in the analysis.
Include a summary table of frequencies and discharges used in the final analysis.

The hydrologic calculations, computer input and output, or documentation obtained from
others used to establish the design flow rates should be included in an appendix.

5.7.1.5 Hydraulics
One-Dimensional Model Setup
Identify and briefly describe the computer program used to calculate the water surface
elevations. Include a figure showing the location of the cross sections used in one-
dimensional models. Figures 5.7-1(1) and 5.7-2 are examples of cross section location
figures. Describe the following aspects of the model development:

 How the data for all the cross sections were obtained and how cross section
locations were selected
 How the starting water surface elevations (tailwater conditions) were determined
 How the Manning’s roughness coefficients were selected

Chapter 5: Bridge Hydraulics 5-173


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.7-1(1): Example Cross Section Location Figure on an Aerial

Figure 5.7-2: Example Cross Section Location Figure on a Quadrangle Map

Chapter 5: Bridge Hydraulics 5-174


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

If warning messages remain in the final output, describe any attempts to eliminate the
warnings and the reasoning for not resolving them. Input and output from the computer
programs used to analyze the crossing should be included in the appendixes. Electronic
copies of the input files also will be provided to the Department.

In some cases, such as bridge widenings that do not affect the water surface profiles,
calculations may not be performed. However, you must still include the flood data at the
site in the plans, per FHWA requirements. If you do not calculate the flood data, then you
must obtain them from another source. Typical sources that can be used are hydraulic
reports for the existing crossing or FEMA Flood Insurance Studies. Document the source
in the report.

Compare water surface elevations for the existing and proposed alternative bridges. The
location of the approach section may vary between the existing bridge and each of the
alternative bridges. For the comparison to be valid, perform the water surface elevation
comparisons at a section that is at a common location in each model. As illustrated in
Figure 5.7-3, make the comparison at the location of the approach section that is farthest
upstream.
Approach
Section

Section

Alternative 1
Exit

Flow
Comparison

Approach
Section

Section

Section

Alternative 2 Flow
Exit

Figure 5.7-3: Water Surface Elevation Comparisons

Include a table that summarizes the water surface elevations for the existing and
alternative bridges. Table 5.7-1 is an example of a table comparing water surface
elevations.

Chapter 5: Bridge Hydraulics 5-175


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Table 5.7-1: Example Water Surface Elevation Comparison


50-Year 100-Year 500-Year
Elevation Elevation Elevation
Existing Conditions 57.4 57.8 59.0
Proposed Conditions 57.2 57.8 59.1
Elevations are NGVD 1929. Elevations shown on the BHRS in the Appendix have been converted to NAVD 88. The
elevations are adjusted by subtracting 0.65 feet.

Two-Dimensional Model Setup and Results


If two-dimensional modeling was performed as part of the hydraulic analysis of the bridge,
the BHR should contain sufficient documentation of the model development and
simulation to provide the reviewer and subsequent readers of the report a clear
understanding of both the modeling process and the results of the modeling. This begins
with a description of the model selected and justification for that selection. The report
should document who or what agency developed the model (e.g., FHWA’s FESWMS
model), as well as the features of either the model or the physical features of the study
area that make the model the appropriate choice.

Documentation of the model development should include the following:

 A description of the survey data employed (including horizontal and vertical


datums)
 A description of the boundary conditions, as well as sufficient documentation of
their development
 Documentation of the selected friction specification
 A listing of other model input parameters (e.g., turbulent closure parameters, time
step size, etc.)
 Graphic representations of the model mesh clearly displaying both elevation
contours and elements (e.g., Figure 5.7-4 through Figure 5.7-6). Figures should
display both the model domain as well as a close-up of the bridge location to
ensure documentation of the resolution of the study area.

Chapter 5: Bridge Hydraulics 5-176


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.7-4: Tampa Bay Model Mesh Domain

Chapter 5: Bridge Hydraulics 5-177


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.7-5: Model Mesh at Tampa Bay

Chapter 5: Bridge Hydraulics 5-178


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.7-6: Model Mesh at the Courtney Campbell Causeway Bridge

Documentation of the two-dimensional model should include:


 A complete description of the calibration process
o Calibration data
o The simulations
o Parameters changed to achieve calibration
o Parameters of the model
 Both a qualitative and quantitative description of the model’s capability to predict
measured data
o Calculation of mean error
o Standard deviation
o Percentage error, etc. over time series, between observed high water
marks, measured stages, or comparison with predicted tidal ranges.

Chapter 5: Bridge Hydraulics 5-179


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Examples of qualitative descriptions are provided in Figure 5.7-7 and Figure 5.7-8, which
show comparisons between measured and modeled water surface elevations and flow
rates.

FayTropical Storm
Tropical Storm

Fay

GustavHurricane

IkeHurricane
Hurricane
Hurricane

Gustav

Ike
Figure 5.7-7: Model Calibration Plot for the US 90 Bridge over Macavis Bayou
Replacement Project at the River Run Marina

Chapter 5: Bridge Hydraulics 5-180


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

80000

60000

40000
Flow Rate (cfs)

20000

Measured
0
Modeled
-20000

-40000

-60000

-80000
4/5/2001 9:00

4/5/2001 12:00

4/5/2001 15:00

4/5/2001 18:00

4/5/2001 21:00
Date Time

Figure 5.7-8: Flow Rate Calibration at Lake Worth Inlet


(Error Bars Indicate 10% Error)

Chapter 5: Bridge Hydraulics 5-181


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Documentation of two-dimensional modeling simulation results should include, at a


minimum:
 Table of max conditions for each simulation at the bridge
 Figures of each simulation (Figure 5.7-9):
o Display contours of velocity magnitude
o Velocity vectors displaying the direction of the flow across bridge
 For long bridges, hydraulic parameters at each pier or groups of piers should list:
o Max stage
o Max flow rate
o Max velocity
o Angle of attack
 Tidal analysis (time-dependent simulation)
o Time series plot of design values for stage, velocity, and flow rate (Figure
5.7-10 through Figure 5.7-12)

Chapter 5: Bridge Hydraulics 5-182


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

Figure 5.7-9: Velocity Magnitude Contours and Velocity Vectors at the Time of
Maximum Velocity during the 100-Year Storm Surge Event at the SR-A1A Bridge
over the Loxahatchee River

Chapter 5: Bridge Hydraulics 5-183


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

50000

40000

30000

20000

10000
Flow Rate (cfs)

-10000

-20000

-30000

-40000

-50000
0 10 20 30 40 50 60
Simulation Time (Hrs.)

100-year 500-year

Figure 5.7-10: Flow Rate Time Series during the Design and Check Event at the
SR-A1A Bridge over the Loxahatchee River

Chapter 5: Bridge Hydraulics 5-184


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

12

10

8
Water Surface Elevation (ft-NAVD88)

-2

-4
15 20 25 30 35 40 45 50 55 60
Simulation Time (Hrs.)

100-year 500-year

Figure 5.7-11: Water Surface Elevation Time Series during the Design and Check
Event at the SR-A1A Bridge over the Loxahatchee River

Chapter 5: Bridge Hydraulics 5-185


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

4
Velocity Magnitude (ft/s)

0
0 10 20 30 40 50 60
Simulation Time (Hrs.)

100-year 500-year

Figure 5.7-12: Velocity Magnitude Time Series during the Design and Check Event
at the SR-A1A Bridge over the Loxahatchee River

Required documentation of two-dimensional wave modeling is almost identical to that for


hydraulic analyses. The only difference is in the parameters themselves. At a minimum,
the wave parameters should include the highest significant wave height at the bridge
cross section, the associated peak period, the maximum wave height, and the maximum
crest elevation with all parameters associated with the 100-year return period conditions.

Alternatives Analysis
You will not need this section for bridge-widening projects. For new and replacement
bridges, this section should document the cost analysis, environmental impacts, and other
impacts on adjacent properties. Each alternative still should meet the design standards,
but if exceptions must be made for an alternative, then the exception should be included
in the comparisons. This section must document the reasons for selecting the
recommended alternative.

Chapter 5: Bridge Hydraulics 5-186


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.7.1.6 Scour
You should plan to include a discussion of the stream geomorphology, the scour history,
the long-term aggradation or degradation, and the scour values, including information on
the methods used to determine each of these items. Plot scour depths in a figure.

Discuss the proposed abutment protection. If you use one of the standard abutment
protection designs given in Section 4.9.3.2 of the Drainage Manual, abutment scour need
not be calculated and plotted. You may use other abutment protection designs in certain
circumstances, but not without prior approval from the District Drainage Office.

5.7.1.7 Deck Drainage


Document the proposed method of deck drainage. Justify the use of longitudinal collection
systems. Include in the appendix spread and interception calculations, as well as capacity
calculations for any longitudinal collection systems.

5.7.1.8 Appendices
Include calculations and other backup documentation as appendixes to the BHR to avoid
disrupting the flow of the main body of the report. Items to consider including in the
appendixes are:

 Hydrology calculations
 Hydrology reports from other sources
 Hydraulic calculations
 Hydraulic reports from other sources
 FEMA report excerpts and maps
 Scour computations
 Cost calculations for alternatives
 Deck drainage calculations
 Regulatory requirements and permits
 Memos, meeting minutes, and phone notes

5.7.2 Bridge Hydraulics Report Process


Figure 125.5.1 of the FDM, gives the Approval and Concurrence Process for the Bridge
Hydraulics Report. FDM 250 specifies the multidisciplinary approach to follow for scour
consideration, along with submittal requirements. Prepare the BHR in conjunction with

Chapter 5: Bridge Hydraulics 5-187


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

the Bridge Development Report and preliminary Structures Plans. Figure 250.2.1 of the
FDM outlines a flow chart for the Structural Plans Development Process.

The process flow chart in Figure 5.7-13 shows the general sequence of events necessary
to prepare a Bridge Hydraulics Report. You also may need to perform additional
coordination, especially for projects involving floodways or for other complex elements.

After you have a relatively good idea of the approximate structure length and location,
you should conduct a field review. Then, submit the preliminary structure length and
location, along with preliminary scour depths and low member elevations to the Structures
Design Office for their preliminary evaluation. After you have developed the proposed
bridge configuration and foundation type and submitted them back for review, perform the
final hydraulic and scour analyses and submit them to the Structures and Geotechnical
Departments.

Have the BHR and BHRS reviewed internally (or by an outside consultant, if necessary).
After you have addressed all comments, approve the BHR and BHRS and submit them
to the Department for concurrence. After the BHR and BHRS receive concurrence from
the Department, the final BHR and BHRS should be submitted to the structural and
geotechnical engineers so that they can complete the BDR and geotechnical reports.

Chapter 5: Bridge Hydraulics 5-188


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

The Hydraulics Engineer reviews


environmental documents for
commitments related to the bridge

The Hydraulics Engineer:


(1)Begins data collection
(2) Field reviews the project site to
define survey needs (with input from
structures)

The Hydraulics Engineer prepares Survey request to


request for survey and, if not available location surveyor
from PD&E, request for geotechnical (directly or through
information for scour calculations PM)

Hydraulics Engineer continues data


collection and starts BHR preparation
with hydrology calculations, conceptual
deck drainage, etc.
Surveys
performed:
*Cross Section
*Tidal
*Others, as
*Coordinate with Structures & Geotech requested
re: structure types, spans, substructure,
known constraints
*Start environmental coordination
(bridge length, deck drainage)
*Discuss MOT constraints with
Construction
2

Figure 5.7-13: Bridge Hydraulics Report Process

Chapter 5: Bridge Hydraulics 5-189


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

2
Hydraulics Engineer starts hydraulic
calcs to establish the minimum
hydraulic structure to avoid high Survey data to
Yes velocities and/or adverse upstream Hydraulics
effects and to establish the minimum Engineer
low member elevation for drift
clearance

Coordination among hydraulics,


structures, environmental, project
manager to develop the most
appropriate structure based on
hydraulic, environmental, structural,
and geotechnical criteria

If necessary, revise Complete deck


hydraulics based on drainage computations
structure size and type and coordinate with
selected in last step design and structures

Geotechnical
Finalize scour calculations
Information

Does scour
calculated dictate a
Yes
structure different
than assumed?

No

Figure 5.7-13: Bridge Hydraulics Report Process (continued)

Chapter 5: Bridge Hydraulics 5-190


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

* BHRS prepared
*Hydraulics Engineer completes BHR
and furnishes needed information to
Structures

* Structures prepares BDR


* BHR, BDR, geotech info are reviewed
for compatibility

Structures Plans Preparation

Final check to be sure final bridge


plans and BHR are still in agreement

Figure 5.7-13: Bridge Hydraulics Report Process (continued)

Chapter 5: Bridge Hydraulics 5-191


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

5.7.3 Common Review Comments


By far, the most frequent comments associated with the BHR and BHRS address
omissions or requests for supporting documentation. The following checklist should
provide an additional resource to ensure a quality product for submission to the
Department:

 Draft Bridge Hydraulics Report


o Verify that the report contains the following information:
 Bridge location
 Bridge number (if available)
 Florida County
 Description of all data collected in the office data collection
 Description of all data collected in the field data collection
 List of relevant datums (e.g., NAVD 88, NGVD 29, etc.); provide the
difference between datums if supporting documents, new data, and
the Plans use different datums
 Description of the model hydrology
 Description of the constructed hydraulic model
 Description of the modeling procedures (inputs, boundary conditions,
etc.)
 Quantitative and qualitative presentation of the calibration simulation
results
 Presentation of the simulation results
 Description of scour calculation procedures
 Aggradation/degradation calculation (methodology and results)
 Channel migration calculation results (methodology and results)
 Contraction scour mode and calculation results (inputs and output)
 Local scour calculations and results (inputs and output)
 Total design scour prediction; total check event scour prediction;
recognize that maximum scour for these events can occur at a flow
less that the associated return interval flow rate, i.e., if overtopping
occurs before either the total design scour or total check event scour
 Wave climate/wave modeling discussion
 Wave force calculation procedure and results (inputs and output)

Chapter 5: Bridge Hydraulics 5-192


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

 Abutment protection recommendations and calculations (inputs and


output)
 Deck drainage discussion
o Check the report for the following:
 Language is clear and concise
 Presentation graphics address the technical aspects of the project
with the public’s point of view in mind
 Report format is consistent
 Units are consistent, with alternative units presented where
appropriate
 Cross referencing of figures, tables, section numbers within the
document have been double-checked
 Draft Bridge Hydraulics Recommendations Sheet
o Verify that the BHRS contains the following information:
 Plan View
 Stationing, scale, and north arrow; include the channel
baseline if one was created
 Existing topography (including existing bridge) and contours
(show elevations)
 The name of the water body
 Arrows showing the direction of the flow
 Proposed bridge begin and end station
 Limits and type of abutment protection
 Right-of-way lines
 Profile View
 Stationing and scale
 Existing surveyed cross section
 Road profile for the proposed structure with stationing and
elevations
 Proposed bridge with begin and end station, low member, and
pier locations
 Abutment locations (toe of slope) and abutment protection
 Design flood elevation

Chapter 5: Bridge Hydraulics 5-193


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

 Normal High Water/Mean High Water


 New bridge number
 Drainage Map and Location Map
 Location map with north arrow
 Range and township and an arrow showing the project
location
 Entire drainage area for the proposed structure
 Calculated drainage area
 Water elevations on date of survey
 Existing Structures, Hydraulic Design Data, and Hydraulic
Recommendations
 Existing structures
 Proposed structure
 Foundation
 Overall length
 Span length
 Type of construction
 Area of opening
 Bridge width
 Elevation of low member
 Hydraulic Information
 Normal High Water (non-tidal)
 Control (non-tidal)
 Mean High Water (tidal)
 Mean Low Water (tidal)
 Maximum event of record
 Design flood information
 Base flood hydraulic and scour information
 Overtopping flood/greatest flood hydraulic and scour
information
 Begin bridge station

Chapter 5: Bridge Hydraulics 5-194


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

 End bridge station


 Skew angle
 Navigation clearances—required and provided
 Drift clearances—required and provided
 Abutment protection description—begin and end bridge
 Deck drainage
 Remarks
 Final Bridge Hydraulics Report
o Verify that the report contains the following information:
 Changes to the report as specified by the responses to comments
following the Department review process
 Final Bridge Hydraulics Recommendations Sheet
o Verify that the BHRS contains the following information:
 Changes to the BHRS as specified by the responses to comments
following the Department review process

5.7.4 Bridge Hydraulics Recommendations Sheet (BHRS)


The Bridge Hydraulics Recommendations Sheet (BHRS) provides a single reference that
summarizes the findings and recommendations of the hydraulic analysis. The BHRS flood
data must match those given in the BHR and computer output.

The BHRS is divided into four sections:

 Plan View
 Profile View
 Location Map and Drainage Area
 Existing Structures, Hydraulic Design Data, and Hydraulic Recommendations
FDM 305 gives the minimum requirements of the first three sections. In addition, consider
the following items:

 In the Plan View, the FDM requires that the limits of riprap be shown. However,
abutment protection other than riprap may be proposed. Show the horizontal
extents and label the protection type in either the plan or profile view.

Chapter 5: Bridge Hydraulics 5-195


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

 Plot and label the profile of the existing natural ground in the Profile View, and note
the existing elevation at each end.
 When practical, you should show the profile of the expected design scour
(contraction and long-term scour along the entire unprotected cross section and
the local scour at the intermediate piers/bents). Display local scour holes as
beginning at the foundation element edges at the design scour depth and
extending up at a 1V:2H slope to meet the profile, illustrating the contraction/long-
term scour profile.
 Although the profile grade line must be plotted in the Profile View, you do not need
to show percent of grade. Plot the PC, PI, and PT of vertical curves using their
respective standard symbols; however, there is no need to note data (station,
elevation, length of curve). Flag begin and end bridge stations.

Figure 5.7-14 shows a larger view of the section of the BHRS that includes Existing
Structures, Hydraulic Design Data, and Hydraulic Recommendations. The hydraulic
design data and hydraulic recommendations are for the proposed structure. The required
data are identified by bold numbers in parentheses and a brief description is provided on
the following pages.

Chapter 5: Bridge Hydraulics 5-196


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

(50)

Figure 5.7-14: BHRS Required Data

(1) Existing Structures: Structure 1 refers to the structure being replaced or modified.
Structures 2, 3, and 4 refer to relief structures, immediate upstream and
downstream structures, and those structures that affect the hydraulics of the
proposed structure.

(2) Proposed Structure: This column should have information pertaining to the
proposed structure.

(3) Foundation: This row should have information describing the type of foundation
(e.g., timber piles, concrete piles, etc.).

Chapter 5: Bridge Hydraulics 5-197


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

(4) Overall Length (ft): This row should give the total length of the structure in feet.
The length should be measured from the top of the abutments. For the proposed
structure, this length should match the total length shown in the final plans.

(5) Span Length (ft): This row should give the span length of the structure in feet.
This length should be based on the length at the main span.

(6) Type of Construction: This row should have information describing the
material(s) used for construction of the structure (e.g., steel, concrete, steel and
concrete, etc.).

(7) Area of Opening (ft2) @ D.F.: This row should have the area of opening in square
feet below the design flood elevation less the assumed pile area, if significant, at
the bridge section.

(8) Bridge Width (ft): The bridge width should be from rail to rail, including the rails,
in feet.

(9) Elev. Low Member (ft): This elevation in feet should be the lowest point along the
low member of the structure.

(10) N.H.W. (Non-Tidal) (ft): The Normal High Water at the bridge. This water surface
elevation in feet only applies to non-tidal areas.

(11) Control (Non-Tidal) (ft): The water surface elevation in feet controlled by the
operation of pump stations, dams, or other hydraulic structures.

(12) M.H.W. (Tidal) (ft): The Mean High Water elevation in feet at the bridge. This
water surface elevation only applies to tidal areas.

(13) M.L.W. (Tidal) (ft): The Mean Low Water elevation in feet at the bridge. This water
surface elevation only applies to tidal areas.

(14) Max. Event of Record: This column provides information related to the maximum
event recorded based on historical information (if available).

(15) Design Flood: This column provides information related to the design flood.

(16) Base Flood: This column provides information related to the base flood.

Chapter 5: Bridge Hydraulics 5-198


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

(17) Overtopping Flood/Greatest Flood: If the overtopping flood has a lower return
period than the greatest flood, then the block indicating overtopping flood is
checked and the information related to the overtopping flood is shown.
Otherwise, the block indicating greatest flood is checked and the information
related to the greatest flood is shown.

(18) Stage Elev. NAVD 88 or NGVD 29 (ft): For freshwater flow, the elevation in feet
typically taken from the hydraulic model at the approach section for the design
flood and/or base flood, overtopping flood, greatest flood. Proper engineering
judgment is required for long bridges since it may not be realistic to use the
elevation at the approach section because the losses between the bridge and
approach section are large.

For tidal flow, the maximum elevation during the flood or ebb storm surge at the
bridge for the design flood and/or base flood, overtopping flood, greatest flood.
Add a remark that stage, discharge, and the velocity described in the flood data
do not occur at the same time.

(19) Discharge (cfs): For freshwater flow, the total discharge in cubic feet per second
used in the simulations for the design flood, base flood, overtopping flood, and/or
greatest flood.

For tidal flow, the maximum discharge during the flood or ebb storm surge at the
bridge for the design flood, base flood, overtopping flood and/or greatest flood.
Add a remark that stage, discharge, and the velocity described in the flood data
do not occur at the same time.

(20) Average Velocity (fps): For freshwater flow, the average velocity in feet per
second taken from the computer simulations at the Bridge Section for the design
flood, base flood, overtopping flood and/or greatest flood.

For tidal flow, the maximum average velocity at the bridge section during the
flood or ebb storm surge for the design flood, base flood, overtopping flood and/or
greatest flood.

(21) Exceedance Prob. (%): The probability that the conditions are exceeded.
Determined as 100% times unity over the return interval (e.g., 100%*(1/100) =
1%).

(22) Frequency (yr): The return period of the conditions in years.

Chapter 5: Bridge Hydraulics 5-199


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

(23) Frequency (yr): The frequency (return period) in years of the worst case scour
condition up through the design return period flow conditions.

(24) Frequency (yr): The frequency (return period) in years of the worst case scour
condition up through the design check period flow conditions.

(25) Pier No.: The pier number or range of pier numbers that correspond to the pier
size and type in Column 26 and the scour elevations in Columns 27, 28, and 29.

(26) Pier Size and Type: The proposed pier size and type that produces the greatest
scour. If necessary for clarity, place a reference to the appropriate details of the
bridge plans. If the space provided is not adequate, place the information in the
plan or profile view.

(27) Long-Term Scour (ft): Applicable only to structures required to meet extreme
event vessel collision load. See Section 6.2 for the definition of long-term scour.
If it is not applicable, state so.

(28) Total Scour Elevation (< 100-year) (ft): The predicted total scour elevation in feet
for the worst-case scour condition up through the scour design flood frequency.
This includes aggradation or degradation, channel migration, local scour (pier
and abutment), and contraction scour.

(29) Total Scour Elevation (< 500-year) (ft): The predicted total scour elevation in feet
for the worst-case scour condition up through the scour design check flood
frequency. This includes aggradation or degradation, channel migration, local
scour (pier and abutment), and contraction scour.

(30) Begin Bridge Station: The station for the beginning of the bridge.

(31) End Bridge Station: The station for the end of the bridge.

(32) Skew Angle (degrees): The angle in degrees at which the structure is skewed
from the centerline of construction. See Standard Plans, Index 400-289, Sheet
1, Schematic “B” for further explanation.

(33) Navigation Clearance (Horiz.) (ft): The actual horizontal navigation clearance in
feet provided between fenders or piers.

(34) Navigation Clearance (Vert.) (ft): The actual vertical navigational clearance in
feet provided between fenders or piers.

Chapter 5: Bridge Hydraulics 5-200


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

(35) Navigation Clearance (Above El.) (ft): For freshwater flow, the elevation (NAVD
88 or NGVD 29, ft) at the normal high water (NHW) elevation or control elevation.

For tidal flow, this is the elevation at mean high water (MHW).

(36) Drift Clearance (Horiz.) (ft): The actual minimum horizontal clearance in feet
provided.

(37) Drift Clearance (Vert.) (ft): The actual minimum vertical clearance in feet provided
above the design flood.

(38) Drift Clearance (Above El.) (ft): For freshwater flow, this is the design flood
elevation (NAVD 88 or NGVD 29, ft) and either of two values is appropriate. In
many cases, it is reasonable to use the elevation at the approach section,
realizing that this will be slightly higher than actual elevation at the bridge.

For tidal flow, use the maximum stage associated with an average velocity of 3.3
fps through the bridge section during the flood or ebb for the storm surge for the
design flood. If the maximum velocity due to the storm surge is less than 3.3 fps,
use the stage associated with the maximum velocity through the bridge section.
If either of these stages causes the profile to be higher than the profile of the
bridge approaches, consider other alternatives. One alternative is to discuss with
personnel in the Structures Design Office the potential of having less drift
clearance and designing the structure for debris loads. Another alternative is to
do a more rigorous and site-specific analysis to set the stage above which to
provide the standard drift clearance. Investigate and address these situations on
a site-specific basis.

(39) Navigation Clearance (Horiz.) (ft): The minimum horizontal navigation clearance
in feet required. See the FDM 210 for the minimum requirements. Other agencies
may have minimum clearance requirements.

(40) Navigation Clearance (Vert.) (ft): The minimum vertical navigation clearance in
feet required. The Department minimum clearances are given in the FDM 210.
Other agencies may have minimum clearance requirements.

(41) Drift Clearance (Horiz.) (ft): The minimum horizontal debris drift clearance in feet
required. The Department minimum clearances are given in the FDM 210.

Chapter 5: Bridge Hydraulics 5-201


January 2019

Drainage Design Guide


Chapter 5: Bridge Hydraulics

(42) Drift Clearance (Vert.) (ft): The minimum vertical debris drift clearance in feet
required above the design flood. The Department minimum clearances are given
in FDM 210.

(43) Rubble Grade: Grade of rubble (e.g., Riprap (Bank & Shore), etc.) to be
constructed at the begin and end bridge abutments. References can be made to
details sheets if non-standard riprap is employed.

(44) Slope: Slope of the abutments at the begin and end bridge (e.g., 1H:2V, etc.).

(45) Non-buried or Buried Horiz. Toe: Indicate whether the toe of the abutment will be
non-buried or buried when extended horizontally from the bridge. See Section
5.5.4 of this document for details.

(46) Toe Horizontal Distance (ft): Horizontal extent in feet of the rubble protection
measured from the toe of the abutment. See Section 5.5.4 of this document for
details.

(47) Limit of Protection (ft): Distance measured parallel to the stationing in feet, from
the edge of the rubble protection to the bridge begin/end station. If the distance
is different on each side, indicate both distances with their corresponding sides.

(48) Deck Drainage: Type of deck drainage to be used for the proposed structure
(e.g., scuppers, storm drain system, etc.)

(49) Remarks: This space is available to record any pertinent remarks.

(50) Wave Crest Elevation (ft): The 100-year design wave crest elevation in feet,
including the storm surge elevation and wind setup. The vertical clearance of the
superstructure must be a minimum of 1 foot above the wave crest elevation. The
Department minimum clearances are given in FDM 210.

Chapter 5: Bridge Hydraulics 5-202


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

CHAPTER 6: STORM DRAINS

6. STORM DRAINS .................................................................................................. 6-1

6.1 FDOT Storm Drain Tabulation Form .................................................................. 6-1

6.2 Hydrology ............................................................................................................ 6-6


6.2.1 Design Frequency .......................................................................................... 6-6
6.2.1.1 Storms of Greater Magnitude .................................................................. 6-6
6.2.2 Time of Concentration .................................................................................... 6-8
6.2.2.1 Peak Flow from Reduced Area ............................................................... 6-9
6.2.2.2 Ignoring Time of Flow in Section ........................................................... 6-12

6.3 Inlets and Pavement Hydraulics ...................................................................... 6-13


6.3.1 Inlets ............................................................................................................ 6-13
6.3.1.1 Apparent Locations ............................................................................... 6-13
6.3.1.2 Sags...................................................................................................... 6-13
6.3.1.3 Continuous Grades ............................................................................... 6-14
6.3.1.4 Back of Sidewalk .................................................................................. 6-14
6.3.1.5 Inlet Capacity ........................................................................................ 6-15
6.3.2 Pavement Hydraulics ................................................................................... 6-16
6.3.2.1 Gutter Grades ....................................................................................... 6-16
6.3.2.2 Cross Slope .......................................................................................... 6-19
6.3.2.3 Shoulder Gutter .................................................................................... 6-19
6.3.2.4 Determining the Spread ........................................................................ 6-21

6.4 Pipe System Placement .................................................................................... 6-33


6.4.1 Plan Layout .................................................................................................. 6-33
6.4.1.1 Retaining Wall Drainage ....................................................................... 6-34
6.4.2 Profile Placement ......................................................................................... 6-34
6.4.2.1 Slopes ................................................................................................... 6-34
6.4.2.2 Minimum Pipe Depth............................................................................. 6-36
6.4.3 Utility Coordination ....................................................................................... 6-37
6.4.4 Pipe-to-Structure Connections ..................................................................... 6-37

6.5 Pipe Hydraulics ................................................................................................. 6-47


6.5.1 Pressure Flow .............................................................................................. 6-47
6.5.2 Partially Full Flow ......................................................................................... 6-48
6.5.3 Lower-End Hydraulic Gradient ..................................................................... 6-48
6.5.3.1 Design Tailwater (DTW)........................................................................ 6-51
6.5.4 Upper-End Hydraulic Gradient ..................................................................... 6-51
6.5.5 Flow Velocity in the Pipe Section ................................................................. 6-53
6.5.6 Utility Conflict Box Losses ............................................................................ 6-55

Chapter 6: Storm Drains 6-i


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

6.5.7 Minor Losses................................................................................................ 6-56

6.6 Procedure .......................................................................................................... 6-57

Chapter 6: Storm Drains 6-ii


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

6. STORM DRAINS
6.1 FDOT STORM DRAIN TABULATION FORM
The primary means of documenting storm drain design is the Department’s storm drain
tabulation form shown in Figure 6.1-1. On this form, record items identified by numbers in
parentheses on Figure 6.1-1. These items are discussed in the description following the
form. This information also is available on the FDOT Drainage web site.

Chapter 6: Storm Drains 6-1


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

FLORIDA DEPARTMENT OF TRANSPORTATION


STORM DRAIN TABULATION FORM
Financial Project Identification: County: Network: Prepared: Date:
Description: Organization: State Road: Checked: Date:

ACTUAL VELOCITY
PIPE NOTES
DRAINAGE HYDRAULIC
STRUCTURE NO.
SIZE SLOPE (%) AND
LOCATION AREA (Acres) GRADIENT

TIME OF FLOW IN SECTION (min)


TIME OF CONCENTRATION (min)
(in.) REMARKS
OF

FULL FLOW CAPACITY (cfs) *


UPPER END
C= (1) ** CROWN ZONE: (38)

NUMBER OF BARRELS *
HYD. GRAD.

(fps)
INLET ELEVATION (ft.)
TYPE OF STRUCTURE
C= (1) ** FLOWLINE RISE FREQUENCY (Yrs): (39)

HGL CLEARANCE (ft.)


MINOR LOSSES (ft.) *
BASE FLOW (cfs) *

TOTAL FLOW (cfs)


ALIGNMENT NAME UPPER C= (1) ** PHYSICAL MANNING'S "n": (40)

INTENSITY (in/hr)

ELEVATION (ft.)

ELEVATION (ft.)
DISTANCE (ft.)

TAILWATER EL (ft): (41)

LOWER END
TOTAL (C*A)
LENGTH (ft.)

UPPER END
INCREMENT

SUB-TOTAL

PHYSICAL
VELOCITY
STATION LOWER SPAN MIN. PHYS.

TOTAL

FALL
(C*A)
SIDE

(fps)
(ft.)
(2) (6) (9) (10) (11) (21) (22) (27) (30) (32) (35)

(7) (8) (9) (10) (11) (12) (13) (14) (15) (16) (17) (18) (19) (20) (23) (24) (29) (33) (37) (42)
(3) (4) (5) (6) (28) (31) (36)
(9) (10) (11) (25) (26) (34)

* Denotes optional information.


** A composite runoff coefficient may be shown in lieu of individual C-values, provided the composite C calculations are included in the drainage documentation.

Figure 6.1-1: Storm Drain Tabulation Form

Chapter 6: Storm Drains 6-2


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Tabulation Form Description: through existing pipes. If so, note it in


the Remarks Column (42) or use the
1. Runoff Coefficients (C): You will be optional Base Flow Column.
limited to three runoff coefficients.
For most projects, this provides Manholes usually do not have
sufficient flexibility. incremental areas as they are
handling areas already tabulated. If
2. Alignment Name: The name of the the incremental drainage area does
alignment that the structure’s station not fit one of the three runoff
and offset references. coefficients selected, mathematically
adjust the size of the area to fit one of
3. Station: The survey station number the selected runoff coefficients. Note
for the structure being used. the adjustment in the Remarks
Column (42).
4. Distance (ft): The offset distance, in
feet, from reference point of the AreaADJ = (CACT/CSELECT) x AreaACT
structure to the reference station.
10. Total: The total area, in acres,
5. Side: The side, Right (Rt) or Left (Lt), associated with each runoff
of the reference station. coefficient and passing through the
structure. Identify all the areas that
6. Structure Number: The structure drain to the structure through pipes
number at the upper end is shown from upstream structures. Add these
above the structure number at the “upstream areas” to the incremental
lower end. Each major row (three drainage areas for the structure (9).
minor rows) of the form identifies an
inlet and the downstream pipe from 11. Sub-Total (C*A): The result of
that inlet. multiplying the total area associated
with each runoff coefficient (10) by
7. Type of Structure: Usually shown the corresponding runoff coefficient.
with abbreviations such as Type P-3
or P-5 for inlets; Type C or E for ditch 12. Time of Concentration (min): Usually,
bottom inlets (DBI); Type P-8 or J-7 the time required for the runoff to
(MH) for manholes; and Type J-7 travel from the most hydraulically
(Junct) for junction boxes. remote point of the area drained to
the point of the storm drain system
8. Length (ft): The length, in feet, from under consideration. This time
the hydraulic center of the structure consists of overland flow, gutter flow,
to the hydraulic center of the next and flow time within the pipe system.
downstream structure. Occasionally, this time is associated
with a reduced area that creates a
9. Increment: The incremental peak flow. If so, note it in the
drainage areas, in acres, Remarks Column (42). Show this
corresponding to the runoff number in minutes.
coefficients being used. It is
normally only the area that drains 13. Time of Flow in Section (min): The
overland to an inlet, but it can time, in minutes, it takes the runoff to
include areas that drain to structures pass through the section of pipe.

Chapter 6: Storm Drains 6-3


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

TSECT=Hydraulic Length (8)/Actual


Velocity (35) 20. HGL Clearance (ft): This value is
determined by calculating the
14. Intensity (in/hr): Determined from one difference between the Inlet
of the 11 Intensity-Duration- Elevation (19) and the Upper End
Frequency (IDF) curves developed Hydraulic Gradient Elevation (21).
by the Department. Intensity depends Show this number to one hundredth
on the design frequency and the time of a foot.
of concentration. Show this number
in inches per hour. 21. Upper End Elevation (ft) (Hydraulic
Gradient): The elevation of the
15. Total (C*A): The sum of the sub-total hydraulic gradient at the upper end of
C*A values (11). the pipe section. The elevation, under
design conditions, to which water will
16. Base Flow (cfs): This is an optional rise in the various inlets and
column to account for known flows manholes. Show this number to one
from underdrains, offsite pipe hundredth of a foot.
connections, etc. Show this number
in cubic feet per second. 22. Lower End Elevation (ft) (Hydraulic
Gradient): The elevation of the
17. Total Flow (cfs): The product of the hydraulic gradient at the lower end of
intensity (14) and the Total C*A (15) the pipe section. Show to one
plus Base Flows (16). Show this hundredth of a foot.
number in cubic feet per second.
23. Upper End Elevation (ft) (Crown): The
18 Minor Losses (ft): This is an optional inside crown elevation at the upper
column to account for minor losses end of the pipe section. Show this
according to Section 3.6.2 of the number to one hundredth of a foot.
Drainage Manual. Show this number
to one hundredth of a foot. 24. Lower End Elevation (ft) (Crown): The
inside crown elevation at the lower
19. Inlet Elevation (ft): The elevation of end of the pipe section. Show to one
the edge of pavement for curb inlets hundredth of a foot.
(Standard Plans, Indexes 425-020
through 425-025 and 425-061). The 25. Upper End Elevation (ft) (Flowline):
elevation of the theoretical grade The flowline at the upper end of the
point for barrier wall inlets of pipe section. Show this number to one
Standard Plans, Indexes 425-030 hundredth of a foot.
and 425-032. The grate elevation as
shown in the Indexes for barrier wall 26. Lower End Elevation (ft) (Flowline):
inlet (Standard Plans, Index 425-031) The flowline at the lower end of the
and gutter inlets (Standard Plans, pipe section. Show this number to one
Indexes 425-040 and 425-041). The hundredth of a foot.
grate elevation for ditch bottom inlets
(Standard Plans, Indexes 425-050 27. Fall (ft) (Hydraulic Gradient): The
through 425-055). The elevation of elevation change of the hydraulic
the manhole cover for manholes. grade line from the upper end to lower
Show this number to one hundredth end of the pipe section. Show this
of a foot. number to one hundredth of a foot.

Chapter 6: Storm Drains 6-4


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

this number to a minimum of one


28. Fall (ft) (Crown/Flowline): The tenth of a foot per second.
physical fall of the pipe section. Show 36. Physical Velocity (fps): The velocity
this number to one hundredth of a produced when the pipe is flowing full
foot. based on the Physical Slope (33).
Show this number to a minimum of
29. Number of Barrels: This optional one tenth of a foot per second.
column should be used for systems
V = (1.49/n)R2/3SPHYSICAL1/2
with pipe segments that have multiple
barrels.
37. Full-Flow Capacity (cfs): This
optional column is the product of the
30. Pipe Size (Rise) (in): The vertical
Physical Velocity (36) and the cross-
distance between the Flowline (25)
sectional area of the pipe. Show this
and the Crown (23) in inches.
number in cubic feet per second.
31. Pipe Size (Span) (in): The horizontal
39. Zone: One of the 11 FDOT Rainfall
distance of the inside of a pipe at its
Zones published in Appendix B.
widest point in inches.

32. Slope (%) (Hydraulic Gradient): For 39. Frequency (Yrs): The Storm Drain
pipes under pressure flow, this is the Design Frequency according to
full-flow friction slope. For pipes Section 3.3 of the Drainage Manual.
flowing partially full, this is: [Upper 40. Manning’s “n”: For storm drains, this
End HG (21) – Lower End HG (22)] value should be 0.012 according to
/Hydraulic Length (8). Show this Section 3.6.4 of the Drainage
number to one hundredth of a Manual. Document any other
percent. Manning’s “n” values used in the
Remarks Column (42).
33. Slope (%) (Physical): Determined
from Physical Fall (28)/Hydraulic 41. Tailwater EL (ft): The water elevation
Length (8). Show this number to one coincident with the outlet pipe and
hundredth of a percent. established by Section 3.4 of the
Drainage Manual. Some districts may
34. Slope (%) (Minimum Physical): The have more stringent criteria.
flattest physical slope to maintain a 42. Remarks: Include such things as:
velocity of 2.5 FPS flowing full, area adjustments, partial flow
obtained from rearranging Manning’s depths, existing pipe connections, or
equation: anything unusual.
S MIN = [Vn / (1.49R2/3)] 2

Show this number to one hundredth


of a percent.

35. Actual Velocity (fps): Determined by


Total Flow (17) divided by the
average cross-sectional flow area.
See discussion in Section 6.5. Show

Chapter 6: Storm Drains 6-5


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

6.2 HYDROLOGY
The rational method is used for pipe sizing, inlet capacity, and spread calculations.

Q=CiA
where:
Q = Runoff, in cubic feet per second (cfs)
C = Runoff Coefficient (see Table B-4, Appendix B)
i = Rainfall intensity, in inches per hour
A = Area, in acres

6.2.1 Design Frequency


The Drainage Manual states the design frequency for storm drains. For the
Department’s facilities, the frequencies range from three-year to 50-year, with three-
year as the most common design frequency. These frequencies apply to pipe hydraulics,
not inlet capacity or spread within the roadway. Section 6.3, below, discusses the criteria
for inlet capacity and spread. If a storm drain system includes both curb inlets and ditch
bottom inlets, check the ditch bottom inlets for a 10-year design frequency and all
structures in the mixed system should meet the three-year design frequency.

6.2.1.1 Storms of Greater Magnitude


You should always consider the intent of the Department’s criteria regarding the flooding
of properties upstream or downstream of the Department’s right of way. In several
chapters of the Drainage Manual, it says that any increases over pre-development
stages must not significantly change land use values. So, you should consider the
impacts of storm events that are more severe than the standard design frequency of the
storm drain. Initially, this should be a qualitative evaluation. Realize there are several
reasons why urban typical sections with storm drains can handle storms of greater
magnitude.

The first reason is conservatism within the storm drain design procedure. The flow rate
calculated for each pipe section is the peak flow rate. This is conservative because we
calculate the hydraulic gradient assuming that peak flow rates exist in all of the pipe
sections simultaneously. In reality, when one pipe section is at peak flow, usually one or
more of the other pipe sections have flow rates less than peak. This is most evident
when considering the differences between the upper and lower parts of a long system.
For example, consider a system where the outlet pipe’s flow is calculated based on a
Time of Concentration of 35 minutes. The flow rates of the first several pipes were based
on Times of Concentration of 10 minutes to 15 minutes. If a 35-minute storm and its
associated intensity is applied to the entire system, the flow rates in the first several

Chapter 6: Storm Drains 6-6


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

pipes would be less than the flow rate we calculated based on Times of Concentration
= 10 minutes to 15 minutes. Therefore, the friction losses in these pipes actually are less
than we calculated. Conversely, during short, intense storms, the upper pipes could
reach their design flow rates, but the downstream portion of the system does not have
the entire area contributing, so downstream pipes do not see the design flows. This
conservatism exists to some degree throughout all pipe systems, but has a minimal
effect on short systems where the differences in Times of Concentration are small.

Another reason an urban typical section can handle storms of greater magnitude is that
the roadway itself can convey substantial flow. A standard pavement section of 0.02 ft/ft
cross slope on a 0.3-percent longitudinal grade can convey approximately 7 cfs 1 with
the depth of the flow at the top of the curb.

The last reason, although less significant, is that when the flow in the road reaches the
height of the curb, there is more pressure on the piping system, thus forcing more flow
through the pipes.

Considering these reasons, look at the system to see if there are any places where the
water elevations or discharge rates could be increased.

● Are there sags in the profile? If so, could the pond water leave the right of way
at these locations? Would water have gone that direction in the pre-developed
condition?

• Is the roadway blocking overland flow in any areas? If so, would the blocked
water substantially change land use values?

• Where back-of-sidewalk inlets are used, should check valves or flap gates be
used to prevent the water in the pipes from backing off of the right of way?

• Would the inlets at the ends of the system bypass flow during a more severe
storm event? If so, would water have gone that direction in the pre-developed
condition?

If you have concerns after considering these factors, it may be appropriate to do a more
detailed evaluation. Perhaps check the operation of the storm drain system with higher
frequency (less frequent) storm. Perhaps the storage in the road and the pipes could be
modeled. You may need a more detailed model of the pre-developed conditions.

If, after evaluating these situations, it is evident there would be increased discharge or
increases over pre-developed stages that would significantly change land use values,

1 Q = (0.56/n)•Sx1.67•S0.5•T8/3= (0.56/0.016) • 0.021.67 • 0.0030.5 • 18.758/3 ≈ 7 cfs where T = curb


height / cross slope = (4.5/12)/0.02 = 18.75’

Chapter 6: Storm Drains 6-7


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

document this conclusion. Then change the storm drain design as necessary to bring
the stages down or to reduce the discharge. This is not saying use a higher design
frequency for the storm drain system. Instead, use larger pipes where necessary.
Increasing pipe size to prevent the adverse impacts to adjacent properties is different
than using a higher design frequency and maintaining the standard hydraulic gradient
clearance.

6.2.2 Time of Concentration


The Time of Concentration (tc) is the time required for the runoff to travel from the most
remote point in the drainage basin to the point of the storm drain system under
consideration. This will be the longer of: (a) the overland flow time to the inlet, or (b) the
sum of the tc to the inlet immediately upstream in the piping system plus the time of flow
through the upstream pipe section. For inlets that have more than one upstream pipe,
you will need to compare the tc and Time of Flow through Section of all the upstream
inlets and pipes with the overland travel time to the subject inlet. Use the longest of
these as the tc. See Figure 6.2-1. For pipe segments that do not have upstream pipes,
the tc will be simply the overland flow time.

Figure 6.2-1: Determining Time of Concentration

Chapter 6: Storm Drains 6-8


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

6.2.2.1 Peak Flow from Reduced Area


Check to see if a portion of the drainage area will produce a larger flow rate than the
entire area. This could occur where a larger portion of the drainage area exists toward
the bottom or outlet, as in Figure 6.2-2. This is even more likely if the land cover of the
area toward the outlet is more impervious than the upstream area. Mathematically, you
will observe this where the reduction in area is more than offset by an increased intensity
and possibly an increased runoff coefficient.

The Department encourages that this check be made at apparent junctions or inlets in a
storm drain system. It is acceptable but not necessary to check every pipe section for
peak flow from reduced area. Some computer programs may do this automatically.

Scott Street

Figure 6.2-2

Chapter 6: Storm Drains 6-9


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Example 6.2-1—Peak Flow from a Reduced Area


Given:
● The partial storm drain system shown in Figure 6.2-3
● Project located in San Antonio, Pasco County, Zone 6

Find:
● The design flow rate for pipe section P31-32.

S30
TC = 11 min.
3.7 Ac @ C =0.95
0.3 Ac @ C =0.20
T sect (P30-31) = 1 min

S29 S31 S32


S29
TC= 26 min. T sect (P29-31) = 2 min
1.3 Ac @ C=0.95
0.7 Ac @ C=0.20

Figure 6.2-3: Example 6.2.1

First, calculate the flow rate using the total drainage area (maximum tc).

1. Add the product of C*A for the upstream areas.


Total C*AS-29 = 0.95 x 1.3 ac + 0.2 x 0.70 ac = 1.38
Total C*AS-30 = 0.95 x 3.7 ac + 0.2 x 0.30 ac = 3.58
Total C*AS-31 = 4.96

2. Determine the time of concentration.

The tc is the time it takes for the entire drainage area to contribute. It is the longer
of:
(tc)S-29 + Time of Flow in Section P29-31 = 26 + 2 = 28 min
(tc)S-30 + Time of Flow in Section P30-31 = 11 + 1 = 12 min
Therefore, (tc)S-31 = 28 min.

3. Determine the intensity.

From the IDF curve, the intensity is 4.0 iph.

4. Determine the flow.

Q = (C*A) x i = 4.96 x 4.0 = 19.8 cfs

Chapter 6: Storm Drains 6-10


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Now, check for a larger flow from part of the drainage area (peak flow from
reduced area).

5. Determine the intensity associated with the shorter tc.

The shorter system time is from S-30 and is 11 + 1 min = 12 min.


The intensity in Zone 6 for 12 min = 6.0 iph.

6. Estimate the area that will contribute from S-29 during a 12-minute storm.

One approach is to reduce the area from the pipes having long times of
concentration by the ratio of the times of concentrations. Ratio = (Short tc)/(tc of
the associated pipe).

AS-29 is reduced by 12 min/28 min = 0.43


AS-29 REDUCED @ C = 0.95 = 1.3 ac x 0.43 = 0.56 ac
AS-29 REDUCED @ C = 0.20 = 0.7 ac x 0.43 = 0.30 ac

7. Add the areas that will contribute to S-31 during a 12-minute storm.

Area TOTAL = AreaS-29 REDUCED + AreaS-30


@ 0.95 = 0.56 + 3.7 = 4.26
@ 0.20 = 0.30 + 0.3 = 0.60

8. Add the product of C*A contributing to S-31 during a 12-minute storm.

Total C*A = 0.95 x 4.26 + 0.2 x 0.6


= 4.05 + 0.12 = 4.17

9. Determine the flow from the reduced area.

QReduced Area = (C x A) x i12 min


= 4.17 x 6.0 = 25.0 cfs

For pipe sections downstream of P31-32, add the incremental drainage areas to the
reduced areas recorded for P31-32. Then, add the time of flow in downstream sections
to the reduced time of concentration for P31-32.

Table 6.2-1 shows a way of presenting these approaches on the Tabulation form.

Chapter 6: Storm Drains 6-11


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Table 6.2-1

TIME OF FLOW IN SECTION


TIME OF CONCENTRATION
STRUCTU

NOTES
NUMBER

TYPE OF STRUCTURE DRAINAGE AREA


AND

Example Approach
RE

(ac.)

INTENSITY (IPH )
REMARKS

TOTAL FLOW
C•A—TOTAL
(MINUTES)

(MINUTES)
C = 0.95 SUB- Zone:

(cfs )
TOTAL
C= C•A Frequency (Yrs):
UPPER
C = 0.2 Manning’s “n”:
Tailwater EL (ft):
LOWER INCREMENT TOTAL

0.2 1.3 1.24


29 J1 26 2 1.38

Total Area – Max tc


31 0 0.7 0.14
n/a 3.7 3.52
30 J1 11 1 3.58
31 n/a 0.3 0.06
5.0 4.75
31 MH 28 - 4.0 4.95 19.8
32 1.0 0.20
0.2 1.3 1.24

Peak Flow from Reduced Area


29 J1 26 2 1.38

31 0 0.7 0.14
n/a 3.7 3.52
30 J1 11 1 3.58
31 n/a 0.3 0.06
Area from S-29 reduced
0.2 4.26 4.05
by 12/28. Intensity based
31 MH 0.0 12 - 6.0 4.17 25.0 on System Time from S-
30
32 0.6 0.12

6.2.2.2 Ignoring Time of Flow in Section


For systems where the pipes are full without a storm event because of normal tailwater
conditions, the time of flow in the pipe section is meaningless. For the runoff to get into
the pipe, the water that is in the pipe has to move out. Since water under the pressures
we are dealing with is essentially incompressible, what goes in the inlet must be coming
out the outlet at the same time. In these situations, it is realistic to ignore the travel time
through pipes submerged by normal tailwater. Note that you should use normal tailwater
(perhaps the control elevation of a wet pond), not the design tailwater, to determine if a
pipe segment is submerged.

The Department realizes that current design software does not use the approach of
ignoring time of flow in section. As such, some districts may not require that time of flow
be ignored through submerged pipes.

Chapter 6: Storm Drains 6-12


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

6.3 INLETS AND PAVEMENT HYDRAULICS


6.3.1 Inlets
Factors controlling the selection of an inlet type include hydraulics, utility conflicts, right-
of-way limits, bicycle and pedestrian safety, etc. Guidelines for selecting inlets are
located in Section 3.7.4 of the Drainage Manual.

6.3.1.1 Apparent Locations


● At low points in the gutter. Double-throated inlets—such as Type 2, 4, and 6—
are symmetrical about the centerline and are intended to accept flow from both
sides. Use these where the minor gutter flow exceeds 50 feet in length or 0.5
cubic feet per second.

● Upstream of pedestrian cross walks.

● Upstream of curb returns.

● 10 feet outside the flat cross sections in super elevation transitions.


Although the flow may be small, the cross slope is nearly flat so the spread
potential is high.

● Outside of driveway turnouts. If the adjacent property is under development or


redevelopment, try to obtain the site plans to identify future driveway locations.

6.3.1.2 Sags
Normally, one inlet at the low point in combination with inlets on each of the approaching
grades is sufficient to meet spread criteria.

Use flanking inlets for sags that have no outlet other than the storm drain system—for
instance, underpasses, barrier wall sections, or depressed sections where the roadway
is much lower than the surrounding ground. Flanking inlets are those placed on one or
both sides of and fairly close to the sag inlet. They provide backup capacity for the sag
inlet if it becomes clogged. The flanking inlets must be located to satisfy spread criteria
when the sag inlet is blocked. Figure 6.3-1 shows a representation of this location
pattern. Figure 6.3-9 provides vertical curve formulae to help determine the flanking inlet
locations.

Chapter 6: Storm Drains 6-13


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Figure 6.3-1

6.3.1.3 Continuous Grades


When deciding about the initial placement of curb and gutter inlets on a continuous
grade, you should base placement on the 300-foot maximum spacing for an 18-inch
pipe (Drainage Manual, Section 3.10.1). After the initial placement of inlets, check the
spread and add or move inlets as necessary to meet the spread standards.

The piping system layout may affect the locations of curb and gutter inlets. As you lay
out the piping system, you may need a manhole to redirect the flow, to provide
maintenance access, or merely to connect stub pipes. If you use an inlet rather than a
manhole, you get the benefit of an additional hydraulic opening for little or no additional
cost. Section 6.4, below, discusses piping system layout.

6.3.1.4 Back of Sidewalk


Locate back-of-sidewalk inlets where concentrated
flows drain toward the road and where the The Field Review is Critical
proposed sidewalk would block overland flow. to Designing
Often, you can identify these areas from the survey, Back-of-Sidewalk Drainage
the back-of-sidewalk profiles, and the proposed
cross sections. Do not rely on these alone! Walk the entire project area looking for places
where concentrated runoff flows to the road and for localized depressed areas that were
not identified in the survey. Development may have changed the existing ground line
since the time the survey was done. Your field review with the back-of-sidewalk profiles
and proposed cross sections will identify areas where you need back-of-sidewalk inlets.

In instances where you may need numerous back-of-sidewalk inlets, check with the

Chapter 6: Storm Drains 6-14


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

roadway designer about modifying the roadway profile grade to better accommodate
overland flow.

Standard Plans, Index 425-060 contains the standard back-of-sidewalk drainage inlets.
Use yard drains and the double four-inch pipes under the sidewalk to correct small
existing flooding problems. For any other back-of-sidewalk drainage, obtain right of way
as necessary to construct a ditch bottom inlet or other substantial back-of-sidewalk
drainage conveyance.

Where the Department’s storm drain system connects to back-of-sidewalk inlets, check
the hydraulic grade line elevation at these inlets to see if water would back up or cause
the system to create adverse impacts to adjacent properties. If so, first consider
increasing the size of some downstream pipe sections. If avoiding adverse impacts by
increasing pipe sizes is not feasible, consider using check valves or flap gates in the
pipe connected to the back-of-sidewalk inlet (see Figure 6.3-2). Flap gates and check
valves are not ideal because they require maintenance; nevertheless, they may be the
most practical option for some situations.

Flap Gate
(Designer to specify)
Yard Drain
(Index 282)
or DBI
To Storm Drain

Underdrain Inspection Box


Index 245

Figure 6.3-2

6.3.1.5 Inlet Capacity


Capacity data for most of the Department’s inlets were developed by laboratory studies
done at the University of South Florida (Anderson, 1972). A graphical presentation of
these data is given in Appendix I. Separate curb inlet capacity charts are presented for
various cross slopes. You also can use methods described in FHWA’s Hydraulic
Engineering Circular HEC 12 or HEC 22 to evaluate the interception capacity of the
Department’s inlets.

Chapter 6: Storm Drains 6-15


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

6.3.2 Pavement Hydraulics


The Department uses driver visibility as a basis for the spread standards. There is a
rainfall intensity that reduces the driver’s sight distance to less than the minimum
stopping sight distance. Removing the water from the road for intensities greater than
this serves no purpose. If a driver’s sight distance is less than the minimum stopping
sight distance when the driver sees an object, the driver cannot stop in time regardless
of how much water is on the road. So removing the water from the road for intensities
greater than this is over-design from a vehicle standpoint.

The Department uses four inches per hour (iph) as the intensity that reduces the driver’s
sight distance to less than the minimum stopping sight distance. This is based on
information summarized in FHWA HEC 21.

Use the integrated form of Manning’s equation to calculate spread in gutters.

0.56 5 3 1 2 8 3
Q= S x SL T
n
where:
Q = Gutter flow rate, in cubic feet per second (cfs)
n = Manning’s roughness coefficient (see Table B-2, Appendix B)
Sx = Pavement Cross Slope, in feet per feet (ft/ft)
SL = Longitudinal Slope, in feet per feet (ft/ft)
T = Spread, in feet (ft)

Figure 6.3-8 provides a nomograph for solving this equation, which is intended for use
with triangular gutter sections. The standard Type F curb forms a composite section
when combined with the pavement cross slope. In most cases, it is reasonable to ignore
the gutter depression and treat the flow section as a simple gutter formed by the cross
slope of the road and the curb. Ignoring the gutter depression is conservative2, but
allows for debris buildup in the gutter. If you need to determine the additional capacity
of the gutter depression, use Figure 6.3-10 or the procedures provided in FHWA’s HEC
12 or HEC 22.

6.3.2.1 Gutter Grades


Standard gutter grades should not be less than 0.3 percent. Some District Drainage
Engineers will approve a 0.2-percent gutter grade in very flat terrain. Use of a saw tooth

2 The gutter depression can add approximately 31% to the conveyance of the flow section in cases
where the pavement cross slope is 0.02 and the spread width is 7.5 feet.

Chapter 6: Storm Drains 6-16


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

profile can maintain minimum grades in very flat terrain.

To provide adequate drainage in sag vertical curves, maintain a minimum gutter grade
of 0.3 percent down to the inlet at the low point. Without this minimum grade, the flat
longitudinal grade near the low point would cause the spread to be greater than
allowable. Maintaining the minimum gutter grade up to the inlet increases the cross
slope at the low point, thus providing additional drainage. To maintain the minimum
gutter grade, develop and show special gutter grades in the plans.

Example 6.3-1—Special Gutter Grade


Given:
● The sag vertical curve described in the figure below.

Length of V.C. = 250'


28'
PG Point @ 0.02

½ Typ. Section

Figure 6.3-3: Example 6.3-1 Given Information

Find:
● The limits of the special gutter grade
● The theoretical gutter elevation at the low point
● The cross slope at the low point

1. Determine the rate of change of longitudinal slope. (Formula from Fig. 6.3-9)

Rate of change = r = (g2 – g1)/L = 0.6 – (-0.5)/2.5 = 0.44

2. Find the location of the low point and the location where the longitudinal slope on
the curve is -0.3 percent and +0.3 percent. Use the equation for longitudinal slope
at any point and rearrange to solve for X.

X-0.3% = (SL – g1)/r


= [-0.3 – (-0.5)]/0.44 = 0.4545 stations

Chapter 6: Storm Drains 6-17


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

∴ Station = 48+00 + 0+45.45 = 48+45.45

X+0.3% = (SL – g1)/r


= [0.3 – (-0.5)]/0.44 = 1.8182 stations
∴ Station = 48+00 + 1+81.82 = 49+81.82
Using the equation for the station of the turning point:

X LOW POINT = (g1 x L)/(g1 – g2)


= (-0.5 x 2.5)/(-0.5 – 0.6)
= 1.1364 stations
∴ Station = 48+00 + 1+13.64 = 49+13.64
So, a special gutter grade of -0.3 percent is needed from Sta. 48+45.45 to Sta.
49+13.64 and a special gutter grade of +0.3 percent is needed from Sta.
49+13.64 to Sta. 49+81.82.

3. Find the elevation of the profile grade line at Sta. 48+45.45 and Sta. 49+81.82.
Both are equal distance from the center, so we only need to find one elevation.

Elev48+45.45 = Elev48+00 +g1 X + ½ r X2


= 35.386 ft. + (-0.5)(0.4545) + ½ (0.44)(0.4545)2
= 35.204 ft.

4. Find the elevation at the gutter at Sta. 48+45.45. (This equals the elevation of the
gutter at Sta. 49+81.82.)

The edge of pavement is 0.56 ft. (28 ft. x 0.02) lower than profile grade line and
the gutter is 1.5 inches (0.125 ft.) below the edge of pavement, so:

ElevGUTTER = Elev PGLSta. 48+45.45 – 0.56 – 0.125 = 35.204 ft. – 0.56 ft. – 0.125 ft.
= 34.519'

5. Find the theoretical gutter elevation at the low point.

Elev = ElevSta. 48+45.45 – (special gutter grade x length of special gutter)


Elev = 34.519 ft. – [0.3 x (49.1364 – 48.4545)] = 34.314 ft.

This elevation would be used to check the hydraulic grade line clearance below
the sag inlet.

6. Find the cross slope at the low point.

The elevation of the profile grade line at the low point is:

PGL Elev49+13.64 = Elev48+00 +g1 X + ½ r X2

Chapter 6: Storm Drains 6-18


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

= 35.386 ft. + (-0.5)(1.1364) + ½ (0.44)(1.1364)2


= 35.102 ft.

The elevation at the edge of pavement at the low point is:


EOP Elev49+13.64 = ElevGUTTER + 1.5 inches
= 34.314 ft. + 0.125 ft. = 34.439 ft.

Cross Slope = (35.102 ft. – 34.439 ft.)/28 ft. = 0.024 ft./ft.

This would be used to check the spread of the inlet at the low. Interpolate
between the values in Appendix I, Figures I-17 through I-19, where the cross
slope value is between the values of the figures.

6.3.2.2 Cross Slope


FDM 210 and 211 gives the standard cross slopes.

6.3.2.3 Shoulder Gutter


Use shoulder gutter on fill slopes and at bridge ends to protect the slopes from erosion
caused by water from the roadway and bridge. Use shoulder gutter in accordance with
Section 3.7.3 of the Drainage Manual. When placed at bridge ends, the gutter should
be long enough to construct the gutter transitions shown on Standard Plans, Indexes
536-001 and 425-040. The terminal shoulder gutter inlet should intercept all of the flow
coming to it for a 10-year storm.

The Drainage Manual gives two spread criteria for sections with shoulder gutter. One is
related to driver visibility (rainfall intensity of four inches per hour) and the other is related
to erosion protection of the fill slope (10-year design storm). Both criteria need to be
met. Consider the potential for future additional lanes in the median when determining
the flow rates in shoulder gutter.

In a typical situation where standard cross slopes and shoulder widths exist, the criterion
for protecting the fill slope has a higher intensity and less allowable spread than the
criterion for driver safety. Thus, the criterion for protecting the fill slope will set the inlet
spacing.

Given the typical situation where both the shoulder and the miscellaneous asphalt
behind the gutter slope upward at 0.06 ft/ft from the gutter, the spread across the gutter
and pavement section should not exceed six feet for the 10-year storm. This section has
a conveyance of approximately 28 cubic feet per second [K = Q/SL½ = 28 cfs]. You can
use the conveyance to determine maximum allowable flow rates for various longitudinal
slopes. Another approach is to treat the shoulder gutter and pavement section as a
triangular gutter with a cross slope of 0.05 ft/ft, designing for 10-year flows, and limiting
the spread to six feet (see Figure 6.3-4).

Chapter 6: Storm Drains 6-19


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

1. The maximum shoulder gutter design conveyance should be K = 28 cfs adjacent


to guardrail, and K = 15 cfs with no guardrail for the 10-year storm. K = 28 cfs is
derived from the flow area being limited to 15 inches outside the shoulder gutter
and n = 0.016. K = 15 cfs is derived from limiting the flow area to the shoulder
gutter section.

2. The maximum shoulder gutter design conveyance approaching a terminal gutter


inlet should be K = 15 cfs to intercept 100 percent of the design storm flow.

3. Consider placing two gutter inlets at the down gradient shoulder gutter terminus
to provide 100-percent interception, unless 100-percent interception by one inlet
(K = 15 cfs) is demonstrated by appropriate calculation.

4. Inlet spacing must meet spread criteria (Drainage Manual, Section 3.9),
maximum pipe length criteria (Drainage Manual, Section 3.10.1), and 10-year
frequency gutter capacity criteria. In most cases, the 10-year frequency storm
may govern inlet spacing.

5. Where applicable, design inlet spacing to accommodate the additional runoff from
future widening.

6. Place gutter inlet(s) at the down gradient end of all shoulder gutter, in lieu of
concrete spillways or flumes, to reduce the potential for erosion.

Traffic Lane
Edge of

Figure 6.3-4

Chapter 6: Storm Drains 6-20


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

6.3.2.4 Determining the Spread


For roads that have uniform longitudinal grades and cross slopes, the spread
calculations may be as simple as calculating the spread and bypass for the inlets with
the largest overland flow. For these projects, you usually can make a reasonable
assumption that if the inlets with the largest overland runoff do not exceed the spread
standards and do not have any bypass, the other inlets will not exceed the spread
standards and will not have any bypass. If you cannot comfortably make this
assumption, you can determine the spread by the following procedure used with Table
6.3-1. In general, the information in Table 6.3-1 is the minimum required for spread
calculations. You may need additional information in certain situations.

Start at the upper-most inlet and work to the low point, then start at the opposite high
side and work back to the low.

1. Determine the drainage area and runoff coefficient of the overland runoff. Record
the product of the area and runoff coefficient (C*A) in column 2.

2. Calculate the overland runoff by multiplying the product of C*A in column 2 by the
appropriate intensity (four inches per hour or 10-year storm design).
Q = C • A • i.

3. Calculate the total flow to the inlet by adding the overland runoff in column 3 to
the bypass from the upstream inlets.

4. Record the cross slope and longitudinal slope in column 6 and column 7,
respectively.

5. Calculate the spread and compare it to the allowable spread, keeping in mind
that allowable spread can vary along the project due to super-elevation slope
toward the median, turn lanes, and design speed. If it is within the standards,
record the number in column 8 and go to the next step. If it is not, move the inlet
(and add and move inlets as necessary) to make the spread acceptable and
repeat Step 1 through Step 5.

6. Calculate intercepted flow and bypass flow. (The figures in Appendix I can be
used in lieu of calculations to determine intercepted flow. Record these numbers
in column 9 and column 10, respectively.

7. Proceed to the next downstream inlet and repeat Step 1 through Step 6.

Chapter 6: Storm Drains 6-21


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Table 6.3-1: Spread Calculations


FLORIDA DEPARTMENT OF TRANSPORTATION
SPREAD CALCULATIONS
Road: Sheet of _______
County: Prepared by: Date _____
Financial Project ID: Checked by: Date _____
System Description: ____________________________________________________

Manning's n = _______

Bypass to
Inlet No. Cross Long Allowable Inlet No.
or C•A Overland Previous Spread or
Runoff By-pass Total Flow Slope Slope Spread Intercepted Bypass
Location (ft/ft) (%) Flow Flow to Inlet @
(1) (2) (3) (4) (5) (6) (7) (8) (8a) (9) (10) (11)

Chapter 6: Storm Drains 6-22


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Example 6.3-2—Sag Vertical Curve Given


Given:
● The sag vertical curve and associated approach grades shown below.
● Four-lane curb and gutter section with 12-foot lanes, 12-foot continuous two-way
left-turn lane, four-foot bike lane, and six-foot sidewalk.
● Offsite drains to the road from 65 feet beyond the sidewalk.
● Offsite area draining to the road is impervious (C = 0.95).
● Type 1 and Type 2 inlets are preferred by the District.
● Inlet location is not restricted by driveways or side streets.
• Design speed = 45 mph, then allowable spread is 11.5 feet [1.5-foot gutter + four-
foot bike lane + six feet (half of a travel lane)].
• A minimum gutter grade of 0.3 percent is used approaching the sag.

900' 900'

320' V.C.

42'

Typical Section
Cross Slope = 0.02 ft / ft

Figure 6.3-5: Example 6.3-2 Given Information

Find:
● Inlet spacing necessary to meet the spread criterion

1. For the first try, place the inlets at the maximum 300-foot spacing out from the
low point. So, the inlets will be placed at Station 44+00, Station 47+00, Station
50+00, Station 53+00, and Station 56+00.

The area to each inlet on the approach grades is:


Area = (½ Rdwy Width + 65 ft.) x 300 ft.
Area = (42 ft.+ 65 ft.) x 300 ft./43,560 = 0.74 ac @ C = 0.95
CxA = 0.95 x 0.74 = 0.70
Q OVERLAND = CAi
= 0.95 x 0.74 x 4 = 2.8 cfs.

Chapter 6: Storm Drains 6-23


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

2. Determine the spread, the intercepted flow, and the bypass (if any) for the
uppermost inlets. (Sta. 44 & 56)

Spread (T) = [(Q x n)/(0.56 x SX5/3 x SL½)]3/8


This conservatively ignores the 1.5-inch gutter depression.

= [(2.8 x 0.016)/(0.56 x 0.025/3 x 0.02½)]3/8

= 9.3 ft. Acceptable (allowable spread is 11.5 ft.)


QINTRCEPT ~ 2.1 cfs From Figure I-1, Appendix I
QBYPASS = 2.8 – 2.1 = 0.7 cfs

3. Determine total flow to the next downstream inlets. (Sta. 47 & 53)

QTOTAL = QOVERLAND +QBYPASS


= 2.8 + 0.7 = 3.5 cfs

4. Determine the spread, the intercepted flow, and the bypass.

Spread = 10.2 ft. Still acceptable


Q INTRCEPT = 2.3 cfs From Figure I-1.
Q BYPASS = 3.5 – 2.3 = 1.2 cfs

5. Determine the spread approaching the sag inlet from either side.

Q TOTAL = QOVERLAND +QBYPASS


= 2.8 + 1.2 = 4.0 cfs

6. Determine the spread approaching the sag inlet. The longitudinal slope is 0.3
percent approaching the sag. For this example, the cross slope at the low point
is 0.021 ft./ft. due to maintaining 0.3-percent gutter grade to the sag inlet. This
was calculated using the approach in Example 6.3-1.

T = [(4.0 x 0.016)/(0.56 x 0.0215/3 x 0.003½)]3/8


= 14.7 ft. Not acceptable (Allowable spread is 11.5 ft.)

The following table summarizes the above calculations.

Chapter 6: Storm Drains 6-24


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Flow in cfs Allowable spread = 11.5 ft. Manning's n = 0.016


1 2 3 4 5 6 7 8 9 10 11
Inlet C•A Overland Previous Total Cross Long Spread Intercepte Bypass Bypass
Location Runoff Bypass Flow Slope Slope d Flow Flow to Inlet @
(Sta.) (ft/ft) (%) Station

44+00 0.70 2.8 -- 2.8 0.02 2.0 9.3 2.1 0.7 47+00

47+00 0.70 2.8 0.7 3.5 0.02 2.0 10.2 2.3 1.2 50+00

56+00 0.70 2.8 -- 2.8 0.02 2.0 9.3 2.1 0.7 53+00

53+00 0.70 2.8 0.7 3.5 0.02 2.0 10.2 2.3 1.2 50+00

50+00 0.70 2.8 1.2 4.0 0.021 0.3 14.7 Exceeds Standard
Approac
h

7. Add and adjust inlets.


There is no direct solution. It is a trial-and-error process of moving inlets to reduce
the spread. Adding an inlet to each side of the sag and adjusting the spacing of
the inlets on the continuous grades should reduce the flow to the sag inlet and
reduce the spread. Try placing the inlets at Stations 44+30, 46+80, 48+80,
50+00, 51+20, 53+20, and 55+70.

Figure 6.3-6

Chapter 6: Storm Drains 6-25


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Example 6-3.2—Second Iteration


The drainage area to the first continuous grade inlets (43+30, 55+70) is:

Area = (½Rdwy Width + 65 ft.) x 250 ft.


Area = (42 ft. + 65 ft.) x 330 ft./43,560 = 0.81 ac @ C = 0.95

The drainage area to the next inlets (46+80, 53+20) is:

Area = (42 ft. + 65 ft.) x 250 ft./43,560 = 0.61 ac @ C = 0.95

The drainage area to the next inlets (48+80, 51+20) is:

Area = (42 ft. + 65 ft.) x 200 ft./43,560 = 0.49 ac @ C = 0.95

The drainage area to each side of the sag is:

Area = (42 ft. + 65 ft.) x 120 ft./43,560 = 0.29 ac @ C = 0.95

The inlets at stations 48+80 and 51+20 are on the vertical curve; therefore, the
longitudinal slope is flatter than 2.0 percent. Using vertical curve formulae:

Rate of change of grade (r) = (g2 – g1)/L


= [2 – (-2)]/3.2 = 1.25 ft./station

Long slope = g1 + r X (X is distance along curve in Sta.)


= -2 + 1.25 (0.4)
= -1.5 percent

For this example, the cross slope at the low point is 0.021 ft./ft. due to maintaining 0.3-
percent gutter grade down to the sag inlet. You can calculate this using the approach in
Example 6.3-1. Using Figure I-17 (cross slope = 0.02) will provide a slight conservatism.

The following table shows the results of the change.

Chapter 6: Storm Drains 6-26


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Flow in cfs Allowable Spread = 11.5 ft Manning's n = 0.016


1 2 3 4 5 6 7 8 9 10 11
Inlet Cross Long
Location Overland Previous Total Slope Slope Spread Intercepted Bypass Bypass to
(Sta) CxA Runoff By-Pass Flow (ft/ft) (%) (ft) Flow Flow Inlet @
44+30 0.77 3.1 0 3.1 0.02 2 9.7 2.2 0.9 46+80
46+80 0.58 2.3 0.9 3.2 0.02 2 9.8 2.2 1.0 48+80
48+80 0.47 1.9 1.0 2.9 0.02 1.5 9.9 2.4 0.5 50+00
50+00 0.28 1.1 0.5 1.6 0.021 0.3 10.5 n/a n/a n/a

55+70 0.77 3.1 0.0 3.1 0.02 2 9.7 2.2 0.9 53+20
53+20 0.58 2.3 0.9 3.2 0.02 2 9.8 2.2 1.0 51+20
51+20 0.47 1.9 1.0 2.9 0.02 1.5 9.9 2.4 0.5 50+00
50+00 0.28 1.1 0.5 1.6 0.021 0.3 10.5 n/a n/a n/a

50+00 0.56 2.2 1.0 3.2 0.021 n/a 6.3 n/a n/a n/a

In an actual project, the inlet location is affected by driveways and side streets.

Example 6.3-3—Shoulder Gutter


Given:
● The bridge approach grades shown below.
● Four-lane rural divided highway, two 12-foot lanes, 10-foot paved outside
shoulder, four-foot slope to gutter under guardrail (3-foot paved)
● Cross slope of shoulder and asphalt under guardrail = 0.06 ft./ft.
● Fill slope is 10 feet high at Station 67+00
● Project located in Zone 7, 10-year/10-minute intensity = 7.4 in/hr
● Additional lanes may be added in the future
● Runoff from bridge = 0.2 cfs (scuppers used on bridge)

Figure 6.3-7: Example 6.3-3 Given Information

Chapter 6: Storm Drains 6-27


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Find:
1. The location of the shoulder gutter inlets.

2. Determine the vertical curve geometry.

Crest Curve:
Rate of change of curve (r) = (g2 – g1)/L = (-2.6 – 3.0)/14 = -0.4
Long Slope at any point X = g1 + r X = 3.0 – 0.4 X

Sag Curve:
Rate of change of curve (r) = (g2 – g1)/L = (3.0 – 0.0)/5 = 0.6
Long Slope at any point X = g1 + r X = 0.0 + 0.6 X

3. Estimate the lowest point at which shoulder gutter is needed.

You should use shoulder gutter on all fill slopes greater than 10 feet long if the
roadway grade is greater than 2 percent. For this example, the fill is
approximately 10 feet at Station 67+00. The longitudinal slope at this station
67+00 is: 0.6 x (67 – 63) = 2.4 percent. Since this is steeper than 2 percent,
shoulder gutter should begin at or before Station 67+00.

4. For the first try at inlet spacing, divide the distance between Station 67+00 and
the beginning of the bridge into equal distances that are less than 300 feet.
Distance = 74+75 – 67+00 = 775 feet

This equates to three spaces at approximately 258 feet. Rounding it to 260 feet,
the first inlet will be located at 74+75 – 260 feet = 72+15. The other inlets are at
69+55 and 66+95.

5. Determine the longitudinal slope at these inlets:

@ 72+15, the longitudinal slope = 3 – 0.4 (72+15 – 70+00) = 3 – 0.4 x 2.15 = 2.14
percent
@ 69+55, the longitudinal slope = 3 percent
@ 66+95, the longitudinal slope = 0.6 x (66+95 – 63+00) = 0.6 x 3.95 = 2.37
percent

6. Determine area and overland runoff to each inlet.

An additional lane may be added toward the median in the future, so use 36 feet
of pavement.
Width = travel lanes + shoulder + gutter + slope* under guardrail.
*Conservatively assume that all four feet sloping back to gutter is paved.

Chapter 6: Storm Drains 6-28


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

= 36 +10 + 3.5 + 4 = 53.5 ft.

Area = 260 ft. x 53.5 ft. = 0.32 ac.


C x A = 0.95 x 0.32 =0.30

The travel time for flow along 260 feet of pavement is small, so we will use the
10-minute intensity for the 10-year storm. i = 7.4 iph
Q = CiA = 0.95 x 7.4 x 0.32 = 2.2 cfs

7. Approximate the shoulder gutter as a triangular section with a cross slope of


0.05 ft./ft. and n = 0.016. The spread must be limited to six feet in this triangular
section to match the capacity of the shoulder gutter section. See previous
discussion of shoulder gutter.

Spread (T) = [(Q x n)/(0.56 x SX5/3 x SL1/2)]3/8

The intercepted flow is determined from Figure I-16. The following table
summarizes the calculations.

All flows (cfs) based on 10-year flow Allowable Spread = 6 ft Manning's n = 0.016
1 2 3 4 5 6 7 8 9 10 11
Inlet Overland Previous Cross Long Intercepte Bypass Bypass
Location C•A Runoff By-pass Total Flow Slope Slope Spread d Flow Flow to Inlet @
(Sta.) (ft/ft) (%) Station

72+15 0.30 2.2 0.2 2.4 0.05 2.14 4.9 2.4 ---
69+55 0.30 2.2 --- 2.2 0.05 3.0 4.5 2.2 ---
66+95 0.30 2.2 --- 2.2 0.05 2.37 4.7 2.2 ---

This inlet spacing meets the spread criterion for protecting the fill slopes and the last
inlet captures all the runoff coming to it. Therefore, this design is acceptable. There is
no need to check the four inches per hour criterion because the intensity would be less
and the allowable spread would be greater.

Chapter 6: Storm Drains 6-29


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Reference: FHWA Hec-12 (1984)


Figure 6.3-8

Chapter 6: Storm Drains 6-30


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

VERTICAL CURVE FORMULAE

g 2 − g1
r=
1. Rate of Change of Grade: L

1
y= rX 2
2. Offset from tangent to curve: 2

1
E x = E pvc + g 1 X + ( rX 2 )
3. Elevation for any point on curve: 2

∂E x
= g1 + rX
4. Grade (longitudinal slope) at any point: ∂X

5. Station from PVC to turning point (local tangent horizontal) on a curve:


gL
X = 1
g 1− g 2

1  Lg 12 
ETP = E pvc −  
2  g 2 − g1 
6. Elevation of turning point:

Where:
All horizontal dimensions (X) are in Stations.
All vertical dimensions (E) are in Feet.
All grades are in percent.
L = Length of vertical curve in Stations.
Epvc = Elevation of the Point of Vertical Curve.

Figure 6.3-9: Vertical Curve Formulae

Chapter 6: Storm Drains 6-31


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Conveyance vs. Spread


for Composite Gutter Sections with Type E or Type F Curb
Manning’s n = 0.016

70

60
[ K ] Conveyance Q / SL^0.5 (cfs)

50

40
Sx = 0.03
Sx = 0.02
Sx = 0.01
30

20

10

0
2 4 6 8 10 12

Spread (ft)

Based on FHWA HEC-12, App. C. SL = Longitudinal Slope (ft/ft)


Sx = Pavement Cross Slope (ft/ft)

Example
Given: Q = 3 cfs, Sx = 0.03, SL = 0.04 ft/ft

Find Spread:
1. K = Q/SL 0.5 = 3 / 0.04 0.5 = 15
2. From Chart at K = 15 & Sx = 0.03, Spread = 6.0 ft

Figure 6.3-10: Conveyance vs. Spread for Composite Gutter Sections with Type
E or Type F Curb

Chapter 6: Storm Drains 6-32


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

6.4 PIPE SYSTEM PLACEMENT


6.4.1 Plan Layout
After the inlets have been placed to drain the pavement adequately, lay out the piping
system to connect the inlets. While laying out the system, you will add manholes as
necessary to redirect the flow, or to provide maintenance access, or merely to connect
stub pipes. At this stage, consider adding an inlet instead of a manhole. When an inlet
is used instead of a manhole, you get the benefit of an additional hydraulic opening for
little or no additional cost.

There are several items to consider that can influence the piping system plan layout.
The most important issues are hydraulics, constructability, and utility conflicts.

• Avoid placing pipes that would oppose flows from other pipes, especially in high-
velocity situations. Impinging flows can be avoided by staggering the elevations of
the pipes entering a junction box.

• Consider right of way necessary to open the trench for the pipes. This is especially
important for deep pipes. You might use temporary sheet piling during installation to
reduce the trench width, but this is very costly, so you will want to explore other
alternatives (e.g., moving the trunk line).

• Use either a manhole or an inlet at changes in flow direction. This will provide
maintenance access where debris and sediment often collect.

• Preferably, place manholes in or behind the sidewalk. This allows access without
closing the travel lanes and is much safer for maintenance personnel. If you must
place manholes in the pavement, avoid putting the lids in the wheel path.

• Minimize interference with major utilities, such as fiber optic lines and sanitary and
potable water lines greater than eight inches in diameter. See the discussion in
Section 6.4.3.

• Where there is one main trunk line, place it on the side of the road constructed first.
This prevents constructing stub lines that can’t be drained.

• Where there is one main trunk line, locate it, if possible, on the low side of super-
elevated roadway sections to minimize the depth of cut.

• Where there is one main trunk line, consider connecting several inlets along the
opposite side of the road from the trunk line, and then running only one pipe laterally
across the road. This will reduce the number of cuts across the road.

Chapter 6: Storm Drains 6-33


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

• Consider using two trunk lines to minimize the number of cuts across the road and
thus simplify maintenance. In such cases, you should weight the gains in improved
maintenance against the increased cost of the additional trunk line.

6.4.1.1 Retaining Wall Drainage


Whenever possible, avoid placing piping within mechanically stabilized earth (MSE)
retaining wall embankments. If placing pipes within the MSE wall section is your only
option, please refer to Section 3.11 of the Drainage Manual.

6.4.2 Profile Placement


6.4.2.1 Slopes
The Drainage Manual states that the minimum physical slope must produce a velocity
of 2.5 feet per second flowing full. The slope is obtained from the velocity form of
Manning’s equation using the full cross sectional area of the pipe:

V = (1.49/n) R2/3 S1/2


rearranging: S = [V n/(1.49 R2/3)]2

where: R is based on the full cross sectional area


V = 2.5 fps

Table 6.4-1 provides the minimum physical slope to produce 2.5 feet per second flowing
full for various pipe sizes with Manning’s roughness coefficient of 0.012. Note that the
velocity of 2.5 feet per second is not necessarily the actual velocity in the pipe under
design conditions. Refer to Section 6.5.5 for a discussion of flow velocity.

Chapter 6: Storm Drains 6-34


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Table 6.4-1

Minimum Physical Slope


n = 0.012
Diameter (inches) Slope (%)
18 0.150
24 0.102
30 0.076
36 0.059
42 0.048
48 0.041
54 0.035
60 0.030
66 0.027
72 0.024
78 0.021
84 0.019

For very flat systems, the minimum physical slope may not be realistic. The overall fall
across the system is based on outlet pipe depth and structural clearances at the upper
end. Most District Drainage Engineers will approve deviation from the minimum pipe
slope in these cases.

Where you cannot attain the minimum slope, try to design the system to avoid
appreciable drops in the velocity. This will help to carry sediment through the system
instead of dropping sediment at the velocity drop point in the system.

The minimum slope is 0.1 percent for systems under pressure flow.

Setting flow lines relates to the slope. Refer to FDM 300 for the accuracy level to which
you must display flow lines.

Chapter 6: Storm Drains 6-35


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

6.4.2.2 Minimum Pipe Depth


The minimum depth of the pipe is controlled either by the minimum pipe cover or by the
need to have clearance above the top of the pipe to maintain strength in a precast
structure. Minimum pipe cover requirements are given in Appendix C of the Drainage
Manual.

The loads placed on precast structures during shipping and handling often are greater
than the loads placed on them in their final location. Since contractors prefer precast
drainage structures and they have become the industry standard, you should consider
the potential for breakage during shipping and handling.

Where pipes are placed high in a structure, the structure has little, if any, strength above
the pipe. This can result in breakage during shipping or handling. For strength reasons,
it is best to maintain a minimum amount of precast concrete section above the pipe.

The ideal amount of precast section varies with the type of inlet and bottom
configuration. Generally, where a pipe is placed in grated inlets or in structure bottoms,
try to maintain a six-inch precast section that has full wall thickness above the pipe
opening, as shown in Figure 6.4-1. For ditch bottom inlets placed on J bottoms, the
recommended minimum precast riser section varies depending on if the unit has slots.
Refer to Structure/Pipe configuration numbers 4 & 5 in Figure 6.4-5. For ditch bottom
inlets without slots, maintain a 10-inch riser section below the grate seat. For ditch
bottom inlets with slots, maintain a 12-inch riser section below the slot.

6" Precast
Grate Seat Section

Precast
Opening*

* Opening is typically 6" larger than pipe O.D.

Figure 6.4-1

Chapter 6: Storm Drains 6-36


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Tables 6.4-4 and 6.4-5 give recommended minimum distances from inlets to pipe flow
lines for most of the Department’s standard inlets. These distances provide the precast
section discussed above and are based on concrete pipe that is centered in the precast
opening. The above discussion represents ideal values that you should try to achieve.
On occasion, you will need to use less precast section than discussed above. This is
acceptable because the contractor has the option to cast structures in place. When using
less precast section, you must add all the appropriate dimensions to assure that no
conflict will exist between pipes and the structure.

6.4.3 Utility Coordination


During the design process, avoid designing storm drain systems in the vicinity of utilities
where practical unless it would substantially increase the cost of the system. Try to
obtain information not only on the location of existing facilities, but on proposed locations
as well. The utility companies (both private and public) will view the design proposed on
the Phase II plans as part of the utility coordination process. You may be asked to attend
utility coordination meetings, which can be very beneficial to the design effort because
the concerned parties gather together to resolve utility placement conflicts. The utility
companies are accustomed to meeting face-to-face with FDOT representatives. The
Department and the utility companies usually negotiate final storm drain design and
utility locations, with the goal to minimize the costs to the public.

Sometimes minor changes in the storm drain design can reduce the cost to a utility
company and minimize the cost to the public. At other times, it may not be practical or
cost effective to accommodate a utility company proposal. Utility companies often take
the opportunity to upgrade their systems or add facilities during the Department’s
construction project. Do not assume they will relocate their systems in the process.

On projects with long storm drain systems in areas of many utilities, include one
additional manhole in the quantities for unforeseen utility conflicts.

6.4.4 Pipe-to-Structure Connections


When a bridge deck piping system connects to a roadway structure, use a resilient
connector to accommodate the expected thermal movement of the bridge and its piping
system.

Check sizes of structure bottoms to make sure that the pipes fit. When doing so, use the
outside diameter of concrete pipe 3. It has the thickest wall of any of the optional pipe
materials. Type P structure bottoms are either 4’-0” or smaller diameter round (Alternate

3
An easy way to remember the wall thickness of the concrete pipe is to take the inside diameter in feet
and add one (1). The result is the wall thickness in inches. Examples: 30” pipe, I.D. = 2.5 feet, Wall
Thickness = 2.5 + 1 = 3.5”.

Chapter 6: Storm Drains 6-37


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

A) or 3’-6” square (Alternate B). 30-inch pipe is the maximum size that will fit in Type P
bottoms.. The contractor has the option of using either Alternate A or Alternate B for
Type P bottoms unless restricted by the plans. Type J structure bottoms are larger than
Type P bottoms and come in various sizes, as described in Standard Plans, Index 425-
010. The plans usually specify the alternate and the size of the J bottoms. Standard
Plans, Index 425-010 gives the minimal structure dimensions for various pipe sizes.
Table 6.4-2 is an excerpt of Standard Plans, Index 425-010. Refer to the latest version
of the index for updates.

Table 6.4-2: Excerpt of Standard Plans, Index 425-010

Chapter 6: Storm Drains 6-38


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

The skew at which a pipe enters a precast rectangular structure is limited by the precast
pipe opening. The maximum opening is six inches larger than the pipe outside diameter
(Standard Plans, Index 425-001). The maximum pipe skew varies with the structure wall
thickness and the pipe size. The maximum skew for various pipe sizes passing through
eight-inch structure walls is shown in Table 6.4-3. Standard Plans, Index 425-010
provides skew values for six-inch structure walls and other pipe sizes. Use round
structure bottoms (Alternate A) where the pipe enters the structure at a larger angle.

Good practice to allow 2" of


Structure Wall contruction tolerance.

Skew

Figure 6.4-2
Standard Plans, Index 425-001 includes a detail of a pipe opening at a corner of a
structure. Although a detail exists for this condition, restrict its use to situations where
other alternatives do not exist. Make every attempt to ensure pipes do not enter the
corner of rectangular structures (“corner-cutouts”).

When placing pipes in existing rectangular structures, the maximum skew is limited by
the dimension of the skewed pipe cut fitting between the walls.

Table 6.4-3

Pipe Size
18" 24" 30" 36" 42" 48" 54" 60"

Max. Skew 19o 17o 16o 16o 15o 14o 14o 13o
These values are based on two inches of construction tolerance, precast structures with eight-inch walls,
and concrete pipe dimensions.
When using round structure bottoms, consider the need to maintain a precast section

Chapter 6: Storm Drains 6-39


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

between the openings of adjacent pipes. Try to maintain at least a two-inch section along
the inside wall between adjacent pipe openings, as shown to the right. Table 6.4-6
provides the minimum angle between adjacent pipe centerlines to maintain the two-inch
precast section along the inside wall. The values in Table 6.4-6 are based on equal pipe
centerline elevations and standard concrete pipe openings. Using these minimum
angles for pipes with offset centerline elevations and other pipe materials is conservative
and would yield more than two inches of precast section.

2"

Precast Opening

Figure 6.4-3

Where large pipes are stubbed into the main line or a large main line pipe makes a 90-
degree turn, rectangular structures can be smaller than round structures given the same
pipe sizes. Figure 6.4-4 shows 48-inch pipes making a 90-degree turn at a structure. An
eight-foot round structure is needed, while a six-foot rectangular structure would work.

Chapter 6: Storm Drains 6-40


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Figure 6.4-4

Chapter 6: Storm Drains 6-41


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Table 6.4-4: Recommended Min. Distance from Inlet Elevation to Pipe Flow Line
PIPE LOCATION RECOMMENDED MIN. DISTANCE (FT.) FROM GRATE (INLET) ELEVATION TO PIPE FLOW LINE
INLET SLOT
TYPE TYPE Wall Wall Dim. 15” Pipe 18” Pipe 24” Pipe 30” Pipe 36” Pipe 42” Pipe 48” Pipe 54” Pipe 60” Pipe
Short 2’-0” 2.2 (1) 2.5 (1) 4.8 (5) 5.4 (5) 5.9 (5) 6.5 (5) 7.0 (5) 7.5 (5) 8.1 (5)
Type A
Long 3’-1” 2.5 (2) 2.8 (2) 4.8 (5) 5.4 (5) 5.9 (5) 6.5 (5) 7.0 (5) 7.5 (5) 8.1 (5)
No Slot 2.6 (2) 2.9 (2) 3.4 (2) 4.0 (2) 6.9 (4) 7.5 (4) 8.0 (4) 8.5 (4) 9.1 (4)
Type B Short
Travers Under Slot 3.4 (3) 3.6 (3) 4.2 (3) 4.7 (3) 6.9 (4) 7.5 (4) 8.0 (4) 8.5 (4) 9.1 (4)
(Note 3)
Long 2.6 (2) 2.9 (2) 3.4 (2) 4.0 (2) 4.5 (2) 7.5 (4) 8.0 (4) 8.5 (4) 9.1 (4)
Short 2’-0” 2.2 (1) 2.5 (1) 4.7 (5) 5.2 (5) 5.8 (5) 6.3 (5) 6.8 (5) 7.4 (5) 7.9 (5)
None
Long 3’-1” 2.4 (2) 2.6 (2) 4.7 (5) 5.2 (5) 5.8 (5) 6.3 (5) 6.8 (5) 7.4 (5) 7.9 (5)
No Slot 2.2 (1) 2.5 (1) 5.3 (4) 5.8 (4) 6.3 (4) 6.9 (4) 7.4 (4) 8.0 (4) 8.5 (4)
Short
Type C Travers Under Slot 2.8 (3) 3.0 (3) 5.3 (4) 5.8 (4) 6.3 (4) 6.9 (4) 7.4 (4) 8.0 (4) 8.5 (4)
(Note 3) Long 2.4 (2) 2.6 (2) 5.3 (4) 5.8 (4) 6.3 (4) 6.9 (4) 7.4 (4) 8.0 (4) 8.5 (4)
No Slot 2.2 (1) 2.5 (1) 5.7 (4) 6.2 (4) 6.8 (4) 7.3 (4) 7.8 (4) 8.4 (4) 8.9 (4)
Non-Trav Short
Under Slot 3.2 (3) 3.5 (3) 5.7 (4) 6.2 (4) 6.8 (4) 7.3 (4) 7.8 (4) 8.4 (4) 8.9 (4)
12” Std.
Long 2.4 (2) 2.6 (2) 5.7 (4) 6.2 (4) 6.8 (4) 7.3 (4) 7.8 (4) 8.4 (4) 8.9 (4)
Short 3’-1” 2.4 (2) 2.6 (2) 3.2 (2) 5.2 (5) 5.8 (5) 6.3 (5) 6.8 (5) 7.4 (5) 7.9 (5)
None
Long 4’-1” 2.2 (1) 2.5 (1) 3.0 (1) 3.5 (1) 4.1 (1) 6.3 (5) 6.8 (5) 7.4 (5) 7.9 (5)
Short 2.4 (2) 2.6 (2) 3.2 (2) 5.8 (4) 6.3 (4) 6.9 (4) 7.4 (4) 8.0 (4) 8.5 (4)
Type D Travers No Slot 2.2 (1) 2.5 (1) 3.0 (1) 3.5 (1) 4.1 (1) 6.9 (4) 7.4 (4) 8.0 (4) 8.5 (4)
Long
(Note 3) Under Slot 2.8 (3) 3.0 (3) 3.6 (3) 4.1 (3) 4.7 (3) 6.9 (4) 7.4 (4) 8.0 (4) 8.5 (4)
Short 2.4 (2) 2.6 (2) 3.2 (2) 6.2 (4) 6.8 (4) 7.3 (4) 7.8 (4) 8.4 (4) 8.9 (4)
Non-Trav
No Slot 2.2 (1) 2.5 (1) 3.0 (1) 3.5 (1) 4.1 (1) 7.3 (4) 7.8 (4) 8.4 (4) 8.9 (4)
12” Std. Long
Under Slot 3.2 (3) 3.5 (3) 4.0 (3) 4.5 (3) 5.1 (3) 7.3 (4) 7.8 (4) 8.4 (4) 8.9 (4)
Short 3’-0” 2.2 (1) 2.5 (1) 3.0 (1) 5.2 (5) 5.8 (5) 6.3 (5) 6.8 (5) 7.4 (5) 7.9 (5)
None
Long 4’-6” 2.4 (2) 2.6 (2) 3.2 (2) 3.7 (2) 4.3 (2) 6.3 (5) 6.8 (5) 7.4 (5) 7.9 (5)
No Slot 2.2 (1) 2.5 (1) 3.0 (1) 5.8 (4) 6.3 (4) 6.9 (4) 7.4 (4) 8.0 (4) 8.5 (4)
Short
Type E Travers Under Slot 2.8 (3) 3.0 (3) 3.6 (3) 5.8 (4) 6.3 (4) 6.9 (4) 7.4 (4) 8.0 (4) 8.5 (4)
(Note 3) Long 2.4 (2) 2.6 (2) 3.2 (2) 3.7 (2) 4.3 (2) 6.9 (4) 7.4 (4) 8.0 (4) 8.5 (4)
No Slot 2.2 (1) 2.5 (1) 3.0 (1) 6.2 (4) 6.8 (4) 7.3 (4) 7.8 (4) 8.4 (4) 8.9 (4)
Non-Trav Short
Under Slot 3.2 (3) 3.5 (3) 4.0 (3) 6.2 (4) 6.8 (4) 7.3 (4) 7.8 (4) 8.4 (4) 8.9 (4)
12” Std.
Long 2.4 (2) 2.6 (2) 3.2 (2) 3.7 (2) 4.3 (2) 7.3 (4) 7.8 (4) 8.4 (4) 8.9 (4)

Notes: 1. The number in parentheses ( ) refers to one of the structure pipe combinations shown in Figure 6.4-5.
2. *** CAUTION *** Where multiple pipes enter a structure, needing a J-bottom because of one pipe could dictate greater distances than shown above for
other pipes entering the structure.
3. The values shown for Type B, C, D, and E inlets are based on Alternate B Bottoms. Alternate A Bottoms have thicker slabs, so add two inches for up
through six-foot diameter bottoms. Add four inches for eight-foot diameter bottoms.
4. The distances are based on precast structures and standard precast openings for concrete pipes.
5. The designer should check that the minimum cover requirements of Drainage Manual, Appendix C are met.

Chapter 6: Storm Drains 6-42


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Figure 6.4-5

Chapter 6: Storm Drains 6-43


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Table 6.4-5: Recommended Min. Distance from Inlet Elevation to Pipe Flow Line
PIPE LOCATION RECOMMENDED MIN. DISTANCE (FT.) FROM GRATE (INLET) ELEVATION TO PIPE FLOW LINE
INLET SLOT
TYPE TYPE Wall Wall Dim. 15” Pipe 18” Pipe 24” Pipe 30” Pipe 36” Pipe 42” Pipe 48” Pipe 54” Pipe 60” Pipe
Short 2’-6” 2.2 (1) 2.5 (1) 4.8 (5) 5.3 (5) 5.8 (5) 6.4 (5) 6.9 (5) 7.5 (5) 8.0 (5)
Type F n/a
Long 4’-0” 2.4 (2) 2.7 (2) 3.3 (2) 3.8 (2) 4.3 (2) 6.4 (5) 6.9 (5) 7.5 (5) 8.0 (5)

Short 3’-0” 2.2 (1) 2.5 (1) 3.0 (1) n/a n/a n/a n/a n/a n/a
None
Long 6’-7” 2.4 (2) 2.6 (2) 3.2 (2) 3.7 (2) 4.3 (2) 4.8 (2) 5.3 (2) 5.9 (2) 6.4 (2)
Type H
Non-Trav Short 3’-0” 3.2 (3) 3.5 (3) 4.0 (3) n/a n/a n/a n/a n/a n/a
12” std. Long 2.4 (2) 2.6 (2) 3.2 (2) 3.7 (2) 4.3 (2) 4.8 (2) 5.3 (2) 5.9 (2) 6.4 (2)
6’-7”

Short 3’-3” 2.6 (2) 2.9 (2) 3.4 (2) 5.5 (5) 6.0 (5) 6.5 (5) 7.1 (5) 7.6 (5) 8.2 (5)
Type J n/a
Long 3’-10” 2.4 (2) 2.7 (2) 3.3 (2) 3.8 (2) 6.0 (5) 6.5 (5) 7.1 (5) 7.6 (5) 8.2 (5)

Short 3’-3” 2.6 (2) 2.9 (2) 3.5 (2) 5.5 (5) 6.0 (5) 6.6 (5) 7.1 (5) 7.7 (5) 8.2 (5)
Type S n/a
Long 3’-10” 2.3 (2) 2.5 (2) 3.1 (2) 3.6 (2) 6.0 (5) 6.6 (5) 7.1 (5) 7.7 (5) 8.2 (5)

Short 3’-3” 2.6 (2) 2.9 (2) 3.4 (2) 5.5 (5) 6.0 (5) 6.5 (5) 7.1 (5) 7.6 (5) 8.2 (5)
Type V n/a
Long 3’-10” 2.4 (2) 2.7 (2) 3.3 (2) 3.8 (2) 6.0 (5) 6.5 (5) 7.1 (5) 7.6 (5) 8.2 (5)

Manhole RECOMMENDED MIN. DISTANCE3 (FT.) FROM TOP ELEVATION TO PIPE FLOW LINE
n/a n/a
Type 8 3.7 (10) 4.0 (10) 4.5 (10) 5.0 (10) 6.3 (11) 6.8 (11) 7.3 (11) 7.9 (11) 8.4 (11)

RECOMMENDED MIN. DISTANCE (FT.) FROM LOW POINT OF GRATE TO PIPE FLOW LINE
Barr-
n/a Short 3’-3” 4.2 (8) 4.5 (8) 5.0 (8) 6.2 (9) 6.8 (9) 7.3 (9) 7.8 (9) 8.4 (9) 8.9 (9)
Wall 218
Long 3’-8” 4.2 (8) 4.5 (8) 5.0 (8) 5.5 (8) 6.8 (9) 7.3 (9) 7.8 (9) 8.4 (9) 8.9 (9)

RECOMMENDED MIN. DISTANCE (FT.) FROM EDGE OF PAVEMENT TO PIPE FLOW LINE
Curb 1-9 n/a n/a
3.9 (6) 4.2 (6) 4.7 (6) 5.3 (6) 6.5 (7) 7.0 (7) 7.5 (7) 8.1 (7) 8.6 (7)

Notes: 1. The number in parentheses ( ) refers to one of the structure pipe combinations shown in Figure 6.4-6.
2. *** CAUTION *** Where multiple pipes enter a structure, needing a J-bottom because of one pipe could dictate greater distances than shown above for
other pipes entering the structure.
3. *** CAUTION *** For curb inlets and manholes, where 30” pipes with similar inverts enter a structure at 90 degrees,a J-bottom is required, thus the
minimum distance may be greater than shown above. This may apply to other inlets also.
4. The distances are based on precast structures and standard precast openings for concrete pipes.
5. The designer should check that the minimum cover requirements of Drainage Manual, Appendix C are met.

Chapter 6: Storm Drains 6-44


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Figure 6.4-6

Chapter 6: Storm Drains 6-45


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Table 6.4-6

Chapter 6: Storm Drains 6-46


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

6.5 PIPE HYDRAULICS


The Drainage Manual states that you must consider friction losses in the computation of
the design hydraulic gradient for all storm drain systems. Energy losses associated with
pollution control structures (weirs and baffles) and utility conflict structures also must be
considered when present in a system. When the hydraulic calculations consider only the
above, the elevation of the hydraulic gradient must be at least one foot below the
theoretical gutter elevation. This is equivalent to 13.5 inches (1.13 feet) below the edge
of pavement for sections with Type E or Type F curb and gutter. For gutter inlets
(Standard Plans, Indexes 425-040 and 425-041), ditch bottom inlets used within the
roadway section (Standard Plans, Indexes 425-050 through 425-055), and barrier wall
inlets (Standard Plans, Index 425-031), the one foot of clearance applies to the grate
elevation. For barrier wall inlets (Standard Plans, Indexes 425-030 and 425-032), the one
foot of clearance applies to the theoretical grade point.

If you calculate all minor energy losses, it is acceptable for the hydraulic grade line to
reach the theoretical gutter elevation. Minor losses include all the losses at inlets,
manholes, and junctions due to expansion, contraction, and changes in flow direction.
Minor losses also include exit losses at the outlet of the system. The Drainage Manual
states that minor losses must be calculated when the velocity is greater than 7.5 fps and
for systems longer than 2,000 feet.

6.5.1 Pressure Flow


Under pressure flow conditions, the pipe section flows full throughout. Calculate friction
losses using Manning’s equation, with the flow area equal to the full cross sectional area
of the pipe.
29n 2LV 2 4.61n 2LQ 2
Head loss [in feet] = 1.33
=
R 2g D5.33

where:
n = Roughness coefficient (refer to the Drainage Manual)
L = Pipe length, in feet (ft.)
V = Velocity, in feet per second (fps)
Q = Flow rate, in cubic feet per second (cfs)
R = Hydraulic radius, in feet (ft.) = Area/wetted perimeter
D = Pipe diameter, in feet (ft.)
g = Gravitational constant = 32.2 ft/s2

Chapter 6: Storm Drains 6-47


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

6.5.2 Partially Full Flow


For pipes that are flowing partially full, the
calculations are more complicated. The
cross-sectional flow area actually changes
as the flow goes through the pipe. For
example, the flow area at the downstream
end of the pipe shown here is the full cross
section area, but at the upstream end the
flow area is much less.

The most accurate approach to calculating this number is to do water surface profile
calculations throughout the pipe section. Although acceptable, these calculations are
tedious and not usually required. The Department accepts the following approach to
calculating the hydraulics of partial full and pressure pipes.

Three values must be determined for each pipe section: (1) the lower-end hydraulic
gradient (lower-end HG), (2) the upper-end hydraulic gradient (upper-end HG), and (3)
the flow velocity.

6.5.3 Lower-End Hydraulic Gradient


Either the downstream hydraulic gradient or the flow conditions in the pipe controls the
lower-end HG. So you must compare the water surface elevations associated with these
numbers and use the higher of the two as the lower-end HG.

Where the downstream HG is above the


lower-end crown of the subject pipe, the The downstream hydraulic gradient
lower-end HG is the downstream HG. elevation is the downstream pipe Upper End
See Detail A of Figure 6.5-1. Pipe flow HG elevation plus junction losses, if they are
conditions will not control in this calculated.
situation, and comparing water surface
elevations is not necessary. If the downstream HG is below the lower-end crown, you will
need to compare the downstream HG with the water elevation associated with the pipe
flow conditions.

Where the downstream hydraulic gradient is low enough, one of two pipe flow conditions
will control the lower-end HG. See Detail C & D of Figure 6.5-1. The appropriate flow
condition is dependent on the relationship of the physical pipe slope and the full-flow
friction slope. If the pipe is sloped steeper than the full-flow friction slope, it is reasonable
to assume that normal depth flow exists at the lower end. Then the lower-end HG is the
normal depth plus the lower-end flow line elevation. (Actual depth could be above normal
depth because the pipe was not long enough to allow normal depth to be reached.)

Chapter 6: Storm Drains 6-48


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Partial depth pipe flow graphs in Appendix E, or the Department’s hydraulic calculator,
can be used to get normal flow depth and associated velocity.

If the pipe slope is equal to or flatter than the full-flow friction slope, the pipe is flowing full
over most of its length. Although the flow may be dropping through critical depth 4 near
the outlet, assuming full flow at the outlet is reasonable and conservative. During very low
flow rates, even flat pipes will not flow full, but such low rates are not typical for design
conditions.

In short, use the higher of the following for the lower-end HG, as shown in Figure 6.5-1.

Condition 1: The downstream pipe upper-end HG (+ junction losses, if calculated)

OR

Condition 2: The normal depth + lower-end flow line elevation (for pipes sloped steeper
than full-flow friction slope) or lower-end crown elevation (for pipes sloped equal to or
flatter than full-flow friction slope).

For the outlet pipe of the system, the lower-end HG elevation is the Design Tailwater
elevation.

4 - For a slightly more refined analysis in this situation, midway between critical depth and the crown of the
pipe of [(Dc+ D)/2} could be used as the Lower End HG.

Chapter 6: Storm Drains 6-49


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Figure 6.5-1: Determining the Lower End Hydraulic Gradient Elevation

Chapter 6: Storm Drains 6-50


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

6.5.3.1 Design Tailwater (DTW)


The Drainage Manual gives the standard design tailwater conditions. In general, it says
use the higher of the crown of the pipe or the downstream condition. Stormwater ponds
are commonly constructed at the outlet of storm drains, so the pond stage may be the
design tailwater. Some Districts may have more stringent criteria than shown in the
Drainage Manual.

You can determine the pond stage by “routing” the storm drain design event (frequency)
through the pond. “Routing” refers to the use of the storage indication method that is
commonly used to simulate runoff hydrographs flowing through stormwater management
facilities. HEC 22 contains a discussion and example of the storage indication method.

6.5.4 Upper-End Hydraulic Gradient


Use the higher of the following, as shown in Figure 6.5-2:

Condition 1: The lower-end HG plus the full-flow friction loss

OR

Condition 2: The elevation of normal depth in the pipe at the upper end

A comparison may not be necessary. First, add the full-flow friction loss to the lower-end
HG. If this is above the upper-end crown, there is no need to calculate normal depth. The
lower-end HG plus full-flow friction loss will control.

Chapter 6: Storm Drains 6-51


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Figure 6.5-2: Determining the Upper-End Hydraulic Elevation

Chapter 6: Storm Drains 6-52


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

6.5.5 Flow Velocity in the Pipe Section


For pressure flow pipes, the velocity is
based on the full cross section area. Where the tailwater conditions submerge
the storm drain without stormwater flow,
Q Q the travel time in the pipe can be ignored
Velocity(fps) = = since the velocity is irrelevant. See the
A πD2
discussion of ignoring travel time in
4 Section 6.2.

where:
Q = Flow rate, in cubic feet per second (cfs)
D = Diameter, in feet (ft.)

For pipes flowing partially full, it is more complicated to determine the velocity. There can
be a water surface profile in the pipe, so the cross sectional flow area can change, thus
changing the velocity along the pipe section. The most accurate velocity should represent
the average velocity through the pipe section, assuming the velocity associated with
normal depth is a conservative assumption. See Figure 6.5-3.

You can use partial depth pipe flow graphs in Appendix E, or the Department’s hydraulic
calculator, to get normal flow depth and associated velocity.

The flow velocity is referred to as the Actual Velocity in the Storm Drain Tabulation Form,
Figure 6-1. The actual velocity is sometimes called the design velocity.

Chapter 6: Storm Drains 6-53


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Figure 6.5-3: Determining the Velocity in Partially Filled Pipes

Chapter 6: Storm Drains 6-54


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

6.5.6 Utility Conflict Box Losses


Calculate the loss through a utility conflict box using the equation:
V2
Head Loss [in feet] = K
2g
Where:
K = Loss factor (or coefficient)
V = Flow velocity in the storm drain, in feet per second (ft/s)
g = gravitational constant = 32.2 ft/s2

Use Figure 6.5-4 to determine the loss factor in conflict boxes where the pipes are flowing full.

Figure 6.5-4: Loss Factors for Conflict Manholes

Chapter 6: Storm Drains 6-55


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

6.5.7 Minor Losses


Minor losses are all the losses that are not due to friction. Generally, these are energy
losses due to changes or disturbances in the flow path. They include such things as
entrance, exit, bend, and junction losses. The losses are calculated from the equation:

2
Vo
Head loss [in feet] = K
2g
where:
K = Loss factor (or coefficient)
Vo = Flow velocity in the outlet pipe of the junction box, in feet per second (ft/s)
g = Gravitational constant = 32.2 ft/s2

FHWA has printed the latest information on computing minor losses in HEC 22, and they
continue to do research on minor losses. A report titled Junction Loss Experiments:
Laboratory Report summarizes work that has been done more recently than the
information published in HEC 22. The report and HEC 22 are available on the Internet at:

http://www.fhwa.dot.gov/engineering/hydraulics/library_listing.cfm?sort=Publication_title&archived=0

It is important to calculate minor losses in high-velocity situations and in long systems. As


the velocity approaches eight feet per second, the velocity head (V2/2g) approaches one
foot (64/64.4). The standard one foot of HGL clearance would be used up where the total
loss coefficient, k, equals 1.0. For long systems, the 1.0 foot of clearance could be used
up by numerous small individual junction losses, e.g., 10 junctions with 0.1 foot minor loss
each.

Chapter 6: Storm Drains 6-56


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

6.6 PROCEDURE
The following is a basic procedure for designing a storm drain system. You can vary
slightly from the procedure and still develop an adequate design. With experience, you
will develop shortcuts and personal preference. The goal is to minimize pipe sizes while
meeting the appropriate standards.

The numbers in parentheses (xx) refer to a space on the Storm Drain Tabulation Form.

IDENTIFY INLET LOCATIONS


1. Define the overall basin Using the drainage map, identify the overall watershed
draining to the project. that drains to the project.
2. Determine the outfalls and
This is typically done as a part of the stormwater
divide the overall basin into sub-
management design.
basins.
3. For each sub-basin, select
inlet locations.
4. Determine the drainage area
to each inlet.

5. Calculate spread and revise


inlet location, as necessary.

LAYOUT PIPES
6. Connect pipes between the
You will use the schematic of the piping system for the
inlets to create a schematic of
rest of the design procedure.
the piping system layout.

Chapter 6: Storm Drains 6-57


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

DETERMINE THE TOTAL “C*A” PRODUCT FOR EACH PIPE SECTION


Begin filling out the Storm Sewer Tabulation Form. Record the Inlet Types (7), Inlet
Locations (3) (4) (5), Inlet Elevations (19), Structure Numbers (6), Incremental Areas
(9), C-Factors (1), and Length (8) on the tabulation form. The incremental areas and
C-factors are those used to calculate the spread.
This involves checking for all the upstream areas.
7. Add the areas that contribute Refer to the piping system schematic to ensure that
flow to the downstream pipe. all the areas are included. Record these in the
space (10) on the tabulation form.

8. Multiply the subtotal areas Record the result in space (11) on the tabulation
by their respective C-factors. form.

9. Add the sub-total (C*A) Record the total in space (15) on the tabulation
values. form.
10. Repeat Step 7 through
Step 9 for the entire system.

PRELIMINARY HYDRAULIC GRADE LINE (HGL) SLOPE


11. Estimate a preliminary For flat terrain, estimate which inlet will be critical.
hydraulic grade line slope. The critical inlet usually will be the lowest inlet in
the portion of the system farthest from the outlet. It
This slope will be used as a may be simply the inlet farthest from the outlet. Use
guide for selecting the trial pipe the following formula to calculate the slope.
size only. It will not control the
final design. Slope =
Critical Inlet Elev - DTW - 1 foot
System Length between Outlet & Critical Inlet

For moderately sloped terrain, an average slope of


the ground line along the project usually is
acceptable for a preliminary HGL slope.

For some systems, there may be two or more


distinct sections of the system that have noticeably
different slopes. For these, calculating a preliminary
HGL slope for each section is advised.

Chapter 6: Storm Drains 6-58


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

CALCULATE RUNOFF FLOW RATES


The following is the beginning of an iterative process of calculating flow rates
as you move down the system and calculating hydraulic grade line elevations
as you move up the system. Step 12 through Step 18 are done on each pipe
segment, beginning at the upper end of the system and working down toward
the outlet.
12. Determine the tc. Record the value in space (12) on the tabulation form.
13. Determine the intensity. Determine the intensity from the appropriate IDF curve
using a storm duration equal to the tc previously
computed. Record the value in space (14) on the
tabulation form.
14. Calculate total runoff for Multiply the total CA times the intensity. Record the
the pipe segment. value in space (17) on the tabulation form.
15. Select a pipe size. For the first pass through the system, select a diameter
that has a full-flow friction slope close to the
preliminary HGL slope. The minimum pipe diameter
will probably control the pipe size of the first few pipe
sections. You will probably not find a pipe diameter that
matches the preliminary HGL exactly. The objective is
to maintain the standard HGL clearance at each inlet.
Matching the preliminary HGL is merely a technique to
begin selecting pipe diameters. Some pipe diameters
will likely be revised later.

Record the pipe size (30) (31) and associated


Minimum Physical Slope (34) on the tabulation form.
Record the full-flow friction slope as the hydraulic
grade line slope (32) during the first pass down the
system. Use the full-flow friction slope in the calculation
of the hydraulic gradient.
16. Determine the pipe flow The flow lines usually will be controlled by such things
lines, fall, and physical as cover requirements, structure clearances, and
slope. minimum physical pipe slope. Record the Flow Line
Elevations (25) (26), Crown Elevations (23) (24),
Physical Slope (33), and Pipe Fall (28) on the
tabulation form.
17. Calculate the flow Actual Velocity: For the first pass through the system,
velocity. assume full flow unless the pipe is obviously flowing
part full. For subsequent passes through the system,
use full-flow velocity or velocity associated with normal

Chapter 6: Storm Drains 6-59


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

depth, as appropriate. See discussion in Section 6.5.


Record the value in space (35) on the tabulation form.

Physical Velocity: Record the value in space (36) of the


tabulation form.
18. Calculate time of flow in Divide the pipe length by the actual flow velocity.
pipe section. Record the value in space (13) on the tabulation form.
19. Repeat Step 12 through Check for peak flow from reduced area. See
Step 18 for the entire discussion in Section 6.2.
system.

Chapter 6: Storm Drains 6-60


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

CALCULATE HYDRAULIC GRADE LINE (HGL) ELEVATION

Step 20 and Step 21 are done on each pipe segment, beginning at the outlet and
working up the system toward the most remote inlet.
20. Determine the Lower- The Lower-End HG for the outlet is the Design
End Hydraulic Gradient Tailwater (DTW). See the discussion in Section 6.5.
Elevation. Record the value in space (22) on the tabulation form
for the outlet pipe and in space (41).
21. Determine the Upper- See the discussion in Section 6.5. Record the Upper-
End HG Elevation, HGL End HG Elevation (21). Record the HGL Clearance
Slope, and HGL Fall. (20).

Where a pipe is flowing full, the full-flow friction slope


recorded in Step 15 is the Hydraulic Grade Line Slope
(32). The HG Fall (27) is calculated by multiplying the
HGL Slope by the pipe length.

Where a pipe is flowing partially full and the Upper-End


HG is based on normal depth, as in Figure 6.5-2 C and
D, the HG Slope and Fall recorded in Step 15 are not
correct. Here, the HG Slope and Fall are not critical to
the design process, but their values can be recorded
as:
HG Fall (27) = Upper-End HG – Lower-End HG
HG Slope (32) = HG Fall/pipe length
Repeat Step 20 and Step 21 for the entire system. For the first pass through the
system, you may want to calculate the HGL elevation only along the main line from
the outlet to the critical inlet. The flow rates and friction losses in the stub lines
usually are small. Calculating the HG through the entire system (i.e., all the stubs)
for the first iteration is acceptable, but may result in extra effort. For subsequent
passes, calculate the HGL elevation for the entire system.

Chapter 6: Storm Drains 6-61


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

COMPARE HYDRAULIC GRADE LINE (HGL) ELEVATION TO STANDARD

22. Compare the HGL The current standard requires that the Hydraulic
Elevation to the standard Gradient be at least one foot below the inlet elevations.
and adjust pipe diameters.
For systems where the distance between the Hydraulic
Gradient and the gutter elevation is greater than the
standard, the diameter of one or more pipe segments
may be reduced to raise the Hydraulic Gradient. Here,
try to reduce the larger-diameter pipe segments first,
since this will provide a greater cost reduction than
reducing the size of the smaller-diameter segments.
For systems where the distance between the Hydraulic
Gradient and the gutter elevation is less than the
standard, you will need to increase the diameter of one
or more pipe segments to lower the Hydraulic Gradient.
Here, increase the smaller-diameter pipe segments
first, since this will provide less of a cost increase than
increasing the size of the larger-diameter pipe
segments. Look for “flow-pipe size” combinations that
have substantial friction losses. For example, there is
very little reduction in the losses by increasing the
diameter of a 24-inch pipe that is carrying only 3 cfs.
Alternatively, if another 24-inch pipe were carrying 15
cfs, increasing the pipe diameter could achieve a
significant reduction in friction losses.

RECALCULATE THE RUNOFF AND HYDRAULIC GRADE LINE ELEVATION

23. Return to Step 14, When you have made enough iterations through the
working the changes system that any changes in diameters of pipe
through the system. segments would cause the distance between the
Hydraulic Gradient and the gutter elevation to be less
than the standard, your design is essentially complete.

Note:
Examples 6.6-1 and 6.6-2 were created before the Plan Preparation Manual (Volume 2,
Chapter 1.3) required that flow lines be shown to two decimal places and before the
Drainage Manual required that HGL Clearance be provided in the storm tab. The
examples have not been revised to reflect these changes. Although the examples have
not been revised, they still represent a valid design procedure.

Chapter 6: Storm Drains 6-62


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Example 6.6-1 Flat System—Determining Appropriate Pipe Sizes


Given:
• Inlets, Pipes, Runoff Coefficients & Details shown in Figure 6.6-1 and Table 6.6-1
• System discharges to a pond that stages to elevation 8.3 during a three-year design
storm

Figure 6.6-1: Example 6.6-1 – Given Information

Chapter 6: Storm Drains 6-63


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Table 6.6-1
LOCATION OF UPPER DRAINAGE AREA NOTES AND REMARKS

STRUCTURE
END (ac) ZONE: 6

STRUCTURE

LENGTH (ft)
ALIGNMENT NAME C = 0.95 FREQUENCY (yrs): 3

TYPE OF
NO.
C= INLET ELEV MANNING’S “n”: 0.012
DISTANCE
STATION

C = 0.20 (ft) TAILWATER EL. (ft) 8.3


SIDE
(ft )

All overland tc < 10 min.


UPPER INCREMENT TOTAL
LOWER
0.3
Ricardo Way 1
P5 110 10.90
40 + 80 46.5 R 2 0.05
0.2
Ricardo Way 2
P5 200 11.10
41 + 25 46.5 L 3 0.03
---
Ricardo Way 3
MH 200 11.40
43 + 25 44 L 4 ---
0.4
Ricardo Way 4
P5 100 11.10
45 + 25 46.5 L 7 0.1
0.4
Ricardo Way 5
P5 110 10.90
46 + 00 46.5 R 7 0.5
---
Frank Blvd 7
MH 25 10.50
30 + 50 10 R 8 ---
0.15
Frank Blvd 6
P5 32 9.60
30 + 75 16 L 8 0.5
0.25
Frank Blvd 8
P5 250 9.60
30 + 75 16 R outlet 0.5

Find:
• The pipe sizes to meet the standard hydraulic gradient clearance of 1.13 feet to the
inlet (edge of pavement) elevation

1. Add the areas that contribute flow to each pipe segment. For each of the pipe
segments, the total impervious area (C = 0.95) is obtained as follows.

Total Area P1-2 = Inc. AreaS-1; no upstream pipes


= 0.3 ac.
Total Area P2-3 = Inc. AreaS-2 + Total Area P1-2
= 0.2 + 0.3 = 0.5 ac
Total Area P3-4 = Inc. AreaS-3 + Total Area P2-3
= 0.0 + 0.5 = 0.5 ac
Total Area P4-7 = Inc. AreaS-4 + Total Area P3-4
= 0.4 + 0.5 = 0.9 ac

Chapter 6: Storm Drains 6-64


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Total Area P5-7 = Inc. AreaS-5; no upstream pipes


= 0.4 ac.
Total Area P7-8 = Inc. AreaS-7 + Tot. Area P4-7 + Tot.
Area P5-7
= 0.0 + 0.9 + 0.4 = 1.3 ac
Total Area P6-8 = Inc. AreaS-6; no upstream pipes
= 0.15 ac.
Total Area P 8-out= Inc. AreaS-8 + Tot. Area P7-8 + Tot.
Area P6-8
= 0.25 + 1.3 + 0.15 = 1.7 ac

The same approach is applied to drainage areas associated with the pervious runoff
coefficient. Table 6.6-2 is a partial tabulation form with the above information.

Chapter 6: Storm Drains 6-65


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Table 6.6-2
DRAINAGE AREA

STRUCTURE
(ac.)
C= 0.95

NO.
C=
C= 0.20

UPPER INCREMENT TOTAL


LOWER
0.3 0.3
1
2 0.05 0.05
0.2 0.5
2
3 0.03 0.08
--- 0.5
3
4 --- 0.08
0.4 0.9
4
7 0.1 0.18
0.4 0.4
5
7 0.5 0.5
--- 1.3
7
8 --- 0.68
0.15 0.15
6
8 0.5 0.5
0.25 1.7
8
outlet 0.5 1.68

2. For each pipe section, multiply the total area associated with each runoff coefficient
by the corresponding runoff coefficient to obtain the subtotal CA values.
P 1-2: 0.3 x 0.95 = 0.29
0.05 x 0.20 = 0.01
P 2-3: 0.5 x 0.95 = 0.48
0.08 x 0.20 = 0.02
Etc.
3. For each pipe section, add the subtotal CA values to obtain the Total CA.
P 1-2: 0.29 +0.01 = 0.3
P 2-3: 0.48 + 0.02 = 0.5
Etc.

Chapter 6: Storm Drains 6-66


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Table 6.6-3 is a partial tabulation form with the above information.

ESTIMATE A PRELIMINARY HYDRAULIC GRADE LINE (HGL) SLOPE

4. Determine the design TW.


The crown of the outlet pipe is not known at this time, so we will use the given
information about the stage (8.3 feet) of the stormwater management facility.
5. Assume which inlet will be critical.
For this example, we will assume that S-1 is critical. Elevation S-1 = 10.9 feet
6. Calculate a preliminary HGL slope.
For this example, we will base the preliminary HGL slope on the following formula.

Table 6.6-3
STRUCTURE
NO.

DRAINAGE AREA (ac.)

SUB-TOTAL

TOTAL
C •A

C •A
UPPE C= 0.95
R C=
C= 0.20
LOW
INCREMENT TOTAL
ER
0.3 0.3 0.29
1
0.30
2 0.05 0.05 0.01
0.2 0.5 0.48
2
0.50
3 0.03 0.08 0.02
--- 0.5 0.48
3
0.50
4 --- 0.08 0.02
0.4 0.9 0.86
4
0.9
7 0.1 0.18 0.04
0.4 0.4 0.38
5
0.48
7 0.5 0.5 0.1
--- 1.3 1.24
7
1.38
8 --- 0.68 0.14
0.15 0.15 0.14
6
0.24
8 0.5 0.5 0.1
0.25 1.7 1.62
8
1.96
outlet 0.5 1.68 0.34

Chapter 6: Storm Drains 6-67


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Critical Inlet Elev - DTW - 1 foot


Slope =
System Length between Outlet & Critical Inlet

The total pipe length is best seen in Figure 6.6-1.


110 + 200 + 200 + 100 + 25 + 250 = 885 feet
Prelim HGL Slope = 10.9 – 8.3 – 1/885
= 0.0018 ft/ft
= 0.18%

CALCULATE RUNOFF FLOW RATES (FIRST PASS DOWN THE SYSTEM)

Starting with pipe section P1-2,


7. Determine the time of concentration. [P1-2]
Since this is the first inlet in the system and it has an overland tc of 10 minutes or
less, use 10-minute minimum.

8. Determine the intensity. [P1-2]


From the IDF curve for Zone 6, the 10-minute intensity is 6.5 in/hr.

9. Calculate the total runoff for the pipe section. [P1-2]


Q = Total CA (Step 3) times the intensity (previous step)
Q = 0.3 x 6.5 = 1.95 cfs

10. Determine pipe size. [P1-2]


For the first pass, we assume full flow.
Using the hydraulic calculator, an 18-inch pipe is acceptable because the friction
slope (0.03 percent) is flatter than the preliminary HGL slope (0.18 percent). The
minimum physical slope is 0.15 percent (see discussion in Chapter 6.4.2.1).
Record the pipe size, and the minimum physical slope. Also, record the full-flow
friction slope as the HGL slope. Although it is not necessary to record the HGL
slope at this step, it will be used later when moving up the system and calculating
the hydraulic gradient. It may save time to record this while the hydraulic
calculator is set for the flow and pipe size.

11. Determine the pipe flow lines, physical slope, and fall. [P1-2]
For this example, we will use 4.5 feet clearance between the inlet (edge of
pavement) elevation and the flow line of an 18-inch pipe. (The minimum clearance
for an 18-inch pipe in a precast Type P-5 structure is 4.2 feet. See Table 6.4-5.)
Then:
Upper-End Flow Line = 10.9 – 4.5 = 6.4 feet

For this example, we will assume there are no constraints such as utilities that
would prevent the pipe from being set at the minimum physical pipe slope (0.15

Chapter 6: Storm Drains 6-68


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

percent). Then:
Minimum pipe fall = 110 ft x 0.15% = 0.17 feet
Pipe fall = 0.2 ft (minimum fall rounded up to nearest 0.1 foot)
Physical Slope = 0.2 ft/110 ft = 0.18 percent
Lower End Flow Line = 6.4 – 0.2 = 6.2 feet

If this were an actual project, you also should check that the minimum
cover heights in Appendix C of the Drainage Manual are satisfied. To
simplify this example, we will assume that adequate cover is provided.

12. Calculate the actual flow velocity. [P1-2]


Vel = Q/A = 1.95 cfs/πD2/4)
The full-flow cross sectional area is used for the first pass down the system. This
is reasonable for a flat system like this example. If you know the pipe is flowing
partially full, use the average cross sectional flow area. See the discussion on
page 56 and the next example.

Using the hydraulic calculator, the velocity of 1.95 cfs flowing full through an 18-
inch pipe is 1.1 fps.

Calculate the physical velocity. [P1-2]


Using the hydraulic calculator, the full-flow velocity for an 18-inch pipe sloped at
0.18 percent = 2.7 fps

13. Calculate the time of flow in pipe section. [P1-2]


Time = Length/Actual Velocity
= 110 ft/1.1 fps = 100 seconds = 1.7 minutes

A partially completed tabulation form is shown in Table 6.6-4.

Chapter 6: Storm Drains 6-69


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Table 6.6-4

HYD. GRADIENT PIPE

VELOCITY
ACTUAL
TYPE OF STRUCTURE CROWN SIZE

(FPS)
STRUCTURE

FLOW LINE (IN.) SLOPE

TIME OF CONCEN-

TOTAL FLOW (cfs)


TIME OF FLOW IN
NUMBER

INTENSITY (iph)
(%)

INLET ELEV. (ft)


SECTION (min)
TRATION (min)

TOTAL C • A
LENGTH (ft)
RISE

LOWER END
UPPER END
ELEV. (ft)

ELEV. (ft)

FALL (ft)
HYD.

PHYSICAL
VELOCITY
GRAD.

(FPS)
UPPER SPAN PHYSICAL
MIN.
LOWER
PHYS.
.03
1 18 1.1
P5 110 10 1.7 6.5 0.30 1.95 10.90 7.9 7.7 .18
0.2
2 6.4 6.2 18 .15 2.7

For pipe section P2-3,


14. Determine the Time of Concentration. [P2-3]
tc overland ≤ 10 min.
tc system = 10 + 1.7 = 11.7 min. therefore
tc = 11.7 min.

15. Determine the intensity. [P2-3]


From the IDF curve for Zone 6, the 11.7-minute intensity is 6.1 in/hr.

16. Calculate the total runoff for the pipe section. [P2-3]
Flow rate = Total CA x Intensity
= 0.5 x 6.1 = 3.1

17. Determine pipe size. [P2-3]


Using the hydraulic calculator, an 18-inch pipe is acceptable because the friction
slope (0.07 percent) is less than the preliminary HGL slope (0.18 percent). As
done for the previous pipe section, record the pipe size, and the minimum physical
slope. Also, record the friction slope as the HGL slope.

18. Determine the pipe flow lines, physical slope, and fall. [P2-3]
Since this inlet is higher than S-1, the potential conflict with the inlet top will not
control the flow lines. For this example, we will attempt to match flow line
elevations across structures. Therefore:
Upper-end flow line = 6.2 (previous pipe section downstream flow line)
Minimum pipe fall = length x min. phys. slope
= 200 ft x 0.15% = 0.3 ft
Pipe fall = 0.3 ft
Physical slope = 0.3 ft/200 ft = 0.15%
Lower-end flow line = 6.2 – 0.3 = 5.9 feet

Chapter 6: Storm Drains 6-70


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

19. Calculate the actual flow velocity. [P2-3]


Vel = Q/A = 1.95 cfs/(π D2/4).
The full-flow cross sectional area is used for the first pass through the system.
Using the hydraulic calculator, the velocity of 3.1 cfs flowing full through an 18-
inch pipe is approximately 1.8 fps.

Calculate the physical velocity. [P2-3]


Using the hydraulic calculator, the full-flow velocity for an 18-inch pipe sloped at
0.15 percent = 2.5 fps

20. Calculate the time of flow in pipe section. [P2-3]


Time = Length/Actual Velocity
= 200 ft/1.8 fps = 111 seconds = 1.9 minutes

A partially completed tabulation form is shown in Table 6.6-5.

Table 6.6-5
HYD. GRADIENT

VELOCITY
PIPE

ACTUAL
CROWN
STRUCTURE

TYPE OF STRUCTURE

(FPS)
SIZE
FLOW LINE
TIME OF CONCEN-

TOTAL FLOW (cfs)


TIME OF FLOW IN

(IN)
INTENSITY (iph)

INLET ELEV. (ft)


TRATION (min)

SECTION (min)
NO.

TOTAL C • A
LENGTH (ft)

SLOPE (%)
LOWER END
UPPER END
ELEV. (ft)

ELEV. (ft)

RISE
FALL (ft)

PHYSICAL
VELOCITY
UPPER

(FPS)
HYD. GRAD.
SPAN
LOWER PHYSICAL
MIN. PHYS.
0.03
1 18 1.1
P5 110 10 1.7 6.5 0.30 1.95 10.90 7.9 7.7 0.18
0.2
2 6.4 6.2 18 0.15 2.7
0.07
2 18 1.8
P5 200 11.7 1.9 6.1 0.50 3.1 11.10 7.7 7.4 0.15
0.3
3 6.2 5.9 18 0.15 2.5

Step 14 through Step 20 are repeated for the remaining pipe sections. Situations that
are different from the above pipe sections are discussed below. Table 6.6-6 shows the
results of doing these steps for the entire system.

Pipe Section P3-4


The manhole contributes no additional flow to the system, nor does it combine
flow from several pipes. The time of concentration and the intensity are not
applicable. The flow through the pipe section is the same as the upstream pipe
section.

Pipe Section P4-7,


The time of concentration is 11.7 + 1.9 +1.9 = 15.5 minutes.

Chapter 6: Storm Drains 6-71


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Pipe Sections P5-7 and P6-8


They receive only overland flow like pipe section P1-2, thus their time of
concentration is based on overland flow time. Their flow lines are determined by
matching the flow lines of the downstream structure, using the minimum physical
slope to the upstream structure, and rounding to the nearest 0.1' such that the
minimum slope is maintained.

Pipe Section P7-8


This is similar to similar to pipe section P3-4 in that the manhole contributes no
flow. It is different from pipe section P3-4 in that two pipes drain to the manhole.
Because of this difference, the pipe section is treated like the other inlets along
the main line. The tc, intensity, and flow are calculated for the section. The time
of concentration is 15.5 + 0.6 = 16.1 minutes.

As stated in Step 18, we will attempt to match flow lines across structures for this
example. The upper-end flow line is set to match the lower-end flow line of pipe
section P4-7. The lower-end flow line is set to match the flow line of S-8.

Pipe Section P8-out


For this example, we will use a 5.1-foot clearance between the inlet (edge of
pavement) elevation and the flow line of a 24-inch pipe. (The minimum clearance
for a 24-inch pipe in a precast Type P-5 structure is 4.7 feet. See Table 6.4-5.)
Then, upper-end FL = 9.6 – 5.1 = 4.5 feet. The lower-end FL is set to match the
minimum physical slope with the FL rounded to the closest 0.1 foot such that the
minimum slope is maintained.

Chapter 6: Storm Drains 6-72


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Table 6.6-6: Results of First Pass Down the System


HYD.
GRADIENT PIPE

VELOCITY
ACTUAL
SIZE
TYPE OF STRUCTURE
STRUCTURE

CROWN

(FPS)
TIME OF CONCEN-

TOTAL FLOW (cfs)


(IN)

TIME OF FLOW IN

INTENSITY (iph)
FLOW LINE

INLETELEV. (ft)
TRATION (min)

SECTION (min)

TOTAL C • A
NO.

LENGTH (ft)
SLOPE (%)

LOWER END
UPPER END
RISE

ELEV. (ft)

ELEV. (ft)
FALL (ft)

PHYSICAL
VELOCITY
(FPS)
UPPER HYD. GRAD.
SPAN PHYSICAL
LOWER
MIN. PHYS.
.03
1 18 1.1
P5 110 10 1.7 6.5 0.30 1.95 10.90 7.9 7.7 .18
0.2
2 6.4 6.2 18 .15 2.7
0.07
2 18 1.8
P5 200 11.7 1.9 6.1 0.50 3.1 11.10 7.7 7.4 0.15
0.3
3 6.2 5.9 18 0.15 2.5
0.07
3 18 1.8
MH 200 N/A 1.9 N/A 0.50 3.1 11.40 7.4 7.1 0.15
0.3
4 5.9 5.6 18 0.15 2.5
0.18
4 18 2.8
P5 100 15.5 0.6 5.4 0.9 4.9 11.10 7.1 6.9 0.2
0.2
7 5.6 5.4 18 0.15 2.9
0.07
5 18 1.8
P5 110 10 1.7 6.5 0.48 3.1 10.90 7.1 6.9 0.18
0.2
7 5.6 5.4 18 0.15 2.7
0.09
7 24 2.4
MH 25 16.1 0.2 5.3 1.38 7.3 10.50 7.4 6.5 3.6
0.9
8 5.4 4.5 24 0.1 14.8
0.02
6 18 1.3
P5 32 10 0.4 6.5 0.24 1.56 9.60 6.1 6.0 0.3
0.1
8 4.6 4.5 18 0.15 3.5
0.18
8 24 3.4
P5 250 16.3 1.9 5.3 1.96 10.4 9.60 6.5 6.2 0.12
0.3
outlet 4.5 4.2 24 0.1 2.7

Chapter 6: Storm Drains 6-73


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

21. Check for peak flow from reduced area.


Reviewing the size and runoff coefficient for each drainage area, it does not
appear that most of the area or most of the imperviousness is concentrated near
the lower end of the system. Since we would not expect to have peak flow from
reduced area, detailed calculations would not be necessary. For this example, we
will check it just to demonstrate an approach.

From the schematic, it


appears that a logical Inlet
reduced area would be area Manhole
32'
S8
flowing overland to S4, S5, S6 25'
S6, and S8. The overland tc 200' 200' S7
S2
to S4 was given as 10 S3 S4
minutes. So let’s apply a 10- Ricardo Way
minute storm to pipe section S5
P4-7. Doing so reduces the S1
contributing area from S3. Schematic

An approach to finding the


reduced contributing area is to multiply the area (or the CA product) from S3 by
the ratio of the times of concentration. From Table 6.6-6, the tc for the flow from S3
is 15.5 minutes. So, reduce the Total CA from S3 by the ratio 10/15.5, or 0.65.

From Table 6.6-6, the Total CA from S3 = 0.5.

So the Total CA from S3 is reduced by: 0.5 (1 - 0.65) = 0.18 This value is not the reduced
CA; it is the amount the Total
CA is reduced by.

Reducing the Total CA for pipe section P4-7 by this amount yields: 0.9 – 0.18 = 0.72

The three-year intensity for a 10-minute storm = 6.5 in/hour.

The flow in the pipe downstream of S4 = CAi = 0.72 x 6.5 = 4.7 cfs.

This is less than the 4.9 cfs calculated for the entire contributing area for P4-7, as
shown in Table 6.6-6. So, peak flow in pipe section P4-7 does not result from
reduced area. Although other pipe sections could be checked for peak flow from
reduced area, this effort shown above is acceptable for this system.

Chapter 6: Storm Drains 6-74


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

CALCULATE THE HYDRAULIC GRADE LINE ELEVATION (first pass up the system)

For pipe section P8-out:


22. Determine the lower-end HG elevation. [P8-out]
For the outlet pipe, the lower-end HG is the design tailwater. For this example, the
design tailwater is the higher of (1) the crown of the pipe (elev. 6.2 feet), or (2) the
peak stage of the stormwater facility during the storm drain design event (elev. 8.3
feet). Thus, the lower-end HG = 8.3 feet

23. Determine the upper-end HG elevation. [P8-out]


For this example, the lower-end HG submerges the entire pipe section; therefore,

Upper-end HG elev. = Lower-end HG elev. + Full-flow friction loss


& Full-Flow Friction Loss = Full-flow friction slope x Pipe length
= 0.18% x 250 ft = 0.45 feet

The full-flow friction slope was previously recorded as the hydraulic gradient slope
in Table 6.6-6 when we moved down the system calculating flow rates.

Then upper-end HG elev. = Lower-end HG elev. + Full-flow friction loss


= 8.3 + 0.45 = 8.75 feet (see table below)

Table 6.6-7
HYD. GRADIENT
STRUCTURE

CROWN
NO.

LENGTH (ft)

FLOW LINE SLOPE (%)

UPPER LOWER
UPPER ELEV ELEV FALL HYD. GRAD.
(ft)
(ft) (ft) PHYSICAL
LOWER
MIN. PHYS.

8.75 8.3 0.45 0.18


8
250 6.5 6.2 0.12
0.3
outlet 4.5 4.2 0.1

Pipe sections P6-8 and P5-7 are stubs and their hydraulic gradient will not be
calculated during the first pass up the system.

Chapter 6: Storm Drains 6-75


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

For pipe section P7-8:

24. Determine the lower-end HG elevation.


The downstream pipe upper-end HG elevation (8.75 feet) is higher than the lower-
end Crown Elevation (6.5 feet); therefore, the lower-end HG elevation =
downstream pipe upper-end HG = 8.75 feet

25. Determine the upper-end HG elevation. [P7-8]


For this example, the lower-end HG submerges the entire pipe section; therefore:
Upper-end HG elev. = Lower-end HG elev. + Full-flow friction loss
Full-flow friction loss = Full-flow friction slope x Pipe length
= 0.09% x 25 ft = 0.02 feet
The full-flow friction slope was recorded previously as the hydraulic gradient slope
in Table 6-6.
Then, upper-end HG elev. = Lower-end HG elev. + Full-flow friction loss
= 8.75 + 0.02 = 8.77 feet

The same steps are repeated for the remaining mainline pipe sections. Table 6.6-8 shows
the results of doing these steps for the entire system.

Chapter 6: Storm Drains 6-76


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Table 6.6-8: Results of First Pass Up the System

HYD. GRADIENT PIPE


SIZE
CROWN (IN)
STRUCTURE

FLOW LINE

INLET ELEV. (ft)


SLOPE (%)

TOTAL FLOW
LENGTH (ft)
NO.

RISE

(cfs) UPPER LOWER


FALL
END END
(ft) HYD. GRAD.
ELEV ELEV
(ft) (ft) SPAN
PHYSICAL
UPPER
LOWER MIN. PHYS.

9.26 9.23 0.03 0.03


1 18
110 1.95 10.90 7.9 7.7 0.18
0.2
2 6.4 6.2 18 0.15
9.23 9.09 0.14 0.07
2 18
200 3.1 11.10 7.7 7.4 0.15
0.3
3 6.2 5.9 18 0.15
9.09 8.95 0.14 18 0.07
3
200 3.1 11.40 7.4 7.1 0.15
0.3 18
4 5.9 5.6 0.15
8.95 8.77 0.18 18 0.18
4
100 4.9 11.10 7.1 6.9 0.20
0.2 18
7 5.6 5.4 0.15
0.07
5 18
110 3.1 10.90 7.1 6.9 0.18
0.2
7 5.6 5.4 18 0.15
8.77 8.75 0.02 0.09
7 24
25 7.3 10.50 7.4 6.5 3.60
0.9
8 5.4 4.5 24 0.10
0.02
6 18
32 1.56 9.60 6.1 6.0 0.30
0.1
8 4.6 4.5 18 0.15
8.75 8.3 .45 0.18
8 24
250 10.4 9.60 6.5 6.2 0.12
0.3
outlet 4.5 4.2 24 0.10

26. Compare the hydraulic gradient to the standard and adjust pipe sizes.
The standard clearance of 1.13 feet between the hydraulic gradient and the inlet
elevation (edge of pavement) is not met for S-8 and probably not met for S-6. The
remaining inlets have adequate clearance. Increasing the size of the outlet pipe
P8-out to 30 inches will reduce the hydraulic gradient at S-6 and S-8, so we will try
that. To reduce costs, we also will try reducing the pipe size of section P7-8 to 18
inches.

27. Calculate the hydraulic gradient using the changed pipe sizes.
Table 6.6-9 shows the new slopes and the recalculated hydraulic gradient for the
entire system.

Chapter 6: Storm Drains 6-77


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Table 6.6-9: Results of Second Pass Up the System


HYD. GRADIENT PIPE
SIZE
STRUCTURE

CROWN (IN)

INLET ELEV. (ft)


TOTAL FLOW
SLOPE (%)
LENGTH (ft)
NO.

FLOW LINE

(cfs)
RISE
UPPER LOWER HYD. GRAD.
END END FALL
ELEV ELEV (ft) PHYSICAL
SPAN
UPPER (ft) (ft)
MIN. PHYS.
LOWER
9.04 9.01 0.03 0.03
1 18
110 1.95 10.90 7.9 7.7 0.18
0.20
2 6.4 6.2 18 0.15
9.01 8.87 0.14 0.07
2 18
200 3.1 11.10 7.7 7.4 0.15
0.30
3 6.2 5.9 18 0.15
8.87 8.73 0.14 0.07
3 18
200 3.1 11.40 7.4 7.1 0.15
0.30
4 5.9 5.6 18 0.15
8.73 8.55 0.18 0.18
4 18
100 4.9 11.10 7.1 6.9 0.20
0.20
7 5.6 5.4 18 0.15
8.63 8.55 0.08 0.07
5 18
110 3.1 10.90 7.1 6.9 0.18
0.20
7 5.6 5.4 18 0.15
8.55 8.45 0.10 0.40
7 18
25 7.3 10.50 6.9 6.0 3.60
0.90
8 5.4 4.5 18 0.15
8.46 8.45 0.01 0.02
6 18
32 1.56 9.60 6.1 6.0 0.30
0.10
8 4.6 4.5 18 0.15
8.45 8.3 0.15 0.06
8 30
250 10.4 9.60 7.0 6.8 0.08
0.20
outlet 4.5 4.3 30 0.08

28. Compare the HG to the standard and adjust pipe sizes.


From Table 6.6-9, the standard 1.13 feet of clearance between the hydraulic
gradient and the inlet elevation (edge of pavement) exists throughout the system.

29. Recalculate the flow.


Changing pipe sizes changes the velocity, thus changing the time of flow in the
section and the time of concentration. These changes affect only the changed
pipes and the pipes downstream of the changed pipes. For this example, only
pipe sections P7-8 and P8-out are affected.

The increased velocity in pipe section P7-8 reduced the time of flow in the pipe by
only 0.1 minute because the pipe is so short. As a result, the time of concentration
of the outlet pipe was reduced by only 0.1 minute from 16.3 to 16.2 minutes. It is
hard to read a change in the intensity from the IDF curve for a change in tc of 0.1

Chapter 6: Storm Drains 6-78


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

minute. Although we changed the size of the two pipes, there was no noticeable
change to the flow rate in the system.
A completed tabulation form is shown in Table 6.6-10.

Chapter 6: Storm Drains 6-79


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Example 6.6-1
Table 6.6-10

VELOCITY(FPS)
LOCATION DRAINAGE AREA HYD. GRADIENT PIPE NOTES AND REMARKS

TIME OF FLOW IN SECTION


TIME OF CONCENTRATION
OF (ac) CROWN SIZE ZONE: 6

ACTUAL
TYPE OF STRUCTURE
UPPER END (IN) SLOPE (%)
STRUCTURE

C • A SUBTOTAL
C = 0.95 FLOW LINE FREQUENCY (Yrs) : 3
ALIGNMENT NAME

C • A TOTAL
TOTAL FLOW
LENGTH (ft)

INTENSITY
NO.
C= RISE MANNING’S “n” : 0.012
INLET

(MIN)

(MIN)

(iph)

(cfs)
C = 0.20 ELEV. HYD. TAILWATER EL. (ft): 8.3
DISTANCE

(ft) UPPER LOWER GRAD


All overland tc < 10 Min.

PHYSICAL
VELOCITY
FALL
SIDE

END END
(ft)

(ft)
STATION

(FPS)
INCRE- ELEV ELEV PHYSICAL
TOTAL SPAN
MENT (ft) (ft)
UPPER
MIN.
LOWER PHYS.
0.3 0.3 0.29 9.04 9.01 0.03 0.03
Ricardo Way 1 18 1.1
P5 110 10 1.7 6.5 0.3 1.95 10.90 7.9 7.7 0.18
0.2
40 + 80 46.5 R 2 0.05 0.05 0.01 6.4 6.2 0.15 2.7
0.2 0.5 0.48 9.01 8.87 0.14 18 0.07
Ricardo Way 2 1.8
P5 200 11.7 1.9 6.1 0.5 3.1 11.10 7.7 7.4 0.15
0.3
41 + 25 46.5 L 3 0.03 0.08 0.02 6.2 5.9 0.15 2.5
--- 0.5 0.48 8.87 8.73 0.14 0.07
Ricardo Way 3 18 1.8
MH 200 N/A 1.9 N/A 0.5 3.1 11.40 7.4 7.1 0.15
0.3
43 + 25 44 L 4 --- 0.08 0.02 5.9 5.6 0.15 2.5
0.4 0.9 0.86 8.73 8.55 0.18 0.18
Ricardo Way 4 18 2.8
P5 100 15.5 0.6 5.4 0.9 4.9 11.10 7.1 6.9 0.2
0.2
45 + 25 46.5 L 7 0.1 0.18 0.04 5.6 5.4 0.15 2.9
0.4 0.4 0.38 8.63 8.55 0.08 0.07
Ricardo Way 5 18 1.8
P5 110 10 1.0 6.5 0.48 3.1 10.90 7.1 6.9 0.18
0.2
46 + 00 46.5 R 7 0.5 0.5 0.1 5.6 5.4 0.15 2.7
--- 1.3 1.24 8.55 8.45 0.1 0.40
Frank Blvd. 7 18 4.1
MH 25 16.1 0.1 5.3 1.38 7.3 10.50 6.9 6.0 3.6
0.9
30 + 50 10 R 8 --- 0.68 0.14 5.4 4.5 0.15 12.3
0.15 0.15 0.14 8.46 8.45 0.01 0.02
Frank Blvd. 6 18 1.3
P5 32 10 0.4 6.5 0.24 1.56 9.60 6.1 6.0 0.3
0.1
30 + 75 16 L 8 0.5 0.5 0.1 4.6 4.5 0.15 3.5
0.25 1.7 1.62 8.45 8.3 0.15 0.06
Frank Blvd 8 30 2.2
P5 250 16.2 1.9 5.3 1.96 10.4 9.60 7.0 6.8 0.08
0.2
30 + 75 16 R Outlet 0.5 1.68 0.34 4.5 4.3 0.08 2.5

Chapter 6: Storm Drains 6-80


January 2019

Drainage Design
Guide

Chapter 6: Storm Drains

Example 6.6-2 Steep System—Determining Appropriate Pipe Sizes


Given:
• Inlets, pipes, runoff coefficients, and details in Figure 6.6-6 and Table 6.6-11
• Designer chooses to match crown elevations across structures

Figure 6.6-6

Chapter 6: Storm Drains 6-81


January 2019

Drainage Design
Guide

Chapter 6: Storm Drains

Table 6.6-11
LOCATION
DRAINAGE AREA
OF NOTES AND REMARKS

TYPE OF STRUCTURE
(ac)
UPPER END

STRUCTURE
ALIGNMENT NAME C= INLET
ZONE: 7

NO.
ELEV.

LENGTH
C = 0.80
(ft)

(ft)
DISTANCE

C= FREQUENCY (Yrs): 3
SIDE

MANNING’S “n”: 0.012


(ft)

STATION
INCRE- TAILWATER EL. (ft): 47.1
TOTAL
UPPER MENT All overland tc < 10 min.
LOWER

Patrick Place 15
P1 300 0.2 59.70
14

Patrick Place 14
P1 300 1.0 59.80
13

Patrick Place 13
P1 300 0.6 59.00
12

Patrick Place 12
P1 300 0.5 54.50
11

Patrick Place 11
P1 300 1.1 50.50
outlet

Example 6.6-2—Given Information


Find:
● The pipe sizes to meet the standard hydraulic gradient clearance of 1.13 feet to
the inlet (edge of pavement) elevation.

1. Add the areas that contribute flow to each pipe segment. For each of the pipe
segments, the total area is obtained as in Example 6.6-1.

2. For each pipe section, multiply the total area associated with each runoff
coefficient by the corresponding runoff coefficient to obtain the subtotal CA
values.

3. For each pipe section, add the subtotal CA values to obtain the total CA value.

Table 6.6-12 is a partial tabulation form complete with the information from the first three
steps.

Chapter 6: Storm Drains 6-82


January 2019

Drainage Design
Guide

Chapter 6: Storm Drains

Table 6.6-12
DRAINAGE AREA
(ac)

STRUCTURE
C=

TOTAL C • A
NO.
C = 0.80
SUB-
C= TOTAL
C•A
UPPER INCRE-
TOTAL
MENT
LOWER

15
0.2 0.2 0.16 0.16
14

14
1 1.2 0.96 0.96
13

13
0.6 1.8 1.44 1.44
12

12
0.5 2.3 1.84 1.84
11

11
1.1 3.4 2.72 2.72
outlet

ESTIMATE A PRELIMINARY HGL SLOPE

4. Determine the design TW.


The crown of the outlet pipe is not known at this time, so we will use the given
information about the lake stage. DTW = 47.1 feet.

5. Assume which inlet will be critical. For this example, we will assume that S-15 is
critical.

6. For this example, we will base the preliminary HGL slope on the following formula.

Critical Inlet Elev - DTW - 1 foot


Slope =
System Length between Outlet & Critical Inlet
Slope = (59.7 – 47.1 – 1)/1,500 = 0.8%

Chapter 6: Storm Drains 6-83


January 2019

Drainage Design
Guide

Chapter 6: Storm Drains

CALCULATE RUNOFF FLOW RATES


FIRST PASS DOWN THE SYSTEM

Starting with pipe section P15-14:

7. Determine the time of concentration. [P15-14]


Since this is the first inlet in the system and it has an overland tc of 10 minutes or
less, use the 10-minute minimum.

8. Determine the intensity. [P15-14]


From the IDF curve for Zone 7, the 10-minute intensity is 6.5 in/hr.

9. Calculate the flow rate for the pipe section. [P15-14]


Q = Total CA (Step 3) times the intensity (previous step)
Q = 0.16 x 6.5 = 1.04 cfs

10. Determine pipe size. [P15-14]


For the first pass, we are assuming full flow.
Using the hydraulic calculator, an 18-inch pipe is acceptable because the friction
slope (<0.04 percent) is flatter than the preliminary HGL slope. The minimum
physical slope is 0.15 percent (see discussion in Section 6.4.2.1). Record the pipe
size and the minimum physical slope. Also, record the full-flow friction slope as the
HGL slope. For this flow rate through an 18-inch pipe, the friction loss is so small
that the Department’s hydraulic calculator does not show the slope. The loss could
be calculated from the equation in Section 6.5.1, but for now we will record the
HGL slope as zero.

11. Determine the pipe flow lines, physical slope, and fall. [P15-14]
For this example, we will use a 4.5-foot clearance between the inlet (edge of
pavement) elevation and the flow line of an 18-inch pipe. (The minimum clearance
for standard precast structures is 4.2 feet. See Table 6.4-5.) Then:
Upper-end flow line = 59.7 – 4.5 = 55.2 feet

For this example, we will assume there are no constraints such as utilities that
would prevent the pipe from being set at the minimum physical pipe slope (0.15
percent). Then:
Minimum pipe fall = 300 ft x 0.15% = 0.45 ft
Pipe fall = 0.5 ft (minimum fall rounded up to nearest 0.1 foot)
Physical Slope = 0.5 ft/300 ft = 0.167%
Lower-end flow line = 55.2 – 0.5 = 54.7 feet

Chapter 6: Storm Drains 6-84


January 2019

Drainage Design
Guide

Chapter 6: Storm Drains

If this were an actual project, you also should check that the minimum
cover heights in Appendix C of the Drainage Manual are satisfied. To
simplify this example, we will assume that adequate cover is provided.

12. Calculate the actual flow velocity. [P15-14]


V = Q/A = 1.04 cfs/(π D2/4)

The full-flow cross sectional area used for the first pass through the system. Using
the hydraulic calculator, the velocity of 1.04 cfs flowing full through an 18-inch pipe
is 0.59 fps.

Calculate the physical velocity. [P15-14]

Using the hydraulic calculator, the full-flow velocity for an 18-inch pipe sloped at
0.17 percent = 2.6 fps

13. Calculate the time of flow in the pipe section. [P15-14]


Time = Length/Actual Velocity
= 300 ft/0.59 fps = 508 seconds = 8.5 minutes

A partially completed tabulation form with the information from this pipe is shown below.

Chapter 6: Storm Drains 6-85


January 2019

Drainage Design
Guide

Chapter 6: Storm Drains

Table 6.6-13
HYD. GRADIENT PIPE

VELOCTIY
STRUCTURE

CROWN SIZE

ACTUAL

(FPS)
TIME OF CONCEN-

TIME OF FLOW IN
FLOW LINE (IN)

SECTION (MIN)
TRATION (MIN)
NO.

TOTAL FLOW
TOTAL C • A
SLOPE (%)

INTENSITY
LENGTH

INLET RISE

(iph)

(cfs)
(ft)

ELEV. UPPER LOWER

FALL
(ft) END END

CELOCITY
PHYSICAL
(ft)
UPPER ELEV ELEV HYD. GRAD.

(FPS)
(ft) (ft)
SPAN PHYSICAL
LOWER
MIN. PHYS.
0
15 18 .59
300 10 8.5 6.5 0.16 1.04 59.70 56.7 56.2 .167
0.5
14 55.2 54.7 18 .15 2.6

Step 7 through Step 13 are repeated for the remaining pipe sections. We have assumed
that all the pipes are flowing full for this pass down the system. Situations that are
different from the above pipe section are discussed below.

For Pipe Section P14-13:


• The time of concentration is 10 + 8.5 = 18.5 minutes
• The upper-end flow line = 54.7 (set by matching the crowns across the structure)
• The lower-end flow line = 54.7 – 0.5 = 54.2 (set by minimum pipe slope as was done
for P15-14)

For Pipe Section P13-12:


• The time of concentration is 18.5 + 1.8 = 20.3 minutes
• The upper-end flow line = 54.2 (set by matching the crowns across the structure)
• The lower-end flow line = S12 gutter elev. – inlet clearance = 54.5 – 4.5 = 50.0
• The physical slope = (54.2 – 50.0)/300 = 1.4 percent

For Pipe Section P12-11:


• The time of concentration is 20.3 + 1.3 = 21.6 minutes.
• The upper-end flow line = 50.0 (set by matching the crowns across the structure)
• The lower-end flow line = S11 gutter elev. – inlet clearance = 50.5 – 4.5 = 46.0
• The physical slope = (50.0 – 46.0)/300 = 1.33 percent

For Pipe Section P11-out:


• Size could be 18-inch pipe or 24-inch pipe based on comparing the full-flow
friction loss slope to the preliminary HGL slope. Try 18-inch pipe, since the other
pipes seem oversized.

Chapter 6: Storm Drains 6-86


January 2019

Drainage Design
Guide

Chapter 6: Storm Drains

• The time of concentration is 21.6 + 1.0 = 22.6 minutes


• The upper-end flow line = 46.0 (set by matching the crowns across the structure).
• Several factors may control the lower-end flow line, such as but not limited to
cover requirements under roads around the lake, the lake bottom elevation,
purposely submerging the outlet to minimize potential erosion at the outlet. For
this example, we arbitrarily chose 44.5 feet.
• The physical slope = (46.0 – 44.5)/300 = 0.5 percent

The full-flow friction slope was recorded as the hydraulic gradient slope for all the pipes.
Table 6.6-14 shows the results of doing Step 7 through Step 13 for the entire system.

Table 6.6-14: Results of First Pass Down the System


HYD. GRADIENT PIPE

VELOCITY
CROWN SIZE

ACTUAL

(FPS)
TIME OF CONCEN-

FLOW LINE (IN)


STRUCTURE

TIME OF FLOW IN
SECTION ( MIN )
TRATION ( MIN )

TOTAL FLOW
TOTAL C•A

SLOPE (%)
INTENSITY
NO.

LENGTH

INLET
( iph )

( cfs )

RISE
(ft)

ELEV UPPER LOWER

FALL
(ft ) END END

( ft )

PHYSICAL
VELOCITY
ELEV ELEV

(FPS)
( ft ) ( ft ) HYD GRAD
UPPER SPAN PHYSICAL
LOWER MIN PHYS
≈0
15 18 .59
300 10 8.5 6.5 0.16 1.04 59.70 56.7 56.2 0.167
0.5
14 55.2 54.7 18 .15 2.6
0.18
14 18 2.8
300 18.5 1.8 5.1 0.96 4.9 59.80 56.2 55.7 0.167
0.5
13 54.7 54.2 18 0.15 2.6
0.38
13 18 4.0
300 20.3 1.3 4.9 1.44 7.05 59.00 55.7 51.5 1.4
4.2
12 54.2 50.0 18 0.15 7.6
0.58
12 18 4.9
300 21.6 1.0 4.7 1.84 8.65 54.50 51.5 47.5 1.33
4.0
11 50.0 46.0 18 0.15 7.5
1.2
18 7.0
11 300 22.6 - 4.6 2.72 12.5 50.50 47.5 46.0 0.50
1.5
46.0 44.5 18 0.15 4.5

CALCULATE THE HYDRAULIC GRADE LINE ELEVATION

For pipe section P11-out


14. Determine the Lower-end HG elevation. [P11-out]

Chapter 6: Storm Drains 6-87


January 2019

Drainage Design
Guide

Chapter 6: Storm Drains

For the outlet pipe, the lower-end HG is the design tailwater (DTW). For this
example, the design tailwater is the higher of: (1) the crown of the pipe (elev. 46.0
feet), or (2) the normal high water stage (47.1) of the lake. Thus:
Lower-end HG = 47.1 feet

15. Determine the upper-end HG elevation. [P11-out]


The upper-end HG is the higher of:
1) Lower-end HG + full-flow friction loss = 47.1 + 1.2% x 300 feet
= 47.1 + 3.6 = 50.7 feet

OR

2) The elevation of normal depth upstream. The above elevation is higher than
the upper-end crown, so normal depth cannot control.

Then: Upper-end HG = 50.7 feet

The standard HG clearance is not met (S-11 inlet elev. = 50.5). We will increase
this pipe size to 24 inches before continuing upstream. To match the crowns at the
upper end, the flow line of the 24-inch pipe will be set 0.5 foot lower than for the
18-inch pipe. [P11-out]
Upper-end flow line = 45.5 ft
Pipe fall = 45.5 – 44.5 = 1 ft (holding lower-end flow line)
Physical slope = 1/300 = 0.33 percent

Starting at the outlet pipe again:

16. Determine the lower-end HG elevation. [P11-out]


Using the same approach as in Step 14, the lower-end HG elevation = 47.1 feet

17. Determine the upper-end HG elevation. [P11-out]

The Upper End HG is higher of:

1) Lower-end HG + full-flow friction loss = 47.1 + 0.26% x 300 ft


= 47.1 + 0.78 = 47.9 ft

OR

2) The elevation of normal depth upstream. The above elevation (47.9 feet) is
higher than the crown (47.5 feet), so normal depth does not apply.

Chapter 6: Storm Drains 6-88


January 2019

Drainage Design
Guide

Chapter 6: Storm Drains

Then, upper-end HG = 47.9 feet

Repeat Step 16 and Step 17 for the other pipe sections.

Table 6.6-16 shows the results of completing these steps for the entire system.
For Pipe Section P12-11:
The lower-end HG is the higher of:

1) Downstream pipe upper-end HG: = 47.9 ft

OR

2) Controlling pipe condition at the lower end. The above elevation (47.9 feet) is
Table 6.6-15
higher than the crown (47.5 feet), so normal depth does not apply.

Then, lower-end HG = 47.9 feet

The upper-end HG is the higher of:

1) Lower-end HG + full-flow friction loss = 47.9 + 0.58% x 300 ft


= 47.9 + 1.74 = 49.64 ft

OR

2) The elevation of normal depth upstream. Using the hydraulic calculator (Q =


8.65 cfs, 18-inch pipe @ 1.33 percent slope), the normal depth = 0.6 x
Diameter

Upper-end normal depth elev. = 50.0 + 0.6 x 1.5 = 50.9 ft (dNORM = 0.6 x D)

Then, upper-end HG = 50.9 feet

For Pipe Section P13-12:


The lower-end HG is the higher of:

1) Downstream pipe upper-end HG = 50.9 ft

OR

2) Controlling pipe condition at the lower end. The pipe slope (1.4 percent) is

Chapter 6: Storm Drains 6-89


January 2019

Drainage Design
Guide

Chapter 6: Storm Drains

steeper than the full-flow friction slope (0.38 percent) so, if the downstream HG
is low enough, the flow depth at the lower end is normal depth. Thus, the
controlling pipe condition is normal depth (Figure 5-1C). Using the hydraulic
calculator (Q = 7.05 cfs, 18-inch pipe @ 1.4 percent slope) the normal depth =
0.52 x Diameter

Lower-end normal depth elev. = 50.0 + 0.52 x 1.5 = 50.78 ft

Then, lower-end HG = 50.9 feet

The upper-end HG is the higher of:

1) Lower-end HG + full-flow friction loss = 50.9 + .38% x 300 ft


= 50.9 + 1.14 = 52.04 ft

OR

2) The elevation of normal depth upstream. = 54.2 + 0.52 x 1.5 = 54.98 ft


(dNORM = 0.52 x D)

Then, upper-end HG = 54.98 feet

For Pipe Section P14-13:


The lower-end HG is the higher of:

1) Downstream pipe upper end HG = 54.98 ft

OR

2) Controlling pipe condition at the lower end. The pipe slope (0.167 percent) is
flatter than the full-flow friction slope (0.18 percent), so use the crown of the
pipe (Figure 6.5-1D) as the controlling pipe condition at the lower end. Lower-
end crown elev. = 55.7 ft

Then, lower-end HG = 55.7 feet

The upper-end HG is the higher of:

1) Lower-end HG + full-flow friction loss = 55.7 + 0.18% x 300 ft

Chapter 6: Storm Drains 6-90


January 2019

Drainage Design
Guide

Chapter 6: Storm Drains

= 55.7 + 0.54 = 56.24 ft

OR

2) The elevation of normal depth upstream. The above elevation (56.24 feet) is
higher than the crown (56.2 feet), so normal depth does not apply.

Then, upper-end HG = 56.24 feet

For Pipe Section P15-14:


The Lower End HG is the higher of:

1) Downstream pipe upper-end HG = 56.24 ft

OR

2) Controlling pipe condition at the lower end. The above elevation (56.24 feet)
is higher than the crown (56.2 feet), so normal depth does not apply.

Then, lower-end HG = 56.24 feet

The Upper End HG is the higher of:

1) Lower-end HG + full-flow friction loss = 56.24 + 0.0 ft = 56.24 ft

OR

2) The elevation of normal depth upstream. Using the hydraulic calculator (Q =


1.0 cfs, 18-inch pipe @ 0.17 percent slope), the normal depth = 0.32 x
Diameter

Normal depth elevation= 55.2 + 0.32 x 1.5 = 55.68 ft

Then, upper-end HG = 56.24 feet

Table 6.6-16 shows the results of doing Step 16 and Step 17 for the entire system.

Chapter 6: Storm Drains 6-91


January 2019

Drainage Design
Guide

Chapter 6: Storm Drains

Table 6.6-16: Results of First Pass up the System


HYD. GRADIENT PIPE
STRUCTURE

CROWN SIZE
TOTAL FLOW SLOPE (%)
LENGTH (ft)

(IN)
NO.

INLET FLOW LINE


(cfs)
ELEV UPPER LOWER RISE

FALL
(ft) END END HYD. GRAD.

(ft)
ELEV ELEV
UPPER (ft) (ft) PHYSICAL
SPAN
LOWER MIN. PHYS.
56.24 56.24 ≈0
15 18
300 1.04 59.70 56.7 56.2 0.167
0.5
14 55.2 54.7 18 0.15

56.24 55.7 0.54 0.18


14 18
300 4.9 59.80 56.2 55.7 0.167
0.5
13 54.7 54.2 18 0.15
54.98 50.9 0.38
13 18
300 7.05 59.00 55.7 51.5 1.4
4.2
12 54.2 50.0 18 0.15
50.9 47.9 0.58
12 18
300 8.65 54.50 51.5 47.5 1.33
4.0
11 50.0 46.0 18 0.15
47.9 47.1 0.78 0.26
11 24
300 12.5 50.50 47.5 46.5 0.33
1.0
outlet 45.5 44.5 24 0.1

The HG slopes shown for pipe sections P15-14, P13-12, P12-11 are the full-flow friction slopes.
The values are not the true HG slopes because these pipes are flowing part full. The
values will be revised in subsequent iterations through the system. The full-flow friction
slopes have been shown in Table 6.6-16 to help follow the discussion of Step 16 and Step
17 for the entire system.

18. Compare the hydraulic gradient elevation to the standard.


Throughout the system, the hydraulic gradient elevation is more than 1.13 feet
below the inlet elevation (edge of pavement), so it meets the current standard. We
will recalculate the flow rates and check again.

19. Recalculate the flow rates.


Several pipes are flowing partly full, so we need to recalculate the velocities and
times of flow in those sections. This will change the times of concentration and the
flow rates. Pipe sections P15-14, P13-12, and P12-11 are flowing part full and the others
are flowing full based on the calculations up to this point. We will assume these
modes of flow as we work downstream recalculating flow. The velocity in the three

Chapter 6: Storm Drains 6-92


January 2019

Drainage Design
Guide

Chapter 6: Storm Drains

pipes flowing partly full will be based on normal depth velocity (see Figure 6.5-3).
Table 6.6-17 shows the results of recalculating the flow rates.

Table 6.6-17: Results of Second Pass down the System


NOTES AND
REMARKS

VELOCITY
ACTUAL
STRUCTURE

PIPE

(FPS)
ZONE: 7
TIME OF CONCEN-

IN SECTION (min)

SIZE SLOPE (%)


TIME OF FLOW
TRATION (min)

TOTAL FLOW
FREQUENCY (Yrs): 3
NO.

LENGTH (ft)

TOTAL C•A
INTENSITY
(IN)
MANNINGS “n”: 0.012
(iph)

(cfs)
HYD. TAILWATER EL. (ft):

PHYSICAL
VELOCITY
GRAD. 47.1
RISE

(FPS)
All overland tc < 10
UPPER PHYSICAL
min.
LOWER SPAN MIN PHYS
≈0 Act Vel based on
15 18 2.1
300 10.0 2.4 6.5 0.16 1.04 0.167 normal
14 18 .15 2.6 Depth. d/D = 0.32
0.25
14 18 3.25
300 12.4 1.5 6.0 0.96 5.76 0.167
13 18 0.15 2.6
0.5 Act Vel based on
13 18 8.0
300 13.9 0.6 5.7 1.44 8.2 1.4 normal
12 18 0.15 7.6 Depth. d/D = 0.55
0.8 Act Vel based on
12 18 8.3
300 14.5 0.6 5.6 1.84 10.3 1.33 normal
11 18 0.15 7.5 Depth. d/D = 0.67
0.38
11 24 4.8
300 15.1 - 5.6 2.72 15.2 0.33
Outlet 24 0.10 4.5

20. Recalculate the hydraulic gradient elevation.


Work up the system, as was done previously in Step 16 and Step 17. Table 6.6-
18 shows the results.

Chapter 6: Storm Drains 6-93


January 2019

Drainage Design
Guide

Chapter 6: Storm Drains

Table 6.6-18: Results of Second Pass up the System


HYD. GRADIENT
PIPE NOTES AND REMARKS
STRUCTURE

SIZE
TOTAL FLOW
SLOPE (%)
LENGTH ( ft )

CROWN (IN) ZONE: 7


NO.

INLET FLOW LINE FREQUENCY (Yrs): 3


(cfs)

ELEV. MANNING’S “n”: 0.012


(ft) UPPER LOWER RISE
HYD GRAD TAILWATER EL. (ft): 47.1

FALL
END END

(ft)
ELEV. ELEV.
UPPER PHYSICAL
(ft) (ft) SPAN
LOWER MON PHYS All overland tc < 10 min.
56.45 56.45 ≈0
15 18 Act.Vel based on Norm
300 1.04 59.70 56.7 56.2 0.167
0.5 Depth. d/D = 0.32
14 55.2 54.7 18 0.15
56.45 55.7 0.75 0.25
14 18
300 5.76 59.80 56.2 55.7 0.167
0.5
13 54.7 54.2 18 0.15
55.03 51.0 0.5 Act Vel & Upper End HG
13 18
300 8.02 59.00 55.7 51.5 1.4 Based on Norm Depth
4.2
12 54.2 50.0 18 0.15 D/D = 0.55
51.0 48.24 0.8 Act Vel & Upper End HG
12 18
300 10.3 54.50 51.5 47.5 1.33 Based on Norm Depth
4.0
11 50.0 46.0 18 0.15 D/D = 0.67
48.24 47.1 1.14 0.38
11 24
300 15.2 50.50 47.5 46.5 0.33
1.0
OUT 45.5 44.5 24 0.10

The HG slopes shown for pipe sections P15-14, P13-12, P12-11 are the full-flow friction slopes.
The values are not the true HG slopes because these pipes are flowing part full. The full-
flow friction slopes have been shown in Table 6.6-18 to help you compare HG elevations
as you work through the system. The values are changed in Table 6.6-19, which reflects
the completed design.

21. Compare the hydraulic gradient to the standard.

Throughout the system, the hydraulic gradient elevation is more than 1.13 feet
below the inlet elevation (edge of pavement), so it meets the current standard. Pipe
section P11-out cannot be reduced in diameter without violating the standard HG
clearance at S-11 (see Step 15). The other pipes are the minimum standard
diameter, so their diameter cannot be reduced.

Pipe section P15-14 is flowing full for about half of its length. Consequently, the flow
velocity is less than the 2.1 fps we estimated in Table 6.6-17. We could make
another iteration through the system recalculating flows based on the reduced
velocity, but there is nothing to be gained from doing that here. None of the pipe
diameters can be reduced. A completed tabulation form is shown in Table 6.6-19.

Chapter 6: Storm Drains 6-94


January 2019

Drainage Design Guide


Chapter 6: Storm Drains

Example 6.6-2
Table 6.6-19
LOCATION DRAINAGE HYD. GRADIENT

VELOCITY (FPS)
OF AREA NOTES AND REMARKS
UPPER END (ac) CROWN PIPE

ACTUAL
STRUCTURE SIZE
ALIGNMENT

TYPE OF STRUCTURE
C= FLOW LINE (IN) ZONE: 7
NAME

TIME OF CONCEN-

IN SECTION ( min )
SLOPE (%)
NO.

SUBTOTAL C•A

TRATION (min )
TIME OF FLOW

TOTAL FLOW
LENGTH (ft)
C = 0.80

TOTAL C•A
INTENSITY
INLET FREQUENCY (Yrs): 3

(iph)

(cfs)
C= ELEV
UPPE
DISTANCE

(ft) LOWER
R RISE

FALL
SIDE

END
(ft)

(ft)
STATION END

PHYSCIAL
VELOCITY
INCRE ELEV HYD GRAD MANNING’S “n”: 0.012
ELEV

(FPS)
TOTA (ft)
UPPER - (ft)
L TAILWATER EL. (ft):
MENT PHYSICAL
SPA 47.1
LOWER N MIN PHYS All overland tc < 10 min

56.45 56.45 ≈0 ≈0
Patrick Place 15 18 2.1 Act Vel based on Norm
P1 300 0.2 0.2 0.16 10 2.4 6.5 0.16 1.04 59.70 56.7 56.2 0.167 Depth.
0.5 d/D = 0.32
14 55.2 54.7 18 0.15 2.6
56.45 55.7 0.75 0.25
Patrick Place 14 18 3.25
P1 300 1.0 1.2 0.96 12.4 1.5 6.0 0.96 5.76 59.80 56.2 55.7 0.167
0.5
13 54.7 54.2 18 0.15 2.6
55.03 51.0 4.03 1.3
Patrick Place 13 18 8.0 Act Vel & Upper End HG
P1 300 0.6 1.8 1.44 13.9 0.6 5.7 1.44 8.2 59.00 55.7 51.5 1.4 based on Norm Depth.
4.2 d/D = 0.55
12 54.2 50.0 18 0.15 7.6
51.0 48.24 2.79 0.9
Patrick Place 12 18 8.3 Act Vel & Upper End HG
P1 300 0.5 2.3 1.84 14.5 0.6 5.6 1.84 10.3 54.50 51.5 47.5 1.33 based on Norm Depth.
4.0 d/D = 0.67
11 50.0 46.0 18 0.15 7.5
48.24 47.1 1.14 0.38
Patrick Place 11 24 4.8
P1 300 1.1 3.4 2.72 15.1 1.0 5.6 2.72 15.2 50.50 47.5 46.5 0.33
1.0
OUT 45.5 44.5 24 0.10 4.5

Chapter 6: Storm Drains 6-95


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

CHAPTER 7: EXFILTRATION SYSTEMS

7. EXFILTRATION SYSTEMS .................................................................................. 7-1

7.1 General ................................................................................................................. 7-1


7.1.1 Hydrology ....................................................................................................... 7-1
7.1.1.1 Time of Concentration............................................................................. 7-1
7.1.2 Hydro-Geology ............................................................................................... 7-1
7.1.2.1 Darcy’s Law ............................................................................................ 7-1
7.1.2.2 Soil Permeability ..................................................................................... 7-2
7.1.2.3 Hydraulic Conductivity of Soils ................................................................ 7-2
7.1.2.4 Hydro-Geologic Tests ............................................................................. 7-3
7.1.2.5 Seasonal High Water Table .................................................................... 7-3
7.1.2.6 Average Antecedent Moisture Conditions ............................................... 7-4
7.1.3 Data Collection............................................................................................... 7-4
7.1.4 Permitting Considerations .............................................................................. 7-5
7.1.5 Construction and Maintenance Considerations.............................................. 7-5

7.2 Exfiltration Trenches .......................................................................................... 7-6


7.2.1 Description ..................................................................................................... 7-6
7.2.1.1 Use ......................................................................................................... 7-7
7.2.2 Design Criteria ............................................................................................... 7-8
7.2.2.1 Water Quality .......................................................................................... 7-8
7.2.2.2 Water Quantity ...................................................................................... 7-10
7.2.2.3 Design Ground Water Elevation ........................................................... 7-10
7.2.2.4 Control Elevation .................................................................................. 7-10
7.2.2.5 Effective Head ...................................................................................... 7-11
7.2.2.6 Recovery Time ...................................................................................... 7-11
7.2.2.7 Safety Factor ........................................................................................ 7-11
7.2.2.8 Dimensions ........................................................................................... 7-11
7.2.2.9 Maximum Length .................................................................................. 7-11
7.2.2.10 Pipe Invert............................................................................................. 7-12
7.2.2.11 Aggregates ........................................................................................... 7-12
7.2.2.12 Filter Fabric ........................................................................................... 7-12
7.2.2.13 Drainage Structures .............................................................................. 7-12
7.2.3 Boundary Conditions .................................................................................... 7-13
7.2.3.1 Ground Water Elevation........................................................................ 7-13
7.2.3.2 Tailwater Elevation ............................................................................... 7-13
7.2.3.3 Headwater Elevation ............................................................................. 7-13
7.2.4 Methodologies to Design Exfiltration Trenches ............................................ 7-13
7.2.4.1 Storage-Recovery Method .................................................................... 7-14
7.2.4.2 Empirical Equations Method ................................................................. 7-17

Chapter 7: Exfiltration Systems 7-i


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

7.2.4.3 FDOT District VI Method ....................................................................... 7-20


7.2.4.4 Other Design Methods .......................................................................... 7-24

7.3 Drainage Wells ............................................................................................... 7-24


7.3.1 Description ................................................................................................... 7-24
7.3.1.1 Use ....................................................................................................... 7-25
7.3.2 Design Criteria ............................................................................................. 7-26
7.3.3 Water Quality ............................................................................................... 7-26
7.3.3.1 Fresh Water-Salt Water Hydrostatic Balance ....................................... 7-27
7.3.3.2 Hydraulic Head ..................................................................................... 7-28
7.3.3.3 Safety Factor ........................................................................................ 7-29
7.3.3.4 Dimensions ........................................................................................... 7-29
7.3.3.5 Exfiltration Rate .................................................................................... 7-29
7.3.3.6 Casing................................................................................................... 7-29
7.3.4 Methodologies of Calculation ....................................................................... 7-29
7.3.4.1 Gravity Wells ......................................................................................... 7-30
7.3.4.2 Pressurized Wells ................................................................................. 7-32

7.4 Modeling Exfiltration Systems ......................................................................... 7-33


7.4.1 Basic Modeling Concepts ............................................................................ 7-33
7.4.2 Exfiltration Trenches Modeling ..................................................................... 7-34
7.4.3 Drainage Well Modeling ............................................................................... 7-35

Chapter 7: Exfiltration Systems 7-ii


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

7. EXFILTRATION SYSTEMS

7.1 GENERAL
7.1.1 Hydrology
Chapter 2 in this design guide and the Drainage Manual encompass the Department’s
general guidance regarding hydrology. Coordinate in advance with the District Drainage
Engineer for approval of the design criteria and calculation methods.

7.1.1.1 Time of Concentration


Chapter 2 (Hydrology) defines and provides methods to calculate the time of
concentration. A longer time of concentration usually reduces the calculated peak
discharge. The Rational Method is very sensitive to changes in the time of concentration
(i.e., if the time of concentration increases from 10 minutes to 60 minutes, the calculated
peak discharge could be reduced by up to 60 percent). The Flow Hydrograph Methods
are less sensitive to the time of concentration changes (i.e., up to 15 percent reduction in
peak discharge if the time of concentration increases from 10 minutes to 60 minutes).

7.1.2 Hydro-Geology
7.1.2.1 Darcy’s Law
Darcy’s Law characterizes the flow through porous media, assuming that the viscosity,
temperature, and density of the fluids are constants. The flow rate is a function of the flow
area, the hydraulic gradient, and the proportionality constant (refer to Figure 7.1-1):

Q=kiA

where:
Q = Flow rate, in ft3/sec
k= Permeability constant, in ft/sec
i= Hydraulic gradient (i = ∆H/L)
A = Cross-sectional area of soil conveying flow, in ft2
∆H = Change in the hydraulic grade line, in ft
L= Distance between points of interest, in ft

Chapter 7: Exfiltration Systems 7-1


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Figure 7.1-1: Saturated Flow through Porous Media

Darcy’s Law was established for saturated flow. As such, it may be adjusted for
unsaturated and multiphase flows.

7.1.2.2 Soil Permeability


The coefficient of permeability (k) in Darcy’s Law is a measure of the rate of water flow
through a saturated soil under a given hydraulic gradient in length unit over time unit (i.e.,
ft/day). The soil permeability is dependent on the grain-size distribution and void ratio.
The coefficient of permeability (k) typically varies from 0.03 ft/sec (43 ft/day) for gravels
to less than 10-8 ft/sec (1.44 x 10-5 ft/day) for clays (refer to Appendix J).

7.1.2.3 Hydraulic Conductivity of Soils


The hydraulic conductivity of a soil (K) measures the relative ease of water transmission
through the soil:

K= Q
A ∆H

where:
Q = Flow rate, in ft3/sec
A = Flow area, in ft2
∆H = Hydraulic head, in ft

The hydraulic conductivity of a soil is the ratio between the discharge through the unit
area of soil perpendicular to the flow per unit of head (i.e., cfs/ft2 – ft of head). This is the
primary factor used to determine the exfiltration rate of a system.

Chapter 7: Exfiltration Systems 7-2


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

The flow transmission through a soil increases as the water content increases. As such,
the maximum hydraulic conductivity occurs under saturated conditions. The hydraulic
conductivity for horizontal-saturated flow (Ks) usually is several times greater than the
hydraulic conductivity for vertical-unsaturated flow (Ku).

7.1.2.4 Hydro-Geologic Tests


The hydro-geotechnical properties of a soil can be measured in tests under falling or
constant head, either in the laboratory or in the field. The most effective soil testing is a
combination of laboratory and field methods. Laboratory tests on undisturbed soil
samples usually provide accurate results representative of only a point among the soil
stratum.

Evaluate the hydrologic and geologic characteristics of the site where the exfiltration
system will be installed to define the test procedures to be used. The following tests are
suggested:

• Laboratory Permeameter Test for saturated hydraulic conductivity on


undisturbed soil samples (ASTM D 5084)
• Double Ring Infiltrometer Test to estimate the initial vertical unsaturated
permeability data of the upper soil layer (ASTM D 3385)
• Constant Head Test in soils with permeabilities that allow keeping the test hole
filled with water during the field test (AASHTO T 215)
• Falling Head Test in areas with excellent soil percolation where keeping the test
hole filled with water is not feasible during the test (FM 5-513)
• DOT Standard Test (constant head) that can be used for the Department’s
projects (FM 1-T 215)
• Well Test Holes are performed to determine relative permeability and water
quality characteristics of the aquifer (ASTM D4050); through continuous water
quality testing, the test hole will indicate the depth at which a minimum of 10,000
milligrams per liter total dissolved solids (TDS) concentration is found; the test also
will indicate the most favorable depth for stormwater discharge
• A Pumping Test is performed at the most favorable depth for stormwater
discharge to determine the design discharge capacity normally in gallons per
minute per foot of head; the test is normally performed in conjunction with well test
holes

7.1.2.5 Seasonal High Water Table


Published information (such as data from the Natural Resources Conservation Service,
formerly known as the Soil Conservation Service) provides preliminary guidance related
to the water table at a specific location, but you will need site-specific water table
information to design exfiltration systems.

Chapter 7: Exfiltration Systems 7-3


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

The initial data to determine the seasonal high water table (SHWT) elevation is the
measurement of the stabilized ground water level in a boring or well. Adjust the initial
(encountered) water table elevation to estimate the SHWT based on antecedent rainfall,
examination of the soil profile (color variations, depth to hardpan, etc.), consistency with
water levels of the adjacent water bodies, vegetative indicators, etc.

7.1.2.6 Average Antecedent Moisture Conditions


The Antecedent Moisture Condition (AMC) indicates the wetness of a soil and its
availability to infiltrate water. The soil moisture ranges from dry to saturated, depending
on the rainfall amount prior to the moisture measurement. The average AMC means that
the soil is neither dry nor saturated, but at an average moisture condition at the beginning
of the design storm event.

7.1.3 Data Collection


The design of an exfiltration system requires a good understanding of the site conditions.
Information required and potential sources include:

• Topographic Data. You usually can find preliminary topographic data in the
United States Geological Survey (USGS) quadrangle maps, topographic LIDAR
data, and previous project construction plans. Supplement this information with a
detailed topographic survey of the project area.

• Geotechnical Data. The Soil Survey Reports by the United States Department of
Agriculture (USDA) Natural Resources Conservation Service (NRCS) provide
general geological and geotechnical properties information. Previous geotechnical
reports for the project area and adjacent developments can provide more specific
information regarding the geotechnical conditions at the project location. After
preliminary evaluation of the available data, request a detailed geotechnical study.
The site-specific geotechnical report should classify the types of soils within the
project location and soil engineering characteristics, including hydraulic
conductivity (refer to Section 7.1.2.3), ground water elevations, etc.

• Receiving Water Bodies. The Water Management Districts, FEMA, and some
local agencies can provide information regarding water elevations under different
storm frequencies for lakes, rivers, canals, and reservoirs. Some agencies also
can provide potentiometric surface maps to assist in determining the ground water
elevations. Tidal information is available on the National Oceanic and Atmospheric
Administration (NOAA) website. You can determine the design tailwater elevation
from the above sources.

Chapter 7: Exfiltration Systems 7-4


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

• Permit Data. Previous permit information for the site and surrounding
developments can provide information related to design criteria, existing wetlands,
possible outfalls, discharge limitations, control elevations, off-site contributors,
geotechnical data, prior soil testing results, etc.

• Right-of-Way Data. You can obtain the right-of-way information from the
Department’s right-of-way maps and county and city right-of-way documents.

• Field Reviews. Visiting and inspecting the site provides first-hand updated
information to the designer regarding the existing drainage system, off-site
contributors, water management facilities, outfalls, and other site conditions.

7.1.4 Permitting Considerations


Generally, you will develop the drainage design in compliance with all applicable federal,
state, and local environmental regulatory programs. The respective permitting authorities
have to be identified and contacted early in the design process. These agencies include
local water control districts and county and state water management districts. Each permit
agency has specific water quality requirements and may impose restrictions on the
construction of exfiltration trenches and well systems.

7.1.5 Construction and Maintenance Considerations


Install stormwater exfiltration systems no less than two feet from parallel underground
utilities and 20 feet from existing large trees that will remain in place. To avoid damaging
adjacent properties, carefully evaluate the existing soils and the excavation method if the
exfiltration system is located in close proximity to the right-of-way line. Implement erosion
control measures to impede the access of sediments and debris into the exfiltration
system during construction, which can clog the filter fabric and diminish the capacity of
the exfiltration trench.

Typically, you would not use exfiltration systems within any type of manmade, compacted
embankment since there is little to no percolation in compacted fill as compared to natural
soils.

Do not install exfiltration systems in close proximity to or behind MSE walls. Install solid
conveyance pipes behind the MSE wall in accordance with the Drainage Manual,
Appendix D, and install exfiltration systems away from the walls. Do not use exfiltration
trenches in locations where a 1H:1V mound could allow the filtrate to impact the MSE soil
reinforcements. In this situation, the potential exists for accelerated corrosion of metallic
reinforcements to occur without warning. Furthermore, seepage forces would need to be
included in the design of the wall, and daylighting filtrate could result in soil washouts,
unsightly mildew, vegetation, staining, and other maintenance problems.

Chapter 7: Exfiltration Systems 7-5


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Physical access devices must be provided with stormwater exfiltration systems to


facilitate maintenance activities. Consider minimum pipe sizes and maximum spacing
between drainage structures (refer to Sections 7.2.2.8 to 7.2.2.13) for the efficient
operation of maintenance equipment. Also consider future expansion of the facilities and
the possible increase of maintenance requirements.

In the case of drainage wells, provide the injection well chamber with physical access
devices for maintenance activities. Maintenance of the injection well includes cleaning,
removing debris, and, in some cases, redeveloping the well to re-establish discharge
capacity. The well location needs to be accessible from the surface to allow these
activities to take place.

7.2 EXFILTRATION TRENCHES


7.2.1 Description
An exfiltration trench is an underground drainage system consisting of a perforated pipe
surrounded by natural or artificial aggregate, which stores and infiltrates runoff (refer to
Figure 7.2-1). Catch basins located at the end of each exfiltration trench segment collect
stormwater runoff; the perforated pipe delivers the stormwater into the surrounding
aggregate through the pipe perforations. The stormwater ultimately exfiltrates into the
ground water aquifer through the trench walls and bottom. As the treatment volume is not
discharged into surface waters, exfiltration trench systems are considered a type of
retention treatment.

Chapter 7: Exfiltration Systems 7-6


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Figure 7.2-1: Typical Exfiltration Trench

The permeability of the soils at the exfiltration trench location and the anticipated water
table elevation determine the applicability and performance of the exfiltration trench
system. This system must exfiltrate the required stormwater treatment volume and draw
down the treatment volume to return to its normal condition within a specific time after the
design storm event. When the trench bottom is at or above the average wet season water
table, the exfiltration trench is considered a dry system.

7.2.1.1 Use
For projects where the areas available for water management facilities are limited and
high right-of-way acquisition costs are anticipated, exfiltration trench systems can provide
the required stormwater treatment if the hydro-geological conditions are suitable for runoff
infiltration (i.e., permeable soils with hydraulic conductivity exceeding 1 x 10-5 cfs/ft2 per
foot of head). Exfiltration trenches, like other types of retention systems, are able to
efficiently remove stormwater pollutants. Additionally, exfiltration trenches contribute to
recharge of the ground water aquifer, thus assisting in combatting saltwater intrusion in
coastal areas.

Chapter 7: Exfiltration Systems 7-7


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Because of the direct infiltration of the surface runoff with its associated pollutant load into
the ground water aquifer, do not install exfiltration trench systems in the proximity of
potable water supply wells. Usually, you are not allowed to install exfiltration trenches
within the 10-day well field protection contour, but you should verify specific requirements
for each well field with the permitting authorities. Do not propose exfiltration trenches
within or near contaminated ground water areas to avoid the potential migration of the
polluted plume due to the direct injection of surface runoff into or adjacent to the
contaminated groundwater plume. In areas with high ground water elevation, the
available hydraulic head for exfiltration trench operation is minimal but the required
hydraulic head can be obtained by pumping, if feasible.

The limited life span of exfiltration trenches is their main disadvantage. The accumulation
of sediments and clogging of the filter fabric and the void spaces of the aggregates usually
shorten the operational life of exfiltration trenches. Consider the need of future
replacement costs in the evaluation of their effectiveness. Prior to replacing existing
systems or using them as part of a new drainage system, test the remaining treatment
capacity of existing exfiltration trenches.

7.2.2 Design Criteria


An exfiltration trench transmits the inflow runoff hydrograph into the groundwater during
small storm events or in land-locked conditions; but in drainage areas with positive outfall,
the fraction of the runoff hydrograph that is not transmitted into the groundwater and
retained within the exfiltration trench is transmitted downstream, usually through an outfall
control structure.

The Standard Specifications for Road and Bridge Construction (Section 443, French
Drains) includes directions, provisions, and requirements for exfiltration trenches.
Standard exfiltration trenches are detailed in Standard Plans, Index 443-001, French
Drains. In the cases where Standard Plans, Index 443-001 is not suitable for a specific
project need, develop a detailed design and include this information in the design
documentation. The following are the Department’s general design criteria. It is
recommended that additional specific criteria from the permitting agencies be evaluated
in the design process.

7.2.2.1 Water Quality


You can install the exfiltration trenches off-line or on-line in the drainage system to provide
water quality treatment to a watershed.

Chapter 7: Exfiltration Systems 7-8


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Figure 7.2-2: Off-Line Exfiltration System

The off-line treatment method (Figure 7.2-2) diverts runoff into the exfiltration trench
designed to provide the required treatment volume; subsequent runoff in excess of the
treatment capacity bypasses the off-line exfiltration trench toward the outfall. For off-line
systems, a diversion drainage structure usually is required.

Figure 7.2-3: On-Line Exfiltration System

Chapter 7: Exfiltration Systems 7-9


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

The on-line exfiltration trench (Figure 7.2-3) provides the required water treatment but the
treatment volume is mixed with the total runoff volume. As such, runoff volume in excess
of the treatment capacity carries a portion of the pollutant load to the receiving water body.

7.2.2.2 Water Quantity


For exfiltration trenches designed to satisfy water quality requirements rather than provide
flood protection, only the fraction of the overall exfiltration trench storage volume,
including pipe and aggregate voids, located above the design ground water elevation and
below the outfall control elevation should be considered for discharge attenuation.

In some special locations (i.e., Miami-Dade County and Monroe County) with very limited
area available for water treatment, the exfiltration trench systems could be credited for
discharge attenuation if the ground water is considered variable, rising from the Seasonal
High Water Table along with the design storm event.

7.2.2.3 Design Ground Water Elevation


Use the elevation to which the ground water can be expected to rise during a normal wet
season to calculate the required exfiltration trench length.

7.2.2.4 Control Elevation


The minimum control elevation for an exfiltration trench system should be at the same
elevation as the top of the perforated pipe. The maximum control elevation should not
violate the base clearance criteria for the project or produce changes in the land use value
of the properties located upstream and downstream of the drainage system. A site-
specific survey, the permit files, and the field reviews are the main sources used to
determine the design control elevation.

Figure 7.2-4: Outfall Control Structure

Chapter 7: Exfiltration Systems 7-10


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

7.2.2.5 Effective Head


Positive Discharge. The effective head of the exfiltration trenches (Heff) with discharge
to an outfall should be the average vertical distance from the SHWT to the outfall control
elevation:

Heff = Control Elevation – SHWT


2

Closed Systems. The effective head for exfiltration trenches with no outfall (self-
contained system) should be the vertical distance from the SHWT to the average distance
between the SHWT and the design high water (DHW):

Heff = DHW + SHWT – SHWT = DHW – SHWT


2 2

7.2.2.6 Recovery Time


If the permitting authorities with jurisdiction over the project location do not have specific
recovery time requirements, the water treatment capacity of the exfiltration trench
systems should be recovered within the 72 hours following the design storm event,
assuming average AMC.

7.2.2.7 Safety Factor


Use a safety factor of two or more to calculate the required length of exfiltration trenches
to consider possible geotechnical uncertainties.

7.2.2.8 Dimensions
The minimum pipe diameter is 18 inches, but 24 inches is preferable; and the minimum
trench width is four feet. The maximum dimensions should depend on site-specific
conditions and construction methods. In general, exfiltration trenches with bottoms wider
than eight feet and/or deeper than 20 feet are not recommended. Perforated pipes with a
diameter of more than 36 inches should be approved by the District Drainage Engineer.

Pipe perforations can be slotted or perforated. Standard locations and dimensions of the
pipe perforations are included in the Standard Specifications, Section 443, French Drains.

7.2.2.9 Maximum Length


a. The maximum length of exfiltration trenches with access through both ends should
be:
For 18-inch to 30-inch pipes 300 feet
Chapter 7: Exfiltration Systems 7-11
January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

For 36-inch and larger pipes 400 feet

b. The maximum length of the exfiltration trenches with access through only one end
should be half of the maximum length of exfiltration trenches with access through
both ends.

7.2.2.10 Pipe Invert


Make the invert elevation of the perforated pipe at least one foot above the trench bottom
elevation. Locate the pipe invert above the SHWT to facilitate maintenance operations.
This criterion may not be feasible in sites where the water table is close to the ground
surface or where a deeply permeable stratum underlies low-permeability soils.

7.2.2.11 Aggregates
Use uniform-graded, natural or artificial coarse aggregate with no more than 3-percent
weight of material passing the Number 200 Sieve at the point of use.

7.2.2.12 Filter Fabric


Enclose the coarse aggregate of an exfiltration trench in filter fabric. The perforated pipe
also could be enclosed in filter fabric to increase the life span of the exfiltration trench if
approved by the District Drainage Engineer.

The filter fabric will comply with the requirements established in the latest FDOT Standard
Specifications, Section 985. Additionally, the permeability of the filter fabric must be equal
to or greater than the permeability of the surrounding soil.

7.2.2.13 Drainage Structures


The minimum side dimension of the drainage structures for exfiltration trenches should
be four feet. Inlets must include sediment sumps to collect sediments and skimmers/
baffles (refer to Standard Plans, Index 241) to prevent oil and floating debris from exiting
the catch basin into the exfiltration trench. The minimum clear distance between baffles
in the same drainage structure will be 2.5 feet. Fiberglass skimmers and baffles are not
recommended due to possible damage from debris impact.

Drainage structures have to provide adequate access to the exfiltration trench for
maintenance operations; the minimum grate size should be two feet and two-piece cast
iron covers (Standard Plans, Index 425-001) are recommended. Provide manholes for
inspection and clean out at the end of each exfiltration trench with no inlet. Inlets Type 1
to 4 (Standard Plans, Index 425-020) are recommended for exfiltration trenches installed
along or from a gutter line.

Chapter 7: Exfiltration Systems 7-12


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Refer to Sections 3.10 and 3.12 of the Drainage Manual for standards related to drainage
structures of exfiltration systems.

7.2.3 Boundary Conditions


The design and performance of an exfiltration trench system depends on the specific
boundary conditions of the site, which are: the ground water elevation, the tailwater
elevation (if positive outfall), and the allowable headwater.

7.2.3.1 Ground Water Elevation


The ground water elevation to design exfiltration trench systems will be the seasonal high
ground water table as defined in Section 7.1.2.5, above.

7.2.3.2 Tailwater Elevation


The receiving water body defines the tailwater conditions for exfiltration trench systems
with positive discharge. Define the design tailwater as per the latest Drainage Manual,
Section 3.4.

7.2.3.3 Headwater Elevation


The maximum allowable stage upstream of the exfiltration trench system will limit the
design high water elevation. The drainage design in general should limit the design high
water during the design storm event to meet the base clearance requirements and cause
no adverse impact to the land use value of the surrounding properties.

7.2.4 Methodologies to Design Exfiltration Trenches


There are several methodologies used to design exfiltration trench systems. All methods
are similar in nature, with specific criteria and requirements set by the regulatory agencies
and FDOT District Drainage Offices. As such, it is important to coordinate and get
approval of the methodology used in each specific project from the District Drainage
Engineer and the permitting authorities with jurisdiction over the area where the proposed
drainage system will be installed.

The equations and formulas included in the following sections present the conceptual
development of the procedures and calculations, which are applicable with any unit
system. As such, the conversion factors are not included, but the designer has to convert
the units of each parameter as required to be consistent.

To illustrate the calculation methods, use/calculate a sample roadway segment (Figure


7.2-5) with a contributing drainage area of 2.3 acres (including 0.8 acres of pavement)
with on-line treatment.

Chapter 7: Exfiltration Systems 7-13


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Figure 7.2-5: Sample Problem

7.2.4.1 Storage-Recovery Method


This method is acceptable for most of the permitting agencies and FDOT Districts
because it provides the required water quality storage capacity within the exfiltration
trench and assures that the treatment capacity will be available again within the required
recovery time.

The storage-recovery method is recommended for exfiltration trenches with bottom


elevation—or at least the perforated pipe invert—above the design ground water
elevation. This method is not applicable if the top of the proposed exfiltration trench is
below the design ground water elevation.

Storage-Recovery Design Procedure

Step 1: Data Collection

Chapter 7: Exfiltration Systems 7-14


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Total drainage area: AT = 2.3 acres

Pavement area: AI = 0.8 acre

Ground water elevation: SHWT = 6.00 ft

Tailwater elevation: TW = 10.00 ft

Headwater elevation: HW = 14.00 ft

Fillable porosity of soil: fs = 0.3

Fillable porosity of aggregate: f = 0.45

Unsaturated hydraulic Ku = 0.00007 ft3/sec


2
conductivity of soil: ft -ft of head

Saturated hydraulic Ks = 0.00025 ft3/sec


2
conductivity of soil: ft -ft of head

Step 2: Calculate the required storage for water quality

Note: The water quality criterion from St. Johns River Water Management District
(SJRWMD) is used for the sample problem.

If off-line treatment is proposed, use the greatest of the following volumes:

VT = 0.5 inch/12 AT VT = 0.096 acre-ft

VI = 1.25 inches/12 AI VI = 0.083 acre-ft Voff = 0.096 acre-ft

If you propose on-line treatment, increase 0.5-inch runoff on the total drainage area to the
off-line treatment volume above:

Von = Voff + VT = 0.096 + 0.096; Von = 0.192 acre-ft

If discharging into shellfish harvesting areas or other receiving water bodies with specific
regulations, provide the water quality volume as required: Vspec = 0 acre-ft

Total required water quality volume: VWQ = Von; VWQ =0.192 acre-ft

Step 3: Define the preliminary characteristics of the exfiltration trench

Outfall control elevation: CE = 13.00 ft Perforated pipe diameter: D = 24 inches

Chapter 7: Exfiltration Systems 7-15


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Trench bottom width: Wtr = 5.00 ft Perforated pipe invert: Pinv= 10.00 ft
Trench bottom elevation: Bel = 8.00 ft

Trench top width: Ttr = 5.00 ft

Trench top elevation: Tel = 13.00 ft

Figure 7.2-6: Sample Exfiltration Trench

Step 4: Calculate the required length of exfiltration trench

The storage capacity of an exfiltration trench is the available void space above the SHWT,
including pipe and aggregate voids.

Storage within pipe:

If the pipe invert is higher than the SHWT, the storage area in the pipe is:
Afull = π D2 Afull = 3.142 ft2
4

Storage within aggregate:

Storage height in trench: Du = Tel − SHWT Du = 7.00 ft


(unsaturated depth)

Storage area in trench: Atrench = f (Wtr∙Du−Apipe) Atrench=14.34 ft2


Note: Engineering judgment will be used regarding whether to consider the thickness of
the pipe walls to calculate Atrench.

Net trench length: Lnet = VWQ


Chapter 7: Exfiltration Systems 7-16
January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Apipe + Atrench Lnet = 477.7 ft

Safety factor to account for hydro-geotechnical uncertainties (SF ≥ 2) SF = 2

Required trench length: Lreq = SF Lnet Lreq = 955 ft

Note: If the required length does not fit within the available location, return to Step 2 and
modify the preliminary characteristics of the exfiltration trench as necessary.

Step 5: Calculate the recovery time of the treatment volume

Effective head: Heff = DHW − SHWT Heff = 4.00 ft


2

Trench height: Htr = Tel − Bel Htr = 5.00 ft

Unsaturated exfiltration will occur only through the walls (2) and the bottom if the trench
bottom is above the SHWT:
Auwb = Lnet (2Du + Wtr) Auwb = 9,076 ft2

Determine the unsaturated exfiltration capacity of the exfiltration trench:

Qu = Ku Auwb Heff Qu = 2.54 ft3/sec

Recovery time (Trec < 72hr): Trec = VWQ/Q Trec= 0.91hr OK

7.2.4.2 Empirical Equations Method


The following exfiltration trench design formulas have been developed by the South
Florida Water Management District (SFWMD) and are included in the SFWMD
Environmental Resource Permit Information Manual, Volume IV. This method calculates
the required length of exfiltration trench based on a one-hour exfiltration time, which is
representative of the majority of small-magnitude and short-duration rainfall events.

Empirical Equations Design Procedure

Step 1: Data Collection (refer to Figures 7.2-1 and 7.2-5)

Total drainage area: A = 2.3 acre

Chapter 7: Exfiltration Systems 7-17


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Pavement area: Ai = 0.8 acre

Ground water elevation: SHWT = 5.00 ft

Tailwater elevation: TW = 6.00 ft

Headwater elevation: HW = 8.00 ft

Saturated hydraulic conductivity: K = 0.00025 ft3/sec


ft2 - ft of head

W.Q. reduction (0.5 for retention): %WQ = 0.5

Safety factor (2 Minimum): FS = 2

Step 2: Calculate the required storage for water quality

The empirical equations have incorporated the adjustment to consider that exfiltration
trenches are retention systems (50 percent of the treatment volume required for wet
detention systems) and have a safety factor of 2. As such, the treatment volume to be
used in these formulas is the greatest of one-inch runoff on the contributing area or 2.5
inches on the pavement area:

VT = 1.0 inch/12 A VT = 0.19 acre∙ft

VI = 2.5 inches/12 Ai VI = 0.17 acre∙ft VWQ = 0.19 acre-ft

Step 3: Define the preliminary characteristics of the exfiltration trench

Outfall control elevation: CE = 6.50

Perforated pipe diameter: D = 24 in

Trench bottom width: W = 5.00 ft

Perforated pipe invert: Pinv = 10.00 ft

Trench bottom elevation: Bel = 1.00

Trench top elevation: Tel = 10.00

Chapter 7: Exfiltration Systems 7-18


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Figure 7.2-7: Sample Exfiltration Trench

Step 4: Calculate the required length of exfiltration trench

Treatment volume in acre-inches: V = 12VWQ V = 2.30 acre-in

Effective head: Heff = CE – SHWT Heff = 1.50 ft


2

Trench height: H = Tel – Bel H = 9.00 ft

Unsaturated depth of trench: Du = Tel – SHWT Du = 5.00 ft

Saturated depth of trench: Ds = SHWT – Bel Ds = 4.00 ft

When the unsaturated depth of trench is greater than the saturated depth or the trench
width is lesser than two times depth:

L1 = FS %WQ.VWQ L1= 33 ft
K (H2W + 2Heff Du – Du2 + 2Heff Ds) + 0.000139 W Du

When the saturated depth of trench is greater than the unsaturated depth or the trench
width is greater than two times depth:

L2 = FS %WQ.VWQ L2=579 ft
K (2Heff Du – Du2 + 2Heff Ds) + 0.000139∙W∙Du

Required length of trench (use the greater of these two lengths):


Lreq = L1 if Du ≥ Ds or L2 if Du < Ds Lreq= 33 ft
Lreq = L1 if W ≤ 2H or L2 if W > 2H Lreq= 33 ft

Chapter 7: Exfiltration Systems 7-19


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

For the sample problem, the saturated depth of trench is greater than the unsaturated
depth, but the trench width is less than two times the trench depth. As such, the required
length of trench is 33 feet.

7.2.4.3 FDOT District VI Method


The technical paper Subsurface Drainage with French Drains, prepared by the District VI
Drainage Section on June 20, 1991 (available on the Department’s ERC system as a
District VI document) includes criteria and procedures to design exfiltration trenches. This
method is acceptable for projects within the FDOT District VI jurisdiction, mainly within
Miami-Dade County, and other areas if approved by the permitting authorities.

This method considers that there is no flow through the bottom of the exfiltration trench
(assuming that the bottom is the first portion of the trench to get clogged), but only through
the vertical areas (walls) of the exfiltration trench. The hydraulic conductivity of the
existing soils at the depth where the exfiltration trench will be located is considered in
calculating the dimensions and required length of the trench. A test procedure has been
developed to determine the hydraulic conductivity (K) of the soils at different depths. The
initial investigation includes soils from 0 to 10-foot depths; if the test hole exfiltration rate
is less than 6 gpm, then soils from 10-foot to 15-foot depths are investigated. If the
accumulated exfiltration rate is still less than 6 gpm, soils from 15-foot to 20-foot depths
are investigated. Deeper exfiltration trenches are not considered economically practical.
Construct the exfiltration trench with its bottom elevation coinciding with the depth of the
selected test results.

FDOT District VI Design Procedure

Step 1: Data collection (refer to Figure 7.2-5)

Total drainage area: A = 2.3 acres Time of concentration: tc = 11 min


Pavement area: Ai = 0.8 acre Runoff coefficient
Pervious area: Ap = A - Ai Impervious: Ci = 0.9
Design frequency: F =10 years Pervious: Cp = 0.3
Ground water elevation: SHWT = 11.00 ft Tailwater elevation: TW=10.00

Chapter 7: Exfiltration Systems 7-20


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Figure 7.2-8: Sample Exfiltration Trench

Saturated hydraulic conductivity in cfs/square foot per foot of head:


Depth from 0 feet to 10 feet: d1 = 10 ft K10 = 0.000152
Depth from 10 feet to 15 feet: d2 = 5 ft K15 = 0.000211
Depth from 15 feet to 20 feet: d3 = 5 ft K20 = 0.000349

Step 2: Determine the maximum polluted volume

Note: The following procedure is applicable only in Miami-Dade County, and it is required
by the Miami-Dade Department of Environmental Resources Management (DERM).

Weighted runoff coefficient: C = Ci∙Ai + Cp∙Ap C = 0.51

Chapter 7: Exfiltration Systems 7-21


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Time to generate 1” of runoff: t1 = 2,940 F-0.11 t1 = 21.07 min


308.5 C – 60.5(0.5895 + F-0.67)

Polluted runoff duration: tT = t1 + tc = 21.07 + 11 tT = 32.07 min

Rainfall intensity: i= 308.5 i = 4.86 in/hr


– 0.11
(For Miami-Dade-County) 48.6 F + TT (0.5895 + F-0.67)

Peak discharge Q=CiA Q = 5.70 cfs

Maximum polluted volume: V = 60 Q TT V = 10,939 ft3

Note: All the runoff generated from a storm event lasting TT or less is assumed to be
polluted or contaminated. As such, the maximum polluted volume (V) usually is greater
than the treatment volume required by the Water Management Districts.

Step 3: Define the preliminary characteristics of the exfiltration trench

Outfall control elevation: CE = 13.00 ft

Trench bottom width: W = 5.00 ft

Trench bottom elevation: Bel = -7.00 ft

Trench top elevation; Tel = 13.00 ft

Perforated pipe diameter: D = 24 in


Perforated pipe invert: Pinv = 10.00

Fillable porosity of aggregate: f = 0.5

Effective head: Heff = CE – SHWT Heff=2.00 ft


2

Trench height: H = Tel – Bel H=20.00 ft

Unsaturated depth of trench: Du = Tel – SHWT Du=2.00 ft

Saturated depth of trench (0’ to 10’) Ds = SHWT – Be Ds=8.00 ft

Step 4: Determine the trench storage

Chapter 7: Exfiltration Systems 7-22


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

The storage capacity of an exfiltration trench is the available void space above the design
ground water elevation (SHWT), including pipe and aggregate voids.

Storage within pipe:

If the pipe invert is higher than the SHWT, the storage area in the pipe is:
Afull = π (D/12)2 Afull=3.14 ft2
4
If the pipe invert is lower than the SHWT, the available storage depth in the pipe is:
Dpipe = Pinv + D – SHWT Dpipe=1.00ft
12

Central angle of the circle segment: θ = 2acos[(D – 2Dpipe)] θ = 0.82 rad


D

Storage area in pipe segment: Aseg = (D/12)2 (θ – sinθ) Aseg=0.04 ft2


8
Apipe = Afull if Pinv l ≥ SHWT
Apipe = Aseg if Pinv < SHWT

In this example: Apipe = Afull Apipe=3.14 ft2

Storage within aggregate:

Storage height in trench:


(unsaturated depth) Du = Tel = SHWT Du = 2.00 ft

Storage area in trench: Atrench = f (W Du – Apipe) Atrench=3.43 ft2

Storage in trench: S = Apipe + Atrench S = 6.57 ft3/ft

Step 5: Determine the exfiltration rate per foot of trench

ET = 2 K10 (Du/2 + Ds) Heff + 2 K15 d2 Heff + 2 K20 d3 H2

ET = 0.0167 cfs/foot of trench

Note: The exfiltration rate values are limited by the District VI Drainage Section to 0.15
cfs/linear foot of trench.

Step 6: Determine the length of exfiltration trench for water quality

Lnet = V Lnet = 283 ft

Chapter 7: Exfiltration Systems 7-23


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

S + 60∙ET∙TT

Safety factor to account for hydro-geotechnical uncertainties (SF ≥ 2): SF = 2

Lreq = SF Lnet Lreq = 566 ft

7.2.4.4 Other Design Methods


Design Curves for Exfiltration Systems. The exfiltration curves provide the ratio
between the trench storage and the exfiltration rate from the trench to the soil. This
method is based on the long-term mass balance of an exfiltration system under local
rainfall conditions. As such, limit your use of exfiltration curves outside of areas where
they have been developed.

Exfiltration Trenches for Discharge Attenuation. With the exception of Miami-Dade


County and Monroe County, exfiltration trenches are not approved for discharge
attenuation by most of the permitting agencies. As such, and only under very special
conditions, the use of exfiltration trenches to attenuate the outfall discharge will be
negotiated with the permitting authorities and with prior approval by the FDOT District
Drainage Engineer.

The exfiltration trench design procedure for discharge attenuation should be similar to the
procedures described above. The difference is the required treatment volume. For closed
basins, the treatment volume should be the pre-development versus post-development
discharge increase instead of the required water quality volume. You could apply the
same criteria for basins with positive outfall if the Rational Method is used because this
method considers the rainfall intensity constant throughout the storm duration. If you use
hydrograph methods, the runoff hydrograph has to be combined with the exfiltration
hydrograph to determine the outflow from the drainage system. Spreadsheets and
modeling programs are available to perform hydrograph calculations.

7.3 DRAINAGE WELLS


7.3.1 Description
The term drainage wells includes all wells that are used to inject surface water directly
into an aquifer, or transfer shallow ground water directly into a deeper aquifer. By
definition, an injection well is any bored, drilled, driven shaft, or dug hole that is deeper
than its widest surface dimension, or an improved sinkhole, or a subsurface fluid
distribution system (refer to Figure 7.3-1).

Chapter 7: Exfiltration Systems 7-24


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Figure 7.3-1: Injection Well

Drainage wells in Florida are grouped into two broad types: surface-water injection wells
and inter-aquifer connector wells. Surface-water injection wells are further categorized as
either Floridan aquifer drainage wells or Biscayne aquifer drainage wells.

7.3.1.1 Use
The Floridan aquifer drainage wells generally are effective as a method of urban drainage
and lake level control. They emplace more recharge into the Floridan aquifer than the
recharge it would receive under natural conditions. The most common use of Floridan
aquifer drainage wells is to supplement surface drainage for urban areas in the karst
terrains of the topographically higher areas of central and north Florida. Be cautious,
however, with regard to the water quality aspects of these wells because they often inject
surface runoff into the same aquifer from which public water supplies are withdrawn.

Chapter 7: Exfiltration Systems 7-25


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

In southeast Florida, drainage wells to the Biscayne aquifer dispose of stormwater runoff
and other surplus water. The majority of these wells dispose of water from swimming
pools or heated water from air-conditioning units. Some of the wells dispose of urban
runoff or wastewaters from business and industry in the area. The use of Biscayne aquifer
drainage wells should have minimal effect on aquifer water quality as long as the injection
of runoff and industrial wastes is restricted to zones where chloride concentrations exceed
1,500 milligrams per liter and 10,000 milligrams per liter total dissolved solids (TDS).

7.3.2 Design Criteria


Designing a storm sewer system using drainage wells requires hydrologic analysis and
determination of the design peak runoff rate of discharge from the project area. See
Chapter 2 (Hydrology) for the procedures to determine the peak runoff rates.

The data collection for drainage well design should include researching similar installed
wells within the project area. Local well contractors can provide an estimate of the
discharge capacity of wells based on previous drainage well installations and pumping
tests. In cases where there is no available data, perform test holes and pumping tests.
The exfiltration capacity of a well will be determined in gallons per minute per foot of head.

7.3.3 Water Quality


Address water quality requirements before discharging into injection wells. The typical
design of drainage wells provides for the use of a retention basin with baffles or skimmers
prior to discharging into the drainage well. Determine the size of the retention basin based
on a 90-second detention time. Other options include the use of exfiltration trenches and
detention/retention treatment swales.

Stormwater pollutant load is very dependent on climatic and topographic features, such
as storm intensity and duration, distribution over the basin, land use, and topographic
features such as hills, swamps, and soil types. Design the type of stormwater treatment
to meet the needs of the particular location (i.e., more stringent water quality measures
should be required for wells discharging into the Floridan aquifer). Pretreatment methods
include physical, chemical, and biological control measures. Physical treatment includes
typical operations like settling and screening. An example of chemical treatment would
be the injection of alum into the stormwater on a storm-by-storm basis. Biological
treatment might be accomplished by using plants, fish, or other types of treatment in
retention ponds. In many instances, a combination of the above methods are used prior
to the discharge of stormwater into the freshwater injection well.

The discharge of the wells needs to occur below the 10,000 ppm Total Dissolved Solids
(TDS) level.

Chapter 7: Exfiltration Systems 7-26


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

7.3.3.1 Fresh Water-Salt Water Hydrostatic Balance


One major consideration in the design of drainage wells in coastal areas is the difference
in density between fresh water and saline water. The hydrostatic balance between fresh
water and saline water can be illustrated by the U-tube shown in Figure 7.3-2. Pressures
on each side of the tube must be equal, therefore:

Ps g Hf = Pf g (Z +Hf)
where:
Ps = Density of the saline water, in lb/ft3
Pf = Density of the fresh water, in lb/ft3
g= Acceleration of gravity, in ft/sec2
Z= Head difference, in ft
Hf = Fresh water height above Mean Sea Level, in ft

Solving for Z, the Ghyben-Herzberg relation is obtained:

Z = (Pf / Ps –Pf ) Hf

Figure 7.3-2: U-Tube Hydrostatic Balance between Fresh and Saline Water

Translating the U-tube to a coastal situation, as shown in Figure 7.3-3, Hf is the elevation
of the water table above the sea level and Z is the depth to the fresh water-saline water
interface below the sea level.

Chapter 7: Exfiltration Systems 7-27


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Figure 7.3-3: Fresh-Saline Water Interface

For typical seawater conditions (Ps = 63.9 lb/ft3 and Pf = 62.4 lb/ft3), the hydrostatic head
balance is approximately: Z = 40 Hf.

The cased portion of drainage wells in coastal areas is placed at least up to the fresh-
saline water interface. You can approximate the head loss due to the difference in the
density of salt water and fresh water as one foot of head loss per 40 feet of casing based
on the relationship Z = 40 Hf. As such, a typical design with 60-foot average casing up to
the interface would require 1.5 feet of head to displace the salt water, which is the usual
rule-of-thumb value used to design drainage wells discharging into the Biscayne aquifer.
This additional head is not required for wells discharging into a freshwater aquifer.

7.3.3.2 Hydraulic Head


The maximum allowable stage upstream of the drainage wells limits the design hydraulic
head for gravity drainage wells. The design high water for the design storm event should
meet the base clearance requirements and cause no adverse impact to the land use value
of the surrounding properties.

In areas where gravity head is not available, you can obtain the required hydraulic head
artificially by pumping. Pressurized drainage wells basically have the same design
requirements as the gravity wells but the hydraulic head is produced by a lift station.
Runoff pretreatment is necessary prior to the lift station, which usually is provided by a
retention basin with baffles and a bar screen for protection of the pumps. Typically, you

Chapter 7: Exfiltration Systems 7-28


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

can combine these features with a finer screen to block smaller debris from discharging
into the well. The FDEP has waived the requirement of the Reasonable Assurance Report
(refer to Appendix K) when a passive by-pass at or below 8.00-foot NGVD ’29 is provided
from the pump’s common header. This requires a vertical stack pipe with a top elevation
at 8.00-foot NGVD ’29. If the head on the pumps is larger, there will be overflow in this
stack. Provide a cap/bird screen to avoid tampering.

7.3.3.3 Safety Factor


Due to some uncertainty related to geological and other factors of drainage wells, a safety
factor of 1.5 is recommended for the design of drainage wells.

7.3.3.4 Dimensions
Drainage (deep) wells usually are 24 inches in diameter and 100 feet to 150 feet deep.
When more than one well is necessary for the system, try to separate them by 75 feet to
100 feet.

7.3.3.5 Exfiltration Rate


A drainage well is drilled as an open hole until the desired level of exfiltration is found,
based on the results of the well test holes and the pumping tests described in Section
7.1.2.4. Common values of well exfiltration rates (in Miami-Dade County) range from 500
gpm to 1,500 gpm per foot of head.

7.3.3.6 Casing
The casing point is usually determined by finding the required minimum total dissolved
solid levels in the aquifer or by finding structurally stable rock formations. Typically, about
70 percent of the well depth is steel encased.

7.3.4 Methodologies of Calculation


The calculation methods to design gravity wells (Section 7.3.4.1) and pressurized wells
(Section 7.3.4.2) are illustrated based on a sample roadway segment (Figure 7.3-4) with
a contributing drainage area of 2.3 acres, including 0.8 acre of pavement to be drained
into an injection well system.

Chapter 7: Exfiltration Systems 7-29


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Figure 7.3-4: Sample Problem

7.3.4.1 Gravity Wells


Step 1: Data collection (refer to Figure 7.3-4)

Total drainage area: A = 2.3 acres Hydrologic location: Zone 10


Pavement area: Ai = 0.8 acre Design frequency: F =3 years
Pervious area: Ap = A - Ai Runoff coefficient
Ground water elevation: SHWT = 1.60 Impervious: Ci = 0.95
Minimum roadway grade: G = 8.00 Pervious: Cp = 0.30

Control elevation: CE = 3.60 ft.

Well design data: (Based on information from other wells near the project site)

Well capacity: QW = 750 gpm/ft of head

Depth to interface: Z = 60 ft

Chapter 7: Exfiltration Systems 7-30


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Step 2: Determine the peak discharge rate into the gravity well system.

Weighted runoff coefficient: C = Ci∙Ai + Cp∙Ap C = 0.53


A

Time of concentration (tc):


Note: You will find methods to calculate the time of concentration in Chapter 2
(Hydrology). Use the Rational Method to solve the sample problem. For larger
projects, you could use the Unit Hydrograph Method to design the system.

Inlet time: ti = 10 min (based on the minimum tc)

Travel time (pipe): tt = Flow path length = 80 ft + 320 ft = 3.3 min


Flow velocity: 2 ft/sec

tc = ti + tt tc = 13.3 min

Rainfall intensity: i= 308.5 i = 5.39 in/hr


– 0.11
(For Miami-Dade County) 48.6 F + TT (0.5895 + F-0.67)

Note: You can determine the rainfall intensity for other project locations using the
appropriate Intensity-Duration-Frequency (IDF) Curves. For IDF curves, go to the
Department’s Internet site at:
http://www.fdot.gov/roadway/Drainage/ManualsandHandbooks.shtm

Peak discharge Q=CiA Q = 6.52 cfs

Note: Consider attenuating the peak discharge when substantial runoff storage capacity
exists or is proposed within the contributing area. In these cases, use the NRCS
(formerly SCS) Method.

Step 3: Calculate the exfiltration capacity of one gravity well.

Effective head: Heff = CE – SHWT – ∆H Heff = 0.50 ft

Note: The head loss (∆H = 1.5 ft.) used above is based on the rule of thumb described
in Section 7.3.3.1. Although usually disregarded, you can consider additional head
loss due to friction along the well casing (approximately 0.001 foot per foot of
casing). You may need a more accurate head loss calculation when dealing with
deep casings and very high flow rates coupled with low available hydraulic heads.

One-well capacity: Qw = 0.00223 qw Heff Qw = 0.84 cfs

Chapter 7: Exfiltration Systems 7-31


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Step 4: Determine the required number of gravity wells.

Safety factor: SF = 1.5

Number of gravity wells: Nw = SF Q Nw = 11.69 wells


Qw

Recommendation: Use 12 gravity wells at a minimum spacing of 75 feet. Distribute


these wells within the project area or at one location. Consider head
losses due to pipe lines to design the stormwater system.

Step 5: Determine the 90-second retention volume for each gravity well.

Required detention volume: V90sec = 90 Q V90sec = 48.9 ft3


(for each gravity well) Nw

Note: Provide the 90-second retention volume right before the drainage wells, below the
top of the well (usually at the SHWT).

7.3.4.2 Pressurized Wells

Step 1: Data collection (refer to Figure 7.3-4)

(The same as Section 7.3.4.1 Step 1)

Step 2: Determine the peak discharge rate into the gravity well system.

(The same as Section 7.3.4.1 Step 2)

Step 3: Determine the net hydraulic head required for a pressurized well system.

Number of pressurized wells: Np = 1 well

Safety factor: SF = 1.5

Pressurized net head: Hp = SF Q + ∆H Hp = 7.34 ft


0.00223 qw H2

Note: The head loss (∆H = 1.5 ft.) used above is based on the rule of thumb described in
Section 7.3.3.1. You can consider additional head loss due to friction along the
well casing (approximately 0.001 foot per foot of casing), but usually you would
disregard it. You may need a more accurate head loss calculation when dealing
with deep casings and very high flow rates coupled with low available hydraulic
heads.

Chapter 7: Exfiltration Systems 7-32


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Required head elevation: Head = SHWT + Hp Head = 8.94 ft NGVD

Note: The sample problem would require approximately 12 wells to function by gravity
(see Section 7.3.4.1, Step 4, above). By pressurizing the system with 7.34 ft head
(at 8.94 ft NGVD), only one well would be needed. Notice that a Reasonable
Assurance Report will be required by FDEP (see Section 7.3.3.2 and refer to
Appendix K) because the head elevation at 8.94 ft NGVD is higher than 8.00 ft
NGVD.

Step 4: Determine the 90-second retention volume for each pressurized well.

Required detention volume: V90sec = 90 Q V90sec = 586.8 ft3


(for each gravity well) Np

Note: The 90-second retention volume has to be provided right before the pump station
or stations for the pressurized drainage wells.

Step 5: Design the pump station.


Required pump discharge: Qp = SF Q Qp = 9.78 cfs

Required head elevation (Step 3 above): HEAD = 8.94 ft NGVD

The pump station would have to deliver a flow of 9.78 cfs with a maximum head of 8.94
ft NGVD. The system design then will consist of pump selection, design of the pump pit,
and the forced line, which is not in the contents of this document. The procedure for the
lift station design could be done by manual methods, spreadsheets, and other computer
software commercially available for lift station design.

7.4 MODELING EXFILTRATION SYSTEMS


7.4.1 Basic Modeling Concepts
Hydraulic modeling idealizes existing and proposed hydraulic systems for calculation
purposes. The hydraulic models generated for a specific project may vary in complexity
based on the accuracy needed by the designer. Several spreadsheets and modeling
software programs are available to design or evaluate open and closed stormwater
systems, but the designers have to use software accepted by the Department and the
permitting agencies with jurisdiction over the area where the project is located.

The main elements in a hydraulic model setup are nodes and links. A node is a point with
a defined location in the drainage system used to simulate inlets, manholes, grade
breaks, bends, outlets, etc. Nodes usually are points that maintain the conservation of
mass during the calculation process. Links are the connections between nodes used to

Chapter 7: Exfiltration Systems 7-33


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

simulate pipes, ditches, and channels. You can use links to transfer or convey water
through the drainage system.

Other important modeling elements are the boundary conditions, which usually simulate
the tailwater conditions as time-stage nodes, and the drainage areas, which are closed
boundaries contributing into their respective nodes. Rainfall and peak runoff
computations, usually related to the drainage areas, are calculated using the Rational
Equation or the NRCS methods.

7.4.2 Exfiltration Trenches Modeling


To simulate the exfiltration trench performance, model the trench pipe as a normal pipe
link between nodes. One of the nodes connected to the trench usually has the stage-area
or the stage-volume characteristics of the exfiltration trench assigned to it. If the modeling
program computes the emptiness of the links, only the voids in the aggregate will be
considered in the stage-area or the stage-volume capacity of the exfiltration trench.
Include a boundary node, usually as a time-stage node, to simulate the design
groundwater elevation, which could be constant or variable with time as per the designer’s
criteria.

It is necessary to include a special link to model the transference of runoff from the node
simulating the exfiltration trench to the node simulating the design ground water (See
Figure 7.4-1). This special link usually is a head-discharge ratio or rating curve, which
could be defined by giving different head values (∆H) to calculate their respective
discharges (Q) using the following equation (refer to Section 7.1.2.3):

Q = (Ku Au + Ks As) ∆H

where:
Q = Flow rate, in cfs
Ku = Unsaturated hydraulic conductivity (cfs/ft2 – ft of head)
Ks = Saturated hydraulic conductivity (cfs/ft2 – ft of head)
Au = Unsaturated flow area (ft2) = Lnet (2Du)
As = Saturated flow area (ft2) = Lnet (2Ds + W)
∆H = Hydraulic head (ft)

If the unsaturated depth of the exfiltration trench (Du) is less than one tenth of the total
trench depth, disregard the unsaturated exfiltration because the unsaturated hydraulic
conductivity usually is several times less than the saturated hydraulic conductivity of the
soils. If the trench width is greater than two times the depth, disregard the exfiltration
through the trench bottom (As = 2Lnet Ds).

Chapter 7: Exfiltration Systems 7-34


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Figure 7.4-1: Sample Nodal Diagram for Exfiltration Trenches

The head-discharge ratio or rating curve may not be accurate if the design ground water
is considered variable and the modeling program uses elevations instead of hydraulic
head for calculations. You can consider the variable ground water to be rising from the
average October ground water elevation to the anticipated ground water elevation at the
end of the design storm event. In these cases, a family of tailwater-headwater-discharge
rating curves is necessary to simulate the transference of runoff from the exfiltration
trench to the ground water. You can calculate the rating curves by giving specific values
to couples of tailwater and headwater elevations to calculate their corresponding head
values (∆H). Having the hydraulic head (∆H), you can calculate the discharges (Q) using
the above equation and the tailwater-headwater-discharge relationship would be
complete.

7.4.3 Drainage Well Modeling


Calculate the inflow discharge using the approved hydrologic method and assign it to a
node that could simulate the retention box. Analyze the drainage well, or a series of them,
using rating curves with an elevation versus discharge relationship of the wells. The rating
curve link will be connected to a time series node representing the water table at the
discharge point. Following is a simple schematic nodal diagram of a well system model.

Figure 7.4-2: Rating Curve Link for Drainage Wells

Chapter 7: Exfiltration Systems 7-35


January 2019

Drainage Design Guide


Chapter 7: Exfiltration Systems

Basin represents the drainage area contributing into the drainage well. Node N-WELL
simulates the retention box prior to discharge into the well. The rating curve link simulates
the well discharge into the water table. Node N-GW is the boundary node representing
the ground water elevation.

Connect the rating curve link to a time series node representing the water table at the
discharge point. If the system includes several drainage wells, you would develop a rating
curve for each well. There are a number of acceptable computer software hydrologic and
hydraulic models capable of analyzing the system.

Chapter 7: Exfiltration Systems 7-36


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

CHAPTER 8: OPTIONAL PIPE MATERIAL

8. OPTIONAL PIPE MATERIAL ............................................................................. 8-1

8.1 Introduction ....................................................................................................... 8-1

8.2 Design Service Life ........................................................................................... 8-1

8.3 Durability ........................................................................................................... 8-5


8.3.1 Project Corrosion Evaluation ........................................................................... 8-5
8.3.2 Project Geotechnical Investigation and Corrosion Tests ................................. 8-8
8.3.3 Project Pipe Service Life Estimation .............................................................. 8-10
8.3.4 Special Cases (Jack and Bore Casings, Ductile Iron Pipe, any Ferrous
Metals) ........................................................................................................... 8-14

8.4 Pipe Structural Evaluation ............................................................................. 8-19

8.5 Documentation ................................................................................................ 8-28


8.5.1 Project Example Considering all Potential Pipe Applications ......................... 8-28
8.5.1.1 Side Drains under Driveways .................................................................. 8-29
8.5.1.2 Cross Drains (including Side Drains under Side Streets) ........................ 8-35
8.5.1.3 Storm Drain ............................................................................................. 8-41
8.5.1.4 Wall Zone Pipe ........................................................................................ 8-43
8.5.1.5 Gutter Drain ............................................................................................. 8-44
8.5.1.6 French Drain ............................................................................................ 8-46

8.6 Specifying Optional Pipe Materials in the Contract Plans .......................... 8-47

Chapter 8: Optional Pipe Material 8-i


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

8. OPTIONAL PIPE MATERIAL

8.1 INTRODUCTION
It is important to consider the array of materials available for culverts. After you complete
the initial hydraulic design, evaluate the list of culvert materials shown in Table 6-1 of the
Drainage Manual to choose among potential options. Chapter 3 of the Drainage Manual
provides the roughness coefficients you can use to evaluate the various materials. Your
evaluation must consider functionally equivalent performance in durability and structural
capacity. If you are constructing a culvert extension, you should match the existing culvert
material to avoid misleading future maintenance assumptions about the type of buried
pipe material. However, if the existing culvert fails a corrosion evaluation when the length
of time of service is factored in or if it shows signs of deterioration, the existing culvert
should be replaced or rehabilitated (e.g., lined).

8.2 DESIGN SERVICE LIFE


The Design Service Life (DSL) is the minimum number of years that a pipe is required to
perform for a particular application in the design of a project. For most applications, a
100-year DSL is required. Specific DSLs for a particular highway type and culvert function
are shown in Table 6-1 of the Drainage Manual. Refer to the example project in Section
8.5 for further guidance on choosing appropriate DSL.

Although Table 6-1 of the Drainage Manual provides comprehensive policy on the
selection of Design Service Life, practical considerations sometimes will override the
guidance material. For instance, gutter drains are listed as a 25-year DSL application, but
if a gutter drain, or any other pipe, is to be located where replacement would require
closure or major traffic disruption during the design life of the facility, then a longer DSL
is appropriate. Any pipe that is beneath or within the soil zone that provides stability to a
structural wall must have a 100-year service life due to the potential for wall damage or
failure and because of the difficulty of replacing that pipe in the future.

Changing the diameter may change the Estimated Service Life (ESL) of concrete and
metal pipe. This occurs because of the change in wall thickness. As the diameter of
concrete and metal pipe increases, so logically does the wall thickness of the pipe. For
concrete pipes, the wall thickness increases as a result of the thickness change in the
cover over the reinforcing wire, and in metal pipes, the increase is due to the thicker gage
metal used for larger-diameter metal pipes.

Chapter 8: Optional Pipe Material 8-1


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Table 6-1 from the Drainage Manual is included on the next page for easy reference by
users of this design guide. Refer to the Drainage Manual for the latest version.

Chapter 8: Optional Pipe Material 8-2


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Table 6-1: Culvert Material Applications and Design Service Life

Chapter 8: Optional Pipe Material 8-3


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Notes for Table 6-1:

1. A minor facility is permanent construction such as minor collectors, local streets


and highways, and driveways, provided culvert cover is less than 10 feet.
Additionally, this category may be called for at the discretion of the District
Drainage Engineer where pipe replacement is expected within 50 years or
where future replacement of the pipe is not expected to impact traffic or require
extraordinary measures such as sheet piling.
2. A major facility is any permanent construction of urban and suburban typical
sections and limited-access facilities. Urban facilities include any typical section
with a fixed roadside traffic barrier, such as curb or barrier wall. Additionally,
rural typical sections with greater than 1,600 annual average daily traffic (AADT)
are included in this category.
3. Temporary construction normally requires a much shorter design service life
than permanent construction. However, treat temporary measures that could be
incorporated into permanent facilities as permanent construction with regard to
design service life determination.
4. Although culverts under intersecting streets (crossroads) function as side drains
for the project under consideration, these culverts are cross drains and you must
design them using appropriate cross drain criteria, not the shorter side-drain
service life criteria. Use Standard Plans, Index 430-022 for end treatment.
5. Replacing this pipe would require removal and replacement of the project’s
pavement or curb.
6. Gutter drains under retaining walls should use a 100-year DSL.
7. F949 PVC pipe has a service life of 100 years. Other PVC pipe has a 50-year
service life. Do not use PVC pipe in direct sunlight unless it meets the
requirements of Standard Specification 948-1.1.
8. Class II HDPE pipe may not be used in the Florida Keys.

9. Any pipes under permanent structures—retaining walls, MSE walls, buildings,


etc.—must use a 100-year DSL.
10. All vertical pipes require resilient connections.
11. Due to the expected high cost of steel pipe, only list steel pipe as an option if no
other pipe material is allowed.

Chapter 8: Optional Pipe Material 8-4


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

8.3 DURABILITY
The requirements for DSL may vary between projects as well as within a project,
depending on the highway functional classification and the application of the culvert.

The projected service life, hereafter referred to as the Estimated Service Life (ESL), of a
culvert is the duration of service time after which significant deterioration is predicted to
occur. After this point, you would need to consider major rehabilitation, lining, or
replacement. For a material to be included in the design of a project, its ESL must meet
or exceed the required DSL.

For metal pipe, the time of first perforation (complete penetration) is the service life end
point. For concrete culvert, the service life ends when the culvert has experienced a
corrosion-related crack in the concrete. The ESL of a specific culvert material is
determined from an evaluation of the corrosiveness, based on the environmental
conditions of both the soil and water, at the intended culvert site.

For plastic pipe (polyvinyl chloride [PVC] and high density polyethylene [HDPE]), the
service life is independent of the environmental conditions. The service life ends when
any crack appears in the pipe. Plastic pipes sometimes crack from initial field loadings,
but can also crack through a creep/rupture mechanism called slow crack growth. The ESL
of plastic pipe is determined by the State Drainage Office rather than by site-specific
corrosion analysis.

8.3.1 Project Corrosion Evaluation


There are several types of corrosion that may occur with metal pipes or culverts
containing steel reinforcement. Some types of corrosion are more severe and you need
to address them in the design stage of a project. You will need to collect environmental
data when designing a culvert system for a specific site. Corrosion rates of culverts
containing metal are governed primarily by the four environmental parameters listed and
discussed below; these site-specific environmental parameters are used to predict the
rate of corrosion and the resultant estimated service life at the site or region of interest:

• pH
• Resistivity
• Chlorides concentration
• Sulfates concentration

Chapter 8: Optional Pipe Material 8-5


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

pH—The measure of alkalinity or acidity. A neutral soil environment has a pH of 7. When


a culvert is placed in an environment in which the pH of the soil is low (≤ 5.0) or high (≥
9.0), the protective layers of the culvert (concrete, galvanizing, aluminizing, etc.) can
weaken, leaving the metal vulnerable to early corrosion. For example, any organic
material from decomposing vegetation will lower the pH of the soil.

Instances of high pH values are extremely rare. Observed pH values in virtually all soils
and waters in Florida are less than 10. pH values between 5.5 and 8.5 are of no concern.
pH values less than 5.5 are common in swampy areas; a pH below 5 is an aggressive
environment for reinforced concrete and a pH below 4 is highly aggressive. Generally, a
low pH is conducive to steel corrosion. Both a low pH and a high pH (> 8.5), coupled with
low resistivity, create a corrosive environment for aluminum.

Resistivity—A measure of the electrical resistance of soils and waters. Resistivity is the
inverse of conductivity. Highly conductive media tend to promote corrosion. Corrosion is
an electrochemical process. For corrosion to occur, charged ions must migrate through
the soil or water from a corroding area (anode) to a non-corroding area (cathode). Soils
with relatively high resistivity values (> 3,000 Ohm-cm) impede the migration of these
ions, which slows corrosion. Environments with low resistivity values (< 1,000 Ohm-cm)
provide an easy path for ions to migrate from anode to cathode, which in turn accelerates
corrosion. In general, clayey soils, organic soils, or chloride-bearing soils would tend to
generate low resistivity values.

Chloride Concentration—A measure of the number of chloride ions present. Chloride


ions react with and break down a protective layer on the surface of metal that otherwise
protects against corrosion. When the chloride concentration is high (> 2,000 ppm), the
protective layer breaks down quickly, leaving the metal vulnerable to corrosion. In
addition, high chloride concentrations result in low resistivity values that allow easy
electrical paths for ion migration and accelerated corrosion. Salt water or brackish water
will be high in chloride concentrations.

Sulfate Concentration—A measure of the number of sulfate ions present. Sulfate can
cause concrete components to deteriorate. If the sulfate concentration is high (> 5,000
ppm), concrete is vulnerable to accelerated deterioration. Sulfate ion concentrations
rarely exceed 1,500 ppm in Florida; therefore, the threat sulfate ions pose is not as
considerable as that of chloride ions.

Elevated chloride values typically are seen only in or near coastal areas. High sulfates
can be seen anywhere, but are more prevalent in coastal areas.

Chapter 8: Optional Pipe Material 8-6


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

There are other factors that may affect service life. These factors are not the primary
factors, and as such are not included in the FDOT Culvert Service Life Estimator (CSLE)
Program. They are mentioned here to alert you to their potential affects.

Microbially Induced Corrosion—Microbially Induced Corrosion (MIC) is the


deterioration of metals resulting from the metabolic activity of microorganisms. MIC
primarily affects metal culverts, but also can affect reinforcing steel in concrete culverts.
Many types of microorganisms can survive in a wide pH and temperature range. MIC
often presents as corroded surfaces covered in slime, black iron sulfide deposits, algal
growth, and as a rotten egg odor. The reactions generally are localized and occur at
cracks, crevices, and welds. Readily available oxygen and organic carbon can increase
the rate of MIC.

Industrial Effluent—Although discharge of industrial effluents to waterways is regulated,


these can occur with accidental spills. Mine tailings or minable geologic formations can
be a source of acidic runoff. Certain land uses—golf courses, dairy farms, farming
operations, coal burning power plants, or cement plants—can all be sources of corrosive
media.

Stray Electrical Current—Electric current in proximity to a pipe can induce corrosion.


Sources of stray current include electrified rail lines, high-tension electric transmission
lines, and cathodically protected gas transmission mains.

Abrasion—Frequent or continuous movement of rapidly flowing, turbulent water


containing a bedload of sands, gravel, and debris can erode protective coatings on pipes
and also erode the pipe material itself. Bedload is the portion of the total transported
sediment that is carried by intermittent contact with the streambed (or culvert invert) by
rolling, sliding, and bouncing. AASHTO’s Highway Design Guidelines (2007) defines
bedload by the two- to five-year return frequency. For these storm recurrences, flow
velocities greater than 5 fps that carry sand bedload are considered abrasive. Velocities
that exceed 15 fps and carry sand, gravel, and rock bedload are considered very abrasive.
The CSLE does not include abrasion, so it is not required to be considered. However, if
you determine that site and hydraulic conditions are likely to produce abrasive flow
conditions, metal pipe suppliers have tables and programs online to estimate loss of wall
thickness due to abrasion.

Chapter 8: Optional Pipe Material 8-7


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

8.3.2 Project Geotechnical Investigation and Corrosion Tests


Because of the varying complexity of projects and soil conditions, it is difficult to establish
a rigid format for conducting subsurface investigations. As stated in the Department’s
Soils and Foundation Handbook, “A subsurface investigation should be performed at the
site of all new structure and roadway construction, and at widening, extensions, and
rehabilitation locations as directed by the District Geotechnical Engineer or project
scope.” Typically, you would perform environmental corrosion tests, as discussed above,
on soil and water at structure locations (e.g., bridge, box culvert, walls), on structural
backfill material, and on subsurface materials along drainage alignments. For drainage
systems parallel to roadway alignments, perform corrosion tests at maximum intervals of
1,500 feet along the project (see Section 3.2.2.6 of the Soils and Foundation Handbook).
To ensure that you collect and analyze sufficient samples, coordinate with the
geotechnical engineer on sample locations and depths. In addition to field review of the
site and the existing culvert conditions, you can use the NRCS Web Soil Survey to help
plan the soils investigation. Soil type parameters—such as pH, steel corrosion potential,
and electrical conductivity—may indicate areas where you should obtain site-specific
information. Test values are seasonally affected by such factors as rainfall, flooding,
drought, and decaying vegetation. Whenever possible, you should perform environmental
tests during periods when no unusual weather conditions exist.

Roadway plans include a “Roadway Soils Survey” sheet, as shown in Figure 8.3-1, which
identifies a range of values of all tests performed. The complete geotechnical report
contains test results for the specific locations sampled, and you can use these data for
culvert analysis. Review the data and correlate them to actual field conditions where
possible. A prediction of the actual service life of a culvert material at a particular site can
be determined by the performance of a similar culvert material in the same or similar
environmental condition. If the test data do not correlate with the observed culvert
conditions, then request additional testing at the site in question. Ultimately, you should
weight conclusive field performance more heavily than predicted service life when field
performance and predicted service life disagree.

Analysis of the test data should take into consideration the most corrosive values of the
native soils. However, with site-specific project environmental test data available, you
won’t need to use the most corrosive individual site data for the entire project or extract
the most aggressive individual parameter results from the testing data to create a worst-
case, project-wide condition. This over-conservatism is unwarranted and unrealistic.
Instead, you can review the soil boring strata and apply test data to those locations that
are most representative of the soil strata and conditions where the corrosion test data
were obtained. There may be particular segments of the project where corrosive
conditions exist, and other segments where the corrosion potential is low.

Chapter 8: Optional Pipe Material 8-8


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Figure 8.3-1

Chapter 8: Optional Pipe Material 8-9


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

8.3.3 Project Pipe Service Life Estimation


Use the CSLE Program and/or the Tables and Figures in Appendix M to determine types
of culvert material that have ESLs that meet or exceed the required DSL. When the DSL,
pipe size, pH, resistivity, chlorides, and sulfates are input, the program provides a listing
of those materials that meet the DSL. Using the program also provides you with an
excellent form of documentation.

An example of the CSLE input data and printout follows:

DSL: This application is to be a storm drain system that is located on a major urban facility
and will function as an urban principle arterial road. The appropriate DSL for this
application is 100 years.

The following data were furnished and a field review gave no indication that these values
were suspect:

pH: 7.6

Resistivity: 2,610

Chlorides: 2,390

Sulfates: 1,120

Diameter (Pipe Size): 42 inches; because this is a storm drain system, an n-value of 0.012
was used.

Figure 8.3-2 illustrates materials you would use in performing the structural analysis.

Chapter 8: Optional Pipe Material 8-10


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Figure 8.3-2

Note that the ‘Hint’ field changes while this screen is open. The colored lines indicate
warnings: a yellow highlight indicates a warning concerning environmental data; a red
highlight indicates a warning based on the structural data [Appendix C of the Drainage
Manual].

If any one of the lines is double-clicked, another window pops up with more detailed
information about the particular material option. The program allows you to print and save
a file with a unique name; this allows you to maintain the corrosion parameters and site
data, but vary the pipe size or roughness coefficient, re-analyze and print to unique output
files. The printouts are very useful for documentation of the analyses. See Figure 8.3-3.

Chapter 8: Optional Pipe Material 8-11


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Figure 8.3-3

Chapter 8: Optional Pipe Material 8-12


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Give no additional consideration to one material over another having less service life if all
meet the minimum required. In looking at the printout, culverts made from steel-reinforced
concrete, aluminum, aluminized steel, polypropylene, and high density polyethylene class
II all meet the 100-year DSL. If the DSL was 50 years, then additional materials may be
allowable.

Pipe size affects materials by invoking different gages of metal pipes. Additionally, pipe
size affects the life of reinforced concrete pipe because, with larger diameters, the wall
thickness increases, as does the cover over the reinforcing steel. See Figure 8.3-4 for
allowable materials when the pipe size is reduced to 24 inches, given the same corrosion
parameters and DSL as the previous example. Reinforced concrete pipe is no longer
allowable, whereas non-reinforced concrete pipe is allowable. The gage of spiral-ribbed
aluminized pipe (SRAP) has changed because of the smaller diameter, and PVC is now
an allowable option because it is available in the smaller diameter.

Figure 8.3-4

Chapter 8: Optional Pipe Material 8-13


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

8.3.4 Special Cases (Jack and Bore Casings, Ductile Iron Pipe, any
Ferrous Metals)
When you choose to install a culvert by jacking and boring instead of open cutting, use
the jacked and bored casing as the conveyance pipe except under railroads or in high-
pressure designs. See Standard Plans, Index 430-001 for railroad company
requirements.

You should specify jack-and-bore installation on high AADT roadways, railroads, or in


areas where open cut for installation or repair causes significant impacts on users. If the
need to install a culvert using jack-and-bore technology was determined after the roadway
soils investigation was made, additional soil borings may be necessary. Determine soil
conditions along the jack-and-bore alignment so that you can evaluate the feasibility of
jack-and-bore installation or of micro-tunneling. You will need corrosion data to estimate
service life.

Because jack-and-bore locations typically have a high AADT, this service life estimation
example will assume a 100-year DSL. Use the following steps to determine the casing
requirements.

1. Run the Service Life Estimator Program or use the figures or tables in Appendix M
with site-specific environmental parameters. If the casing or pipe will be exposed
to water (surface or ground) for extended periods of time, compare the
environmental parameters of the water with those of the soil. Use the test results
that produce the shortest service life for the galvanized steel option. Note that
when using the CSLE Program, reduce the DSL as needed for the galvanized steel
option (corrugated steel pipe [CSP] or spiral rib steel pipe [SRSP]) to show up as
an allowable option. Although the required DSL may be 100 years, you must first
obtain the service life for the particular corrosion parameters, or you can use Figure
M1 to determine estimated service life.

2. To be conservative, deduct 10 years from the ESL of the galvanized steel option
generated by the program (or determined by service life tables/figures).

3. Determine the pitting rate by dividing the wall thickness of the galvanized steel
option estimated by the program by the estimated service life determined in Step
2 (ESL – 10 years). From the Drainage Manual, Appendix C, identify the wall
thickness of the gage pipe called out on the output.

Chapter 8: Optional Pipe Material 8-14


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Gage Thickness 0.xxx inches


Pitting Rate = =
ESL (years) year

Knowing the pitting rate, you can determine the required wall thickness by multiplying the
DSL for the application by the pitting rate.

Required wall thickness = Pitting rate x (DSL)

Using the galvanized steel option shown on Figure 8.3-5:

DSL lowered until the


Galvanized Steel
option appears

Figure 8.3-5

Estimated service life for galvanized steel culvert = 57 years

Deduct 10 years from this: 57 years – 10 years = 47 years

Chapter 8: Optional Pipe Material 8-15


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

14 gage = 0.079 inches (thickness of 14 gage galvanized steel culvert per Drainage
Manual, Appendix C)

Therefore, the pitting rate = 0.079/47 = 0.00168 inches/year

Required wall thickness = 100 years (DSL) x 0.00168 (pitting rate) = 0.1681 inches

Note that the pitting rate determined when other DSLs are input and resultant service
lives are obtained will be approximately the same. For example, if a DSL of 25 years is
input in the above example, the gage of SRSP allowed is 16 (0.064 inch), with a
corresponding service life of 46 years. This results in a pitting rate of 0.00177 inches/year.

In summary, you would need to use a steel casing with a wall thickness of at least 0.17
inches. In the plans, include a note such as: “For corrosion purposes, steel casing must
have a minimum wall thickness of 0.17 inches.”

The required wall thickness is for corrosion purposes only. Typically, you will need greater
wall thicknesses for the structural loadings associated with the jacking of the steel casing.
The CSLE program has an option for “Jack and Bore Steel Casing Pipe” under “Analysis”;
the window will show the approximate wall thickness of pipe suitable for jack and bore
and the associated service life. Check the pipe size in the pop-up window; it may not be
the size input for the initial analysis. The metal thickness shown is that of a typical steel
pipe that you would use for jack and bore. The program uses the thickest galvanized steel
pipe gage to determine the pitting rate (even if that particular gage is not available in the
given size). That pitting rate is applied to the typical jack and bore pipe wall thickness to
estimate service life. If the ESL of the jack-and-bore casing pipe is less than that required,
input the service life in the portion of the window that says “Calculate Metal Thickness…”
and then press return/enter on the keyboard for the required thickness to meet service
life. The print output from this pop-up window contains only the pitting rate analysis, not
the pipe size, nor the corrosion data; so a screen print is a better option for documentation.

The minimum thickness that meets service life is that determined by the pitting rate
equation. Show the thickness in the construction plans with a note stating it is the
minimum thickness to meet service life. The contractor is responsible for determining the
wall thickness required for the jack-and-bore pipe to meet Specification 556.

For the example shown in Figure 8.3-6, the minimum metal thickness based on service
life is thicker than the typical jack-and-bore pipe because of the aggressive
environmental parameters.

Chapter 8: Optional Pipe Material 8-16


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Figure 8.3-6

The printout for this jack-and-bore service life analysis is shown in Figure 8.3-7.

Chapter 8: Optional Pipe Material 8-17


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Figure 8.3-7

When using the casing alone is not allowed, you should place a note disallowing this
practice in the plans to communicate to the Contractor that a VECP (Value Engineering
Change Proposal) eliminating the interior pipe will not be approved.

Chapter 8: Optional Pipe Material 8-18


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

8.4 PIPE STRUCTURAL EVALUATION


After performing the corrosion analysis, the next step in determining the allowable
optional material is to determine the acceptability and structural adequacy of these
materials. If the pipe is within a walled embankment area—“Wall Zone” as illustrated in
the Drainage Manual, Appendix D—then the pipe material considered must be within the
Wall Zone Pipe column in Table 6-1 of the Drainage Manual. All acceptable material types
must be evaluated for anticipated loads on the pipe. The Drainage Manual, Appendix C,
contains cover height tables for the various pipe materials. The information provided in
Appendix C was developed based on criteria found in AASHTO LRFD Bridge Design
Specifications, Section 12.

For each of the acceptable pipe materials based on the corrosion analysis and pipe
location within the embankment, verify that the depth of backfill over the pipe is between
the minimum and maximum fill heights in the appropriate table in the Drainage Manual,
Appendix C. If the cover height is outside the limits, the following options are available:

1. Adjust the flow line of the pipe as long as this adjustment does not violate any other
design criteria.

2. Increase the gage of metal pipe or the class of concrete pipe. For metal pipe, verify
that the specified gage thickness is available for the corrugation specified.

3. Eliminate the material as an option for the job.

The FDM requires that you call out all the acceptable types of pipe materials in the plan.
You can establish the required class of concrete, or gage and corrugation for metal pipe,
using the CSLE program or the service life tables/figures in Appendix M, and the Drainage
Manual, Appendix C. The tables in Appendix C have been incorporated into the CSLE;
however, you will need to back check the results of the CSLE structural check against the
tables in Appendix C for final verification. Generally, it is more efficient to look at the tables
when determining the structural suitability than to input discrete height values in the
Structural Check option of the CSLE. The tables allow you to readily see the lower and
upper limits of allowable cover, whereas the CSLE output provides only a “pass” or “fail”
for a particular value.

For example, given the allowable pipe materials shown in Figure 8.3-2, find the pipes that
are structurally sufficient for a minimum fill height of 23 inches and maximum fill height of
25 feet.

Chapter 8: Optional Pipe Material 8-19


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Using the CSLE program with corrosion data and DSL shown on Figure 8.3-2, click on
the box next to “Structural Check.” A window pops up for input of pavement type and
cover thickness. The cover thickness here is input to the flowline of the pipe whereas the
cover is shown to the outside crown of the pipe in the Appendix C tables. Input a pipe
depth of 68 inches (from finished grade to pipe flowline), and 3 inches for the thickness
of flexible asphalt pavement. The CSLE program results in the elimination of HDPE and
PP pipe (Figure 8.4-1). Referring to the Drainage Manual, Appendix C, the table for Plastic
Pipe, you can see that the minimum cover from top of base course to top of pipe is 24
inches for both HDPE and PP pipe. Based on our inputs to the CSLE program, the cover
is [(68-3)-(42)] = 23 inches, not including deduction for pipe wall thickness. The Appendix
C table for SRAP shows minimum cover of 21 inches for the 42-inch diameter, 14 gage
pipe. Referring to the respective tables for the remaining allowable materials, we find that
12 inches is the minimum cover for both round and elliptical concrete, as well as for the
SRSP (SRASP has the same structural properties as SRSP).

Figure 8.4-1

Chapter 8: Optional Pipe Material 8-20


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Minimum cover and maximum fill heights obtained from the tables in the Drainage
Manual, Appendix C, are summarized in Table 8.4-1 for the acceptable materials shown
in Figure 8.3-2.

Table 8.4-1 Example Project Minimum Cover and Maximum Fill Heights
Allowable material for 42-inch Appendix C Appendix C
diameter, minimum DSL = 100 Allowable minimum Allowable maximum fill height (ft.)
years cover (in.) to finished grade
14 gage SRAP 21; from top of base 25 ft; special installation
Reinforced Concrete Pipe
12; from finished grade 21 ft to 33 ft; CL IV pipe required
(RCP); Typical Dry Cast
RCP; Elliptical Only 12; from finished grade max 25 ft; CL HE-IV required
14-gage SRASP
12; from top of base 54 ft
(use SRSP table)
HDPE CL II 24; from top of base 13 ft

Polypropylene 24; from top of base 15 ft

Preparing a table with the allowable materials and their minimum cover and maximum fill
height allows one to quickly ascertain where the materials can be used within the project
if locations of the minimum cover and maximum fill are known. Plastic pipe would not be
acceptable for installation where the minimum cover is less than 24 inches, and fill heights
are greater than 13 feet, but there may be many locations throughout the project where
plastic pipe installation would be within the allowable structural limits. That is why it is
more efficient to use the tables directly rather than use the CSLE for the structural check.

Note that the fill heights shown in the Drainage Manual, Appendix C, are calculated using
a very conservative approach. In those cases where you encounter very high or very
shallow fill heights, you can use methods set forth in AASHTO LRFD Bridge Design
Specifications, Section 12.Where you must locate pipes within close proximity to walled
embankment areas of any type, review the figures in the Drainage Manual, Appendix D,
to determine what limitations are imposed on pipe location and material. The figures in
Appendix D show Wall Zones A, B, and C. Wall Zone criteria allow both longitudinal and
transverse Wall Zone Pipe (as listed in the Drainage Manual, Table 6-1) in Wall Zone A.
You are allowed to use Transverse Wall Zone Pipe conveyances in Wall Zone B, and you
may not use pipe conveyances of any type in Wall Zone C. A few of the figures in
Appendix D are reproduced here, with examples of where you may place pipes. Wherever
possible, it is best to avoid pipe placement in any of the Wall Zones.

Chapter 8: Optional Pipe Material 8-21


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Figure 8.4-2 shows MSE wall at a bridge abutment on shallow foundation and on deep
foundation. Wall Zone B extends under the deep foundation whereas Wall Zone C (no
pipes) extends under the shallow foundation. The figure shows pipes only within the
allowable zones.

Chapter 8: Optional Pipe Material 8-22


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Figure 8.4-2

Chapter 8: Optional Pipe Material 8-23


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Another example of pipe within MSE Wall Zone is shown in Figure 8.4-3. Longitudinal
conveyances are shown in Wall Zone A. Even though the pipe to the left of the wall is
only partially within Wall Zone A, it must meet the Wall Zone pipe requirements. If right of
way allows, this pipe should be aligned fully outside of Wall Zone A. The pipe that is within
the wall fill embankment can run longitudinally only within the top five feet. Minimize
longitudinal runs of pipe in Wall Zone A to the greatest extent practicable. Where inlets
are required that would extend into Wall Zone B, the preference would be to outfall
transversely to a trunk line located outside of the wall zones. If this is not feasible, then
you could use a deeper structure to allow the pipe to outfall transversely through or under
the wall; these configurations are not ideal. Any structure or pipe within the reinforcement
strap zone must be coordinated with the wall designer.

Chapter 8: Optional Pipe Material 8-24


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Figure 8.4-3

Chapter 8: Optional Pipe Material 8-25


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Figure 8.4-4 shows a gravity wall and its associated wall zones. Ascertain the wall scheme
proposed (there are other schemes) to ensure that any proposed drainage structures
meet the wall zone criteria.

Figure 8.4-4

Chapter 8: Optional Pipe Material 8-26


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Figure 8.4-5 shows wall zones associated with a cantilever wall. Note that this
configuration, as well as all supporting walls on shallow foundation, has a no-pipe zone,
i.e. Wall Zone C, which is directly under the structure and extends out in a trapezoidal
shape below the structure. The depth of the trapezoid is dependent upon the particular
structure.

Figure 8.4-5

Chapter 8: Optional Pipe Material 8-27


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

8.5 DOCUMENTATION
You are required to provide justification if you allow or eliminate a pipe material. You can
find documentation requirements in Chapter 6 of the Drainage Manual. Requirements
include the Design Service Life required for the application, environmental data, and the
results of the structural evaluation.

The CSLE program provides an excellent form of corrosion analysis documentation in the
printout. The printout documents the site specific environmental parameters, the ESL,
and the materials that fail to meet or exceed the DSL. Also, you can add further comments
for documentation purposes. An example of additional comments would be: “not allowed
per Drainage Manual, Appendix C,” “minimum cover not available,” or “maximum cover
exceeded.”

8.5.1 Project Example Considering all Potential Pipe Applications


The project consists of widening and resurfacing a state road in northern Leon County,
Florida. This particular section of roadway contains both rural and urban sections. The
urban section occurs where the roadway approaches and crosses an arterial roadway.
The AADT for the roadway within the project limits is 1,500. The roadway project includes
widening for bike lanes and turning lanes. You conducted a field review prior to final
design; you observed all culverts and side drains for signs of deterioration, siltation, and
erosion. All were in reasonably good condition given their current 40-year time of service.
The project design includes the following applications:

• Side drain
• Cross drain (replacement and extensions)
• Storm drain
• Wall Zone Pipe
• Gutter drain
• French drain

For this example, each application will be addressed and will include a determination of
the design service life, commonly asked questions, and proposed solutions to those
questions.

Chapter 8: Optional Pipe Material 8-28


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

8.5.1.1 Side Drains under Driveways


Due to widening of the roadway, all existing side drains are affected. Referring to Table
6-1, we locate the Side Drain column. From the table, we see that all highway side drains
require a 25-year design service life, and that all but three of the listed materials
(fiberglass, steel (J&B), and ductile iron) may be applicable.

Check the hydraulics at typical locations to determine if materials with N-values of 0.020
or greater should be included. Generally, if the hydraulic evaluation indicates the structure
is outlet controlled, only those materials with N-values equal to 0.012 need be considered.
In this case, the roadside ditches have minimal longitudinal slope and hydraulic
evaluations of a typical location showed the culvert operated in outlet control, so only
materials with N = 0.012 are suitable hydraulically. (See Cross Drains for side drains
under side streets.) The design calculations result in pipe sizes that include 18-inch, 24-
inch, and 30-inch.

Soil corrosion data obtained at shallow depths along the project were fairly consistent and
are shown in Table 8.5-1.

Table 8.5-1
Station Boring # pH Resistivity Chlorides Sulfates
527+50 A-1 5.6 32000 20 108
592+00 A-2 6.6 9500 20 118
610+00 A-3 6.8 17000 20 20

You can enter the information for each site into the CSLE program, or you can use the
tables or figures in Appendix M to determine suitability for the 25-year DSL. Let’s use the
figures to more quickly evaluate these three sets of test data. Looking at the tables and
figures, we can see that the environmental parameters that affect steel and aluminum are
pH and Resistivity. Those that affect reinforced concrete are pH, Chlorides, and Sulfates.
This example shows some of the Figures and Tables in Appendix M as Figures 8.5-1
through 8.5-4.

From Figure 8.5-1, we see that—as long as the pH is above 5.5 and the resistivity is
above 9,000—the DSL of 25 years is met for 16-gage galvanized steel. For 16-gage
aluminized steel, Type II, we see on Figure 8.5-2 that the 25-year service life is met for
pH between 4.5 and 9, with resistivity greater than 1,500. Figure 8.5-3 reflects the service
life of 16-gage aluminum pipe; we can see that low resistivity values (<5,000) with pH
values lower than 6 or higher than 8 adversely affect the service life of aluminum. In our
case, the pH values range from 5.5 to 6.8 and the resistivity values are all greater than
9,000; therefore, the required DSL is met.

Chapter 8: Optional Pipe Material 8-29


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

We will use the table in Figure 8.5-4 to evaluate the suitability of reinforced concrete pipe.
The table shows that service life decreases with increasing chloride concentrations and
as pH drops below 6. If sulfate concentrations go above 1,500 ppm, the service life should
be discounted as noted. Since all chloride values from the samples are 20 ppm or below,
there is no adverse effect on reinforced concrete. However, the table values are for 60-
inch pipe, which has a thick pipe wall. To estimate service life for 18-inch pipe, the service
life of 360 years must be multiplied by 0.36. The minimum service life anticipated for
reinforced concrete pipe on this project is, therefore, approximately 130 years.

Figure 8.5-1

Chapter 8: Optional Pipe Material 8-30


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Figure 8.5-2

Figure 8.5-3

Chapter 8: Optional Pipe Material 8-31


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Figure 8.5-4

All roadside ditches are 3.5 feet below the roadway edge of shoulder, and ground at the
right-of-way line is at or above the edge of shoulder, so available depth is 3.5 feet.
Referring to the cover height tables in the Drainage Manual, Appendix C, we see that
plastic pipe up to 48 inches in diameter requires 24 inches of cover below the top of base
course under flexible pavement. SRAP requires 12 inches of cover for pipe diameters up
through 24 inches and 15 inches of cover (below the top of base course) for 30-inch pipe.
SRSP requires 12 inches of cover below the top of base course for pipe diameters up
through 48 inches. Concrete pipe requires 12 inches of cover from finished grade of
flexible pavement.

For 18-inch pipe, 24 inches is available from inside crown of pipe to finished grade, so
the plastic pipe minimum cover is not met. Where side drains are 24-inch diameter, there
is 18 inches from finished grade to inside crown of pipe and 15 inches from the top of
base course. So for 18-inch and 24-inch side drains, you are allowed to use all pipe
materials with an N value of ≤ 0.012 except plastic. For 30-inch diameter side drains,
there is 12 inches of cover from inside crown of pipe to finished grade; therefore, the
SRAP and SRSP pipe cover requirements are not met. Concrete pipe wall is much thicker
than other types of pipe materials and must be taken into account. In this case, there will
be only 8.5 inches of cover on 30-inch round concrete pipe. The wall thickness for elliptical
concrete pipe is slightly greater for a size equivalent to round concrete pipe. A 24-inch by

Chapter 8: Optional Pipe Material 8-32


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

38-inch elliptical concrete pipe wall is 3.75 inches thick, so there is 14.25 inches of
clearance (42 – 27.75). The elliptical concrete pipe is the only pipe that will meet the
structural requirement where a 30-inch side drain is needed.

Note that if the structural check is used in the CSLE program, it does not use the 30-inch
round equivalent dimension (24-inch by 38-inch) for the elliptical pipe; it uses the 30-inch
diameter input and will, therefore, show that all pipes will fail the structural check where
there is 42 inches from pipe invert to finished grade under flexible pavement. (See the
CSLE output for 24-inch and 30-inch pipe in Figure 8.5-5 and Figure 8.5-6.) So, for pipe
arch or elliptical pipe, it is best to use the cover height tables and to calculate the cover
available based on pipe dimensions.

Figure 8.5-5

Chapter 8: Optional Pipe Material 8-33


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Figure 8.5-6

To present the results in the plans, use notes under the side-drain table stating the
allowable materials. In this case, an appropriate note might read:

Allowable pipe materials for 18-inch and 24-inch side drains, unless otherwise
noted, are: RCP, NRCP, SRSP, SRASP, SRAP and ERCP. Allowable pipe
material for 30-inch side drain is ERCP (30-inch-other) only.

See Section 8.6 for examples of plan quantity presentation.

In cases where there is minimal cover and the structural requirements could not be met
by using elliptical pipe or pipe arch in lieu of the hydraulically required round pipe, then
you will need to analyze alternate pipe configurations. This could include multiple smaller-
diameter pipes or, possibly, a larger diameter pipe buried deeper so that the flow area is
from normal ditch line to crown. The latter also may require adjustment of the roughness
coefficient in the analysis.

For example: if the cover condition at the side drain resulted in less than 12 inches, even
with the elliptical concrete pipe, two 24-inch pipes could be used as long as they fit within
the ditch. The Mitered End Section (MES) width for a double 24-inch pipe installation is

Chapter 8: Optional Pipe Material 8-34


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

8.92 feet out to out (Standard Plans, Index 430-022). A 42-inch pipe half buried has the
equivalent capacity to a 30-inch (fully open) pipe if the fill has a roughness coefficient only
slightly greater than the pipe wall. This installation would be 20 inches narrower than the
double 24-inch, with an MES width of 7.25 feet. If you were to use this option, place a
note in the plans stating that the pipe size is hydraulically necessary and used to meet
cover and dimensional limitations. Additionally, you would need to specify the fill.

8.5.1.2 Cross Drains (including Side Drains under Side Streets)


A cross drain conveys flow under a public roadway. A side street that crosses over a
roadside ditch is, therefore, subject to cross drain design criteria, both hydraulic and
structural. According to the Drainage Manual, Table 6-1, the minimum DSL for a cross
drain is 50 years. That minimum applies to minor collectors and local streets, provided
culvert cover is less than 10 feet. All other cross drains must have a DSL of at least 100
years. If the cross drain hydraulics show that the structure is outlet controlled, then you
may consider only pipe materials with N = 0.012. However, if the cross drain is in inlet
control, materials with higher roughness coefficients should be included in the analyses.
Where pipe options are limited by minimum cover requirements, consider using multiple
smaller-diameter pipes that have sufficient cover as an alternative to a single larger-
diameter pipe that requires less cover. However, multiple pipe configurations are more
susceptible to debris problems and also may require more extensive endwalls. If you have
particular concerns about allowing multiple pipes in lieu of a single pipe cross drain
installation, then you should document the rationale for selection of a particular pipe
configuration that limits material options.

For the project example, these are local streets with AADT less than 1,000, so the
side/cross drains must meet a 50-year DSL. The pipes for the four locations where the
side streets cross roadside ditches were hydraulically checked to ensure that the
appropriate design frequency flow could be passed without damage to the roadway or
offsite properties. The side drains at these locations are 24 inches in diameter. You can
classify these “side” drains as cross drains because they are under public roads.
Therefore, these structures are not included in the side drain summary table, but are
instead included with structure numbers in the Summary of Drainage Structures table.
The materials for these structures were checked using the CSLE to determine which met
the 50-year DSL. All materials previously determined are acceptable, but the SRSP has
changed from 16-gage to 10-gage so that the service life can be met.

Chapter 8: Optional Pipe Material 8-35


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Figure 8.5-7

There are two cross drains that convey offsite runoff under the design roadway. One of
the cross drains is located within a sag vertical curve in the rural section of the design
roadway; the other is located within the urban section.

The location of the first cross drain has a history of roadway over-topping due to basin
diversions and increased runoff from upstream development. It is a 36-inch corrugated
galvanized steel culvert with straight endwalls that are located 18 feet from the edge of
travel lanes. During the Project Development and Environmental (PD&E) study, it was
determined to raise the roadway profile, along with cross drain replacement, to minimize
risk to motorists. The location of this cross drain warrants a DSL of 50 years. The new
cross drain analysis determined that the cross drain is inlet controlled; therefore, both
rough and smooth wall pipe materials were considered. The hydraulic analysis showed
that a 54-inch culvert opening is needed to pass the design discharge of 160 cfs.

The location of the second cross drain, a 36-inch RCP, has exhibited no hydraulic
insufficiencies. Since it is within an urban section, it should have a DSL of 100 years. The
cross drain was analyzed hydraulically and this analysis determined that you could extend
the cross drain with no adverse upstream effects.

Chapter 8: Optional Pipe Material 8-36


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Specific environmental data were obtained at the site of the two cross drains as follows:

Table 8.5-2
Station Boring # pH Resistivity Chlorides Sulfates
505+00 Rural A-4 5.2 17,000 51 6
589+00 Urban A-5 6.9 18,000 42 6

The CSLE program was used to find culvert materials that meet the DSL and structural
requirements for the 54-inch culvert. The depth from finished grade to flow line of pipe is
96 inches. The CSLE output file is shown in Figure 8.5-8. Figure 8.5-8 does not contain
all data for each metal option; see the screen shot of the output in Figure 8.5-9.

Chapter 8: Optional Pipe Material 8-37


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Note that the 54-inch


option with Max N =
0.020 yields 10
acceptable pipe
material/thickness
options. All data about
culvert type do not
appear; for instance,
the two lines that show
CSP, 10 GA., also
include the corrugation
of the material within
the active CSLE
window.

Figure 8.5-8

Chapter 8: Optional Pipe Material 8-38


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Note that the


corrugations
for the CAP,
CSP, and
CASP are
shown here,
but not in
print-out.

Figure 8.5-9
So, there are 10 options for CD-1 at station 505+00. There are metal pipe arch options
that you also could include. However, it is unlikely that the elliptical or arch options would
be economical to construct. Generally, round pipe is easier to construct, so only the round
options need to be shown. If there were utility conflicts, then they could be avoided with
a maximum pipe height of 48 inches; then the 3” x 1” Corrugated Aluminum Pipe Arch,
16 gage pipe; the Corrugated Steel Pipe Arch, 10 gage in both the 2-2/3” x ½” and 3” x
1” corrugations; and the Aluminized Corrugated Steel Pipe Arch, 12 gage, would be
acceptable alternates.

You can extend the second cross drain and still meet the hydraulic requirements. Check
to verify that this cross drain still will have the required DSL. The corrosion data obtained
for this location were input to the CSLE program, along with the depth of cover. The
results, shown in Figure 8.5-10, show that the RCP has a service life of 360 years.
However, Table M-4 of Appendix M indicates that you should adjust the service life of 360
years by multiplying by 0.54 for 36-inch diameter pipe. This results in 194 years. This pipe
has been in service only 40 years, so the DSL of 100 years is met.

When extending cross drains, try to use the same existing pipe materials. If, upon
inspection, the existing pipe shows corrosion or has structural cracking, then you should
replace or line the existing pipe. When the extension of an existing pipe results in a minor

Chapter 8: Optional Pipe Material 8-39


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

transgression of the structural clearance criteria, consider providing additional structural


support for the pipe extension rather than replacing the entire cross drain. Encasing the
extension in flowable fill typically provides the needed additional support.

Figure 8.5-10

Chapter 8: Optional Pipe Material 8-40


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

8.5.1.3 Storm Drain


Referring to Table 6-1 in the Drainage Manual, we find that storm drains require either
50-year or 100-year DSL. A combination of the two DSLs could exist within a project. An
example would be where the main storm drain has to be designed to meet the 100-year
DSL and you could design the outfall to meet the 50-year DSL.

When choosing the appropriate DSL, use the same steps as those previously stated.
Remember, all storm drains do not require the 100-year DSL criteria. Refer to the notes
on Table 6-1 for guidance on the selection of DSL.

The 100-year DSL is required for our example project because the storm drain system is
located within a curb-and-gutter section (see Note 2, Table 6-1). The corrosion data
produced by the geotechnical survey were correlated to the field review observations and
these values are not suspect. The corrosion data values are from the most aggressive
test site (not the most aggressive parameter value from all test sites) along the applicable
subsection of the project.

Table 8.5-3
Station To Station Boring # pH Resistivity Chlorides Sulfates
550+00 630+00 B-1 thru B-5 5.2 17,000 51 6

Because this is a storm drain system, only the smooth wall pipe options may be
considered. The corrosion test data were input to the CSLE program for pipe sizes of 18,
24, 30, and 36 inches. Minimum cover is 25 inches and maximum cover is 11 feet. The
CSLE program provided the materials described below for both 18-inch and 24-inch pipe
sizes (Figure 8.5-11). For 30-inch and 36-inch, 12-gage SRAP is an additional option. A
thicker gage of SRAP is needed because of the low pH and 12 gage is not available in
diameters of less than 30 inches. See Figure 8.5-12 for a screen shot of the CSLE
program for 36-inch pipe.

Chapter 8: Optional Pipe Material 8-41


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Figure 8.5-11

Figure 8.5-12

Chapter 8: Optional Pipe Material 8-42


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

8.5.1.4 Wall Zone Pipe


The example project contains an area of elevated, MSE-walled embankment for a divided
highway with median barrier wall in super-elevation. The roadway profile is at 0.5 percent
and the storm drain piping will follow the slope of the roadway. The roadway storm drain
system outfalls under the wall to a shoulder gutter inlet (SGI) within a parallel storm drain
system. To ensure that all piping meets the Wall Zone Pipe material restrictions and
requirements, sketch the wall zones on the drainage structure sections (see Figure 8.5-
13). You can see that the longitudinal pipe coming into the median barrier inlet is within
Zone A. The 18-inch lateral pipe from the barrier wall inlet along the MSE wall goes
through Zone B transversely and through Zone A.

Figure 8.5-13

Zone C is below the median barrier cantilever wall, which is back and ahead of the median
barrier inlet. The pipe options, as shown in the Drainage Manual, Table 6-1, are:
polypropylene, PVC, and J&B Steel. The steel pipe will cost substantially more than the
PVC or polypropylene, so it would be included as an option only if structurally necessary.
In this case, the 24-inch pipe from the barrier wall inlet to the SGI has 13.5 feet of cover,
so polypropylene may be acceptable. However, because this pipe is under the wall with
roadway on both sides of the wall, you are encouraged to use the methodology in the
AASHTO LRFD (Load Reduction Factor Design) Bridge Design Specifications, Chapter
12, to ensure that the pipe installation will withstand the load conditions. Consider the

Chapter 8: Optional Pipe Material 8-43


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

need for resilient connections at structures, particularly where there may be some
differential settlement.

8.5.1.5 Gutter Drain


From Table 6-1 in the Drainage Manual, you identify a required DSL of 25 years for the
gutter drain. The process is the same as for performing the analysis for the side drain
application discussed previously. However, when sizing gutter drain and choosing
materials, only use materials having an N-value of > 0.020. A gutter drain is defined as a
pipe used along steep slopes to convey stormwater from shoulder gutter inlets on
elevated roadways to drainage conveyance systems at a much lower elevation. These
pipes should be configured so that they can be replaced without disturbing the roadway
and so that they are not placed too deep within the embankment to prohibit future
excavation. Minimize joints where possible. See Figure 8.5-14 for an illustration of gutter
drains.

Chapter 8: Optional Pipe Material 8-44


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Figure 8.5-14

Chapter 8: Optional Pipe Material 8-45


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

8.5.1.6 French Drain


A French drain is used for stormwater treatment and/or attenuation. The Drainage
Manual, Table 6-1, shows that either a 50-year or a 100-year DSL is used for French
drains. The location of the French drain system determines which DSL to use. Consider
a case where you place a French drain in an urban location along the trunk line located
under the sidewalk, parallel and adjacent to the roadway. The French drain is not under
the roadway, but replacement of the French drain would require reconstruction of the
outside lane due to the depth of cut and angle of repose of the soil. Even though the
French drain reconstruction might be performed using sheeting to avoid impacting the
roadway, the cost of the sheeting makes this installation expensive enough to elevate the
service life to 100 years. A similar situation occurs when a pipe installation is adjacent to
buildings. In these cases, sheeting required during replacement is costly; thus, the pipe
should have a longer, 100-year DSL. Conversely, if the French drain is located in a swale
along a rural roadway, the lower DSL may be appropriate.

Figure 8.5-15

For these applications, consider whether a roughness coefficient ≥ 0.020 will result in a
hydraulically acceptable design. Where the French drain also is the primary storm
conveyance system, only materials with N-value = 0.012 need to be considered. Where
the French drain is “offline” or is a secondary conveyance, analyses should consider N
value ≥ 0.020. After ascertaining the hydraulic needs, use the CSLE program or the
figures and tables in Appendix M to select materials that meet the required DSL based
upon the corrosion data and pipe size. Then determine minimum and maximum cover
and use the CSLE program or Appendix C of the Drainage Manual to select pipe materials
that meet the structural requirements.

Chapter 8: Optional Pipe Material 8-46


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

8.6 SPECIFYING OPTIONAL PIPE MATERIALS IN THE CONTRACT


PLANS
Show the optional pipe materials for cross drains, storm drains, French drains, and gutter
drains in the project plans, as illustrated in the FDM. The Optional Pipe Tabulation Sheet
includes: the size; class of concrete; gage, corrugation, and type of metal; and type of
plastic pipe that may be applicable to particular pipes within the project.

Side drains are listed in a Summary of Side Drains table. The two formats within Basis of
Estimates, Chapter 8, have columns for five round and five “other” pipe sizes with
corresponding MES widths; one has additional columns for offset and flowline. You
shouldn’t modify the tables except to “hide” columns not used or to change pipe sizes as
needed. As noted in the side drain example in Section 8.5, you should place notes stating
allowable side drain options below the table. Include any particular exceptions in the
Design Notes column. Do not use the Construction Remarks column, since that is
reserved for the construction phase of the project. The two Summary of Side Drain tables
are shown in Figure 8.6-1. See the current version of Basis of Estimates for the most
current form of these tables.

Figure 8.6-1

French drains also may be listed in a Summary Table; however, that table form has only
the actual limits of pipe/French drain and does not have a column for the structures or the
non-perforated pipe without the gravel envelope that extends a minimum of four feet on
each side of the drainage structure. If you decide to use this Summary Table, then the
Summary of Drainage Structures (SDS) tabulation should include the drainage structure
and non-perforated pipe and a separate column for the French drain segment. You can
note the pipe options allowable for French drains below the French Drain Summary Table
or within the Design Notes Column. You cannot use dissimilar types of pipe within a

Chapter 8: Optional Pipe Material 8-47


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

continuous run of pipe, and the non-perforated pipe should be of the same material as
the perforated pipe.

All other pipes that are listed in the SDS tabulation should have options listed in the
Optional Materials table. When elliptical pipe or arch pipe are the only allowed options,
these are listed in the Summary of Drainage Structures under the column heading “Other”
with the round equivalent size. The Optional Materials tabulation for these pipes should
include elliptical or arch pipe configurations that meet the DSL and structural
requirements. There are two Optional Materials tabulation formats; one includes flowlines
and one does not. Generally, flowlines for all options will be the same. However, in some
cases, minimum cover will control storm drain flowlines and it will be necessary to list the
required alternate flow lines. If round pipes meet the required clearances, there is no need
to even list elliptical and arch pipes as options since they usually are more expensive than
their round equivalents. For instance, where a round concrete pipe and arch metal pipe
meet all requirements, it is not necessary to list elliptical concrete as an option.

You can group pipe options by size and, if necessary, by location (station to station) or
by structure numbers. The structure numbers are listed as “Exceptions” in the “STR No.”
column next to the corresponding pipe size column. If the exceptions all have the same
limited options, the options can be listed with that group; otherwise, show the exceptions
individually with allowable material. Ideally, you can group the options by pipe size and
you can use the suitable materials for a spectrum of sizes. The intent of allowing options
is for the contractor to choose acceptable materials from a fair, competitive pipe supply
market, not to have numerous materials installed within a particular storm drain system.
In general, if you group material options by pipe size, one Optional Pipe Tabulation Sheet
is sufficient to describe allowable options for most projects.

Figure 8.6-2 is a spreadsheet format of the Optional Pipe Tabulation sheet containing the
optional materials determined for the examples in Section 8.5.

Chapter 8: Optional Pipe Material 8-48


January 2019

Drainage Design Guide


Chapter 8: Optional Pipe Material

Figure 8.6-2

Note that both of the Optional Materials Tabulation forms have a “PLOTTED” column. It
is important to check the material used for determining clearances at drainage structures.
If spiral rib pipe was assumed/used to determine clearances at structures and concrete
is listed as a pipe option, then the thicker wall of concrete pipe may not fit into the
structure. Structure fit may be another rationale for choice of pipe material. For example,
pipes with thinner walls would allow for smaller precast openings, which in turn allows for
smaller angle between pipes entering a round structure.

For design/build projects, you still need to create materials analyses to demonstrate
suitability of the pipe to be installed, and then you can include the analyses in project
documentation. You can include either an optional materials tabulation sheet in the
construction plans or make sure that the pipe material to be installed is noted somewhere
in the plans, such as on the plan sheets or on the Summary of Drainage Structures.

Chapter 8: Optional Pipe Material 8-49


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

CHAPTER 9: STORMWATER MANAGEMENT FACILITY

9. STORMWATER MANAGEMENT FACILITY .........................................................9-1

9.1 Selecting a Pond Site ........................................................................................9-1


9.1.1 Estimating Right-of-Way Requirements ...........................................................9-3
9.1.1.1 Typical Factors Controlling Surface Area Requirements...........................9-3
9.1.2 Access and Conveyance..................................................................................9-7
9.1.3 Joint-Use (Regional) Facilities..........................................................................9-8
9.1.4 Facilities on Forest Lands ................................................................................9-8
9.1.5 Coordination with Property Owners..................................................................9-9
9.1.6 A Suggested Evaluation Process ...................................................................9-10
9.1.6.1 Start Final Design ....................................................................................9-16

9.2 Maintenance, Construction, Aesthetics, and Other Concerns....................9-20


9.2.1 Maintenance...................................................................................................9-20
9.2.1.1 Pond Configurations ................................................................................9-20
9.2.1.2 Diversion Structures ................................................................................9-21
9.2.1.3 Conveyance to and from the Pond ..........................................................9-22
9.2.1.4 Vehicle Access ........................................................................................9-23
9.2.1.5 NPDES Permits .......................................................................................9-23
9.2.2 Construction ...................................................................................................9-23
9.2.2.1 Structure Tolerances ...............................................................................9-24
9.2.2.2 Earthwork Tolerances..............................................................................9-24
9.2.3 Aesthetics.......................................................................................................9-25
9.2.3.1 Fence.......................................................................................................9-26
9.2.3.2 Debris Collection......................................................................................9-27
9.2.4 Aviation Safety Requirements ........................................................................9-27

9.3 Stormwater Quality..........................................................................................9-29


9.3.1 Design Criteria................................................................................................9-29
9.3.1.1 Treatment Volumes .................................................................................9-29
9.3.1.2 Special Conditions ...................................................................................9-29
9.3.2 Concerns of Off-Line Systems .......................................................................9-30
9.3.3 Seasonal High Water Table ...........................................................................9-31
9.3.4 Treatment Methods ........................................................................................9-31
9.3.4.1 Wet Detention Systems ...........................................................................9-32
9.3.4.2 Retention Systems...................................................................................9-35
9.3.4.3 Filtration/Underdrain Systems .................................................................9-35
9.3.4.4 Stormwater Re-Use Systems ..................................................................9-41

Chapter 9: Stormwater Management Facility 9-i


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

9.3.4.5 Regional Stormwater Pond Systems .......................................................9-41


9.3.4.6 Harper (2007) Methodology for Nutrient Loadings Computation .............9-41
9.3.4.7 Protection of Springsheds from Nitrates ..................................................9-43

9.4 Stormwater Quantity Control .........................................................................9-50


9.4.1 The Department’s Design Storms ..................................................................9-50
9.4.1.1 Critical Duration .......................................................................................9-51
9.4.1.2 Storm Frequencies ..................................................................................9-54
9.4.2 Estimating Attenuation Volume ......................................................................9-54
9.4.2.1 Pre Versus Post Runoff Volume ..............................................................9-55
9.4.2.2 Simple Pond Model Procedure ................................................................9-58
9.4.2.3 Other Techniques ....................................................................................9-60
9.4.3 Tailwater Conditions.......................................................................................9-60
9.4.4 Routing Calculations ......................................................................................9-61
9.4.5 Discharges to Watersheds with Positive Outlet (Open Basins) Using Chapter
14-86..............................................................................................................9-62
9.4.6 Discharges to Watersheds without Positive Outlet (Closed Basins) Using
Chapter 14-86 ................................................................................................9-71
9.4.6.1 Retention System Groundwater Flow Analysis........................................9-72
9.4.7 Off-Site Inflows ...............................................................................................9-89
9.4.8 Commingling of Untreated Onsite Runoff.......................................................9-90

9.5 Outlet Control Structures ...............................................................................9-90


9.5.1 Weirs ..............................................................................................................9-90
9.5.2 Discharge Coefficients ...................................................................................9-91
9.5.2.1 Submerged Control Devices....................................................................9-93
9.5.3 Skimmers .......................................................................................................9-94
9.5.4 Miscellaneous.................................................................................................9-94

Chapter 9: Stormwater Management Facility 9-ii


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

9. STORMWATER MANAGEMENT FACILITY


9.1 SELECTING A POND SITE
Selecting the most appropriate pond site for a stormwater management facility requires
the work of many different offices and professionals within the Department. You, as a
drainage designer, will provide critical information, but because of the many factors to
consider, a team approach is essential.

There are numerous design features (depth, size, shape, treatment method, landscaping,
etc.) that you can modify to accommodate a pond site. However, hydraulic constraints
may preclude the use of some sites. Alternate sites and their different design features
usually will result in different costs and impacts. As a result, an evaluation of alternates
must be made to select the most appropriate pond site. The purpose of the evaluation is
twofold: (1) it will show that alternate sites were considered and that the selected site was
the most appropriate, and (2) when you combine the evaluation with the final design
details, they become the documentation that justifies the need to acquire property rights.

In the case where one person owns all the property in the area and that person is
agreeable to any pond location proposed by the Department, evaluating alternates may
not seem necessary. In these situations, the evaluation will not be as extensive as in other
situations; nevertheless, you should perform some amount of evaluation to show that the
site selected results in the lowest total cost.

The evaluation should weigh and balance numerous factors, such as cost; maintainability;
constructability; public opinion; aesthetics; and environmental, social, and cultural
impacts. The costs associated with right of way, environmental impacts, construction, and
long-term maintenance usually are the easiest factors to estimate and compare. Other
factors are more subjective and qualitative.

Because the evaluation involves a broad range of subjects, you should put together a
multi-functional team to select the most appropriate pond site. Teams should have
representatives from right of way, design, drainage, landscape architecture,
environmental management, maintenance, construction, and eminent domain. At times,
other units may provide critical information to the evaluation process. Although all of the
team members may not participate in the entire process, they will likely provide critical
information at some stage. The project manager, with support from the Drainage and
Right-of-Way offices, will be responsible for coordinating the team effort and ensuring that
the appropriate personnel participate.

Chapter 9: Stormwater Management Facility 9-1


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Consider the value of existing vegetation during site selection and pond siting. In some
cases, the need to preserve existing vegetation for aesthetic purposes may justify
additional project expenses (retaining walls, acquisition of additional right of way, etc.).

Perform pond site evaluations during the Project Development Phase. Often, you will re-
evaluate pond sites during the Design Phase. Before doing a design reevaluation, check
what commitments have been made and what work has been done during the Project
Development Phase.

CONSIDERATIONS WHEN SELECTING A POND SITE

1. Use existing FDOT properties or other 6. Avoid wetlands.


state-owned property, if feasible.
7. Avoid archaeological sites and historic
2. Minimize the number of parcels structures listed on or eligible for listing
required. For example, avoid using on the National Register of Historic
parts of two parcels when the pond will Places.
fit within one parcel.
8. Consider a joint-use facility (on the
3. Generally, property owners prefer to Department and another entity share) as
place ponds toward the rear of their an alternate, if one is feasible.
property. For parcels that abut the
roadway right-of-way, the portion of the 9. Generally, do not consider an option that
parcel next to the road usually is the requires water quality monitoring.
most expensive. Historically, this has been very
expensive.
4. Avoid splitting a parcel, thus creating
two independent parcel remainders. 10. Stormwater treatment systems must be
at least 30 meters (100 feet) from any
5. Consider the parcels identified by the public water supply well. (Chapter 62-
right-of-way office. Even if a parcel is 555, F.A.C.).
not large enough to provide all the
stormwater management, it may be 11. Locations with billboards usually are
large enough to provide the treatment expensive.
for stormwater quality. Or it could
12. Locations with mature, attractive trees
replace treatment and attenuation for
that will fit into the pond design increase
parcels adjacent to the road that will
the aesthetic value of the pond site.
have their ponds removed because of
the road improvements.

Chapter 9: Stormwater Management Facility 9-2


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

9.1.1 Estimating Right-of-Way Requirements


The right of way required for a pond site varies with the amount of additional impervious
area and associated additional runoff, the ground line and groundwater elevations at the
pond, the proposed road elevations, the existing on-site natural features, and sometimes
the soil conditions and other factors. During the pond site evaluation stage, the accuracy
to which you estimate these items and the resulting pond size varies with several factors.

Sometimes the acquisition schedule dictates that results of the pond site evaluation form
the basis for the final pond site right-of-way requirements. For these projects, you should
determine the pond size as accurately as if doing the final detailed design.

There are other projects where the determination of the final right-of-way requirements
occurs shortly after the pond site evaluation. After establishing the final right-of-way
requirements, the acquisition process starts. For these projects, you would perform a
pond site evaluation only to compare alternate sites or drainage schemes. Make your size
estimates accurate enough to minimize changes to the right-of-way requirements during
the final design.

In a third category of projects the right-of-way acquisition is scheduled for several years
after the pond site evaluation, or the acquisition is not even funded in the Department’s
work program. For these projects, changes in pond size and location from that established
in the original evaluation will not affect production schedules or the right-of-way
acquisition process substantially. Therefore, your pond size estimates need not be very
accurate. For these projects, you typically would perform a pond site re-evaluation shortly
before right-of-way acquisition.

Other factors that affect the level of accuracy for pond size estimates are property costs
and the existing and anticipated development of the project area. In a rural area with
relatively large tracts of land, changes to pond size and location will have less impact to
property owners and the Department than in an expensive urban area that is rapidly
developing and has relatively small parcels. As a result, the pond size estimates you use
for these evaluations do not need to be as accurate as in urban, rapidly developing areas.

9.1.1.1 Typical Factors Controlling Surface Area Requirements


Typically, the need to fit storage volumes within upper and lower constraints dictates the
amount of surface area required for a pond. The following items could control the surface
area requirements for a pond:

 The ground line at the pond (or the berm elevation) minus the freeboard dictates
the top of the treatment and attenuation volume.

Chapter 9: Stormwater Management Facility 9-3


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

 For urban projects, the low point in the gutter minus the hydraulic gradient
clearance dictates the hydraulic gradient of the storm drain. This constraint often
is critical in flat terrain but not steep terrain.

 High groundwater elevations or sometimes discharge tailwater elevations can


constrain storage volumes. The groundwater elevation constraint will vary with the
method of treatment used and the requirements of the regulatory agency.

 Retention ponds must recover a certain volume in a certain time. The size of the
pond bottom area sometimes controls the recovery or drawdown time. This may
be particularly critical for ponds discharging to closed basins.

 For wet detention facilities, most regulatory agencies limit the treatment volume
depth to 18 inches and you must provide the required permanent pool volume.

 To contain a substantial portion of the pond volume in rolling or steep terrain, you
would berm the low side of the pond site. The horizontal distance of the
embankment from the berm top to natural ground dictates how much right of way
you will need in this direction. The embankment slope must be flat enough to be
stable. For example, a 1 (vertical) to 2 (horizontal) slope in sandy soil with seepage
may not be stable. In this case, it would be appropriate for you to conduct a slope
stability analysis. Discuss these situations with the geotechnical engineer to
establish an acceptable slope and thus a reasonable estimate of the surface area
requirements.

 You might adjust the shape of the pond—and, therefore, the surface area—due to
existing on-site natural features (mature vegetation, a significant stand of
vegetation on a slope, a visual landscape barrier, etc).

Example 9.1-1. Estimating Pond Right-of-Way Requirements

Given:

 Flat terrain, approximately 1-percent slope


 Proposed pond discharges to open basin
 Proposed curb and gutter section with gutter elevation at the low point in profile =
59.9 ft
 Ground elevation at pond site = approx. 59 ft
 Estimated seasonal high water table (SHWT) = 2.5 ft – 3.6 ft below ground
(based on NRCS soil survey)
 Treatment volume = 10,950 ft3
 Estimated peak attenuation volume = 19,567 ft3 (from Example 9.4-1)
 Estimated 3-year attenuation volume = 10,243 ft3 (storm drain design frequency)

Chapter 9: Stormwater Management Facility 9-4


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Find: Estimated surface area requirements for a pond

1. Since the SHWT is so close to the surface, you choose a wet detention pond.

For these conditions, one of two requirements typically control the surface area. Both
involve spreading the treatment and attenuation volumes over a large enough area
to keep the height of the volume within limits. The height (H) of the treatment and
peak attenuation volume is constrained on the top, by the ground elevation minus the
freeboard, and on the bottom, by the controlling groundwater elevation. Although
some Water Management Districts (WMDs) allow treatment below the SHWT, this
example will assume that treatment is above the SHWT. First, determine the surface
area necessary to meet these constraints. The other requirement that may control
the surface area is discussed after Step 5.

2. Conservatively, assume the SHWT is 2.5 feet below ground. The standard
freeboard is given in Section 5.4.4.2 of the Drainage Manual. The treatment and
peak attenuation volume are constrained to the following height (H).

H = Depth to SHWT – Freeboard


H = 2.5 – 1.0
H = 1.5 ft.
3. The total peak storage volume required is:

VolumePEAK = Treatment Volume + Est. Peak Attenuation Volume


VolumePEAK = 10,950 + 19,957 = 30,907 ft3

You will need to make assumptions about the pond configuration.

Shape: Assume the shape of the pond will be rectangular. Irregular shapes usually
can be approximated by a rectangular shape so this is a reasonable assumption and
it greatly simplifies estimating the surface area.

Length to Width Ratio (L/W): The property lines may suggest a preferred ratio to
make best use of a parcel. Without other guidance, assume L/W = 2.

Side Slopes: Assume flat slopes, such as 1 (vertical) to 5 or 6 (horizontal) for sites
required to be aesthetically pleasing. Assume 1 (vertical) to 4 (horizontal) for most
other conditions.

Chapter 9: Stormwater Management Facility 9-5


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

4. Use the formula for a rectangular box to determine the water surface area of a
pond with vertical sides.

Volume = LRECT WRECT H

where:
H= Height (m) = 1.5 ft for the above condition
LRECT = Length (ft) of vertical-sided pond
WRECT = width (ft) of vertical-sided pond
Assume for this example that L/W = 2, then

30,907 ft3 = LRECT x (0.5 LRECT) x 1.5 ft, then


LRECT = 203 ft
WRECT = 101.5 ft

5. Increase these dimensions to account for sloped sides by adding: 2 x (0.5 x H x


Side Slope).

For this example, assume side slope = 5, thus adding 7.5 ft to each dimension.

Length @ top of slope = 210.5 ft


Width @ top of slope = 109 ft

Then,
Water Surface at Peak Design Stage = 210.5 x 109 = 22,944.5 ft2 = 0.53 ac
The other requirement that may control the surface area in flat terrain is the
requirement to maintain the clearance between the low point in the gutter and the
hydraulic gradient in the storm drain system. On the top, the low point in the gutter
minus both the hydraulic gradient clearance and the energy losses in the storm drain
system constrain the treatment volume and three-year attenuation volume. On the
bottom, the groundwater elevations (SHWT for this example) constrains these
volumes. The standard hydraulic gradient clearance is given in the Drainage Manual.

You can estimate the energy losses in the storm drain system in two ways. Assume
a hydraulic gradient slope. Slopes of 0.05 percent to 0.1 percent are common in flat
terrain. Multiply the length between the pond and the low point by the assumed slope
to obtain the losses. Another approach is to assume a fixed energy loss, ignoring the
length between pond and low point. In flat terrain, a reasonable value for this
purpose is two feet.

Chapter 9: Stormwater Management Facility 9-6


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

6. The SHWT elevation is 56.5 feet (59 – 2.5). For this example, you can assume
the energy loss in the storm drain to be 0.7 ft. Then, the treatment and three-year
attenuation volume are constrained to the following height (H):

H = Low Point in Gutter – Clearance – Estimated Energy Losses – SHWT


Elevation
H = 59.9 – 1.0 – 0.7 – 56.5
H = 1.7 ft

This is greater than the height (1.5 feet) available to “stack” the peak attenuation
volume (Step 2). Since the three-year attenuation volume is less than the peak
attenuation volume, this constraint will not control the water surface area. If the
height was less than determined in Step 2, you would estimate the water surface
area as done in Step 4 except using different values for H and the total volume.

The water surface area dimensions determined in Step 4 apply.


7. Add the maintenance berms to the water surface dimensions. The standard
maintenance berm width is given in Section 5.4.4.2 of the Drainage Manual.

Length = LTOP + 2 (berm width) = LTOP + 2(20) = 210.5 + 40 = 250.5 ft


Width = WTOP + 2 (berm width) = WTOP + 2(20) = 109 + 40 = 149 ft
Area = 250.5 x 149 = 37,324.5 ft2 = 0.86 acre
8. Increase the value by 10 percent to 20 percent to account for the preceding
information being preliminary. For this example, we will increase it by 10 percent.

Area = 0.86 x 1.1 = 0.95 ac

Realize that this is only the pond size estimate. You also must make estimates for
access and conveyance, as discussed in the next section.

9.1.2 Access and Conveyance


The right of way required to convey the project’s runoff to and from a pond and to provide
access can affect which alternate pond site is the most appropriate. Determine these
requirements for each alternate and include the costs and impacts in the evaluation.

Sites placed far from the project will require more right of way to get stormwater to the
pond than sites adjacent to the project. Similarly, different pond sites can have different
right-of-way requirements for the outfall (discharge) from the pond. Guidelines for
establishing the width or “footprint” of the right-of-way requirements for conveyance are
provided in Section 9.2 of this document.

Chapter 9: Stormwater Management Facility 9-7


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

The Department often provides access through the same property obtained for conveying
the project’s runoff. For pond sites placed far from the project, providing access from a
local road closer to the pond is sometimes more reasonable.

Usually, the Department will obtain the right of way required for access and conveyance
as a perpetual easement. Fee simple right of way may be appropriate sometimes. The
opinion of the District Maintenance Office, balanced with property owner preference and
right-of-way costs, is the primary factor for determining which type of right of way is
appropriate.

Refer to Appendix B of the Drainage Manual and the FDM113. Both contain additional
information about acquisition of property rights.

9.1.3 Joint-Use (Regional) Facilities


Sometimes the Department and other entities can share a stormwater management
facility. Both the Department and the other entities receive the stormwater management
benefits of the facility and share in its construction, operation, or both. The Department
and the other entities create a written agreement describing the responsibilities of each
party. Typically, these agreements are made with local governments, but sometimes
private entities enter joint-use agreements. For example, the Department shares several
facilities with golf course owners.

Advantages of a joint-use or regional facility are that: (1) the Department often can relieve
itself of the maintenance requirements, (2) water quality improves downstream, and (3)
stormwater re-use is incentivized when a larger volume of water is available. A joint-use
facility can have disadvantages, such as affecting production schedules, a more complex
permitting process, and resolving any non-complying discharges, if they occur.

When developing a joint-use agreement, avoid commitments that hold the Department to
completing construction of the site by a certain date because there often are unforeseen
delays in permitting and funding. Developing an acceptable joint-use agreement often
requires an extensive coordination effort involving the project manager and
representatives from numerous other offices. Discuss this option with the project manager
or District Drainage Engineer.

9.1.4 Facilities on Forest Lands


Occasionally, projects are bounded by state and/or national forest lands and ponds must
be located within these public preserves. In such cases, advanced coordination with the
owning agency and the WMD can result in cost-effective designs that will not degrade the
public purpose of the forest lands. This cooperative process can sometimes take longer
to complete and should, therefore, be started early in the PD&E phase.

Chapter 9: Stormwater Management Facility 9-8


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

9.1.5 Coordination with Property Owners


Often, contacting the property owner to get his or her preference regarding the shape and
location of the pond and location of the access road is beneficial from a right of way
standpoint. This coordination is especially important where the Department needs only
part of a parcel for a pond.

Consider contacting the owner during the evaluation of alternate sites. A situation where
contacting the owner during the evaluation may be appropriate is where one person owns
all the property in the area. If a contact is not made during the evaluation process, it is
recommended that a contact be made shortly afterward and before starting final design.
For example, perhaps the property owner may prefer a shallower pond although it would
require more right of way, or the owner may be interested in re-acquiring and maintaining
the pond. A certain pond shape could give the owner better use of the remainder of the
parcel. This is important information to know before starting final design. In some
instances, contacting homeowner’s associations or abutting property owners may be
beneficial to find out if a negative perception of the proposed pond exists.

Sometimes, contacting the owner may not be appropriate. Where the Department needs
an entire parcel, there is no need to obtain the owner’s preference about pond location.

The project manager, with participation from the right-of-way office, should decide
whether to contact the property owner based on individual circumstances.

The Department’s project manager or a right-of-way specialist or both could make the
contact. As a drainage designer, you are the best source to answer technical questions
and will likely be asked to be present when the contact is made. You cannot provide
specifics early in the design process, but you can speak about general principles of
stormwater management facilities.

When obtained in writing, you should accommodate the property owner’s preference to
the greatest degree possible. The Department may not be able to accommodate all of the
owner’s preferences in the design of the pond due to hydraulic constraints or other
limitations. However, after weighing and balancing the owner’s requests with other
factors, it is likely that some aspects of the owner’s preference can be satisfied, thus
improving relations during the right-of-way acquisition process.

If a commitment is made to a property owner, follow through or notify the owner that the
Department cannot meet the commitment. Usually, you will not have enough information
to commit to anything during the first contact with the owner. Remember that the purpose
of the initial contact is to learn the owner’s preference regarding the shape and location
of the pond and location of the access road. The most that you can commit to is to try to

Chapter 9: Stormwater Management Facility 9-9


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

accommodate the owner’s requests. If, during any discussion, the property owner is told
about the operation, shape, or location of the pond, this is a commitment. If you
subsequently design the pond differently, you should notify the property owner. If the
owner is not notified, the right-of-way specialist is placed in the difficult situation of
approaching the owner with a proposed pond configuration that is different than what was
discussed previously.

This holds true for changes that occur through the detailed design phase. The owner must
be notified if the shape, size, and location of the pond are going to be different than what
was discussed previously.

9.1.6 A Suggested Evaluation Process


An outline for evaluating alternate sites follows, and a flow chart is provided in Figure 9.1-
5. The process is divided into seven main steps of work, as follows:

Step 1 Coordinate with the right-of-way office

Step 2 Identify alternate drainage schemes

Step 3 Estimate the right of way required for each alternate

Step 4 Get team buy-in on the proposed alternates

Step 5 Estimate costs and assess impacts

Step 6 Summarize findings

Step 7 Select site

The steps listed below are directed toward you, the drainage designer, but there also is
information about activities that team members from other offices should perform.
Normally, you should proceed through the steps in order, but, often, doing certain steps
earlier in the process or doing several steps concurrently will be reasonable and prudent.
The most important issue is to maintain the coordination necessary to ensure that pond
sites are selected using a multi-functional team.

The degree of detail will vary with individual projects and between FDOT districts. It is
essential that you discuss this with the project manager or the District Drainage Engineer
before starting the evaluation.

Chapter 9: Stormwater Management Facility 9-10


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Step One Coordinate with the Right-of-Way Office

The purpose of this coordination is to provide a preliminary pond size and a general
location to the right-of-way office and to ask the right-of-way office to identify potential
sites.

Shortly after the roadway typical section is set, provide the right-of-way office with a
preliminary estimate of the size and a general location of the pond. Use aerial contour
maps, old construction plans, available surveys, and other data to identify the primary
basins and the general outfall locations (discharge points). Identifying the high points
along the project usually separates the primary basins. At this stage, assume that the
pond site will be near the lows in the terrain and will be close to the existing outfalls. As a
preliminary size estimate, use 20 percent of the roadway right of way draining to the
outfall. The area identified for the general location should be large enough to allow for
several alternates to be developed. Refer to Figure 9.1-1. The project manager should
relay this information to the right-of-way office so it can include the preliminary costs for
pond sites in the cost estimates.

Figure 9.1-1: Size and Location for Initial Coordination with the Right-of-Way
Office

When the corridor and alignment (left, right, or center) are set, the project manager should
request the right-of-way office to identify parcels along the roadway that could be
economical for a pond, due to the impacts of the roadway footprint. The right-of-way office
also should identify existing excess property in the area.

At this stage, impacts of the roadway footprint at intersections and interchanges may be
uncertain still simply because the geometry has not been set. These areas may warrant
discussions with the right-of-way office at a later time.

Chapter 9: Stormwater Management Facility 9-11


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

When the right-of-way office completes this task, the project manager should arrange a
meeting with the team to discuss all potential pond sites, aesthetic concerns, and possible
contacts with property owners. Representatives from right of way, drainage, landscape,
and environmental management should attend.

Refer to tax maps while discussing potential pond sites. The project manager should have
these; if not, the local government should.

Step Two Identify alternate drainage schemes

Before developing the alternates, familiarize yourself with soils and groundwater
conditions in the area and with the various stormwater quality treatment methods. Use
the Natural Resources Conservation Service (NRCS) (formerly Soil Conservation
Service) soil surveys to obtain the soil information. The treatment methods are discussed
in Section 9.3, below.

It may be reasonable to start this step by qualitatively eliminating areas that are not
hydraulically feasible. For example, some areas may be too high in elevation, or may be
at the beginning of the drainage system rather than at the end.

For projects in developing areas, consider contacting the Planning (or Development)
Department of the local government to find out the zoning for the area, the planned land
use, and if proposed developments exist. Although this information should not
automatically eliminate a site from being evaluated, it may help you to identify viable
alternatives, such as a joint pond use with future land developers.

Identify two or three alternate drainage schemes for each primary basin. If two or three
vacant sites are not available, then consider developed sites. Familiarize yourself with the
list of considerations in Section 9.1 when identifying your drainage schemes. Also
consider the sites identified by the right-of-way office in Step One. This is not to say that
these sites need to be evaluated as alternates, but all of the alternates evaluated must
be viable. You should consider these sites during the evaluation.

The alternates may be as simple as two different locations for a wet detention pond, or a
wet detention pond compared with a dry pond with underdrain at the same location. A
system using two ponds, one for off-line quality treatment and one for attenuation, could
be compared with a single pond designed for both quality treatment and attenuation. In
areas with expensive right of way, identifying an alternate that uses a non-standard
approach—such as sand box filters or pumping stations—may be prudent. Check with
the District Drainage Engineer before doing so. See Figure 9.1-2.

Chapter 9: Stormwater Management Facility 9-12


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Figure 9.1-2: Alternate Drainage Schemes

Step Three Estimate the right of way required for each alternate

A. Consider the need for additional soils and groundwater information. Most of the
Department’s districts accept the NRCS soil surveys for pond site evaluations. For
alternates using retention or exfiltration in areas where there are poor soils and for
projects discharging to a closed basin, site-specific data may be appropriate. If you
feel that additional information is warranted, discuss this with the District Drainage
Engineer.
Steps B through G apply to ponds discharging to open basins. Ponds
discharging to closed basins have the additional complication of assuring that
the drawdown requirements are met (see Section 9.4).

B. Determine the required treatment (quality) volume. See the discussion of treatment
volumes in Section 9.3. Refer to the appropriate regulatory agency’s rules or meet
with the agency at this time.
C. Estimate the required attenuation volume. See the discussion of Estimating
Attenuation Volume in Section 9.4.2.
D. Coordinate with the Landscape Architect to perform a preliminary identification of
existing landscape, natural and aesthetic features, and opportunities and
constraints that could impact the placement and design of the pond.
E. Estimate the low point in the proposed roadway. Discuss the grade with the
roadway designer as necessary.

Chapter 9: Stormwater Management Facility 9-13


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

F. Obtain ground elevations around each alternate site. Using a contour map with
one-foot intervals usually is sufficient. In flat terrain where one-foot contour maps
are not available, obtaining a survey of the ground elevations around each
alternate site may be appropriate.
G. Determine the pond surface area necessary to satisfy all applicable criteria. Refer
to the typical controlling factors in Section 9.1.1.1. If you know of aesthetic
preferences that will affect the surface area, such as shape, side slopes,
landscaping, or preserving existing vegetation, account for them in the surface
area determination. Example 9.1-1 goes through this and the following two steps.
H. Add the maintenance berms to the above area.
I. Increase this area by 10 percent to 20 percent to account for the preceding
information being preliminary.
J. Place these surface area requirements within parcel boundaries in a way that
minimizes the number of parcels required. For example, avoid using part of two
parcels when the pond will fit within one.
K. Determine the right-of-way requirements for access to the pond and for
conveyance to and from the pond.
L. Sketch each alternate site and its requirements for conveyance and access on the
tax maps (preferably on aerial background). Refer to Figure 9.1-3.

Figure 9.1-3: Sketch of Each Alternate’s Estimate Requirements

Chapter 9: Stormwater Management Facility 9-14


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Check with the project manager or District Drainage Engineer to see if they want to review
the above work before proceeding to the next step.

Step Four Get team buy-in on the proposed alternates

The project manager should arrange a meeting with the team to discuss the alternates.
The meeting has several purposes: (1) discuss how the right-of-way requirements fit
within parcel boundaries, (2) confirm that alternates being considered are viable, (3)
consider the need to contact property owners to obtain their preference of pond shape
and location, (4) confirm that the access and conveyance requirements are reasonable,
and (5) discuss social, cultural, and environmental impacts, including the existing
landscape, natural and aesthetic features, and opportunities and constraints of each
alternate.

If the property owners are contacted, their preferences should be discussed among the
appropriate team members, and the sites appropriately adjusted before proceeding to the
next step.

Step Five Estimate costs and assess impacts

When the team agrees on the alternate drainage schemes, the project manager should
request environmental assessments, right-of-way cost estimates, and utility impact
assessments for each alternate site. The purpose of the environmental assessments is
to determine potential hazardous material contamination and potential impacts to
environmental resources such as threatened, endangered or significant species and
cultural resources. Environmental specialists from the Environmental Management Office
usually do the assessments, which should include cost estimates associated with any
mitigation and environmental cleanup.

The purpose of the utility assessment is to determine the existence of utility corridors
through each alternate site.

You, as the drainage designer, should estimate the construction cost of each alternate,
including the conveyance requirements to and from the pond. Usually, the largest costs
are associated with earthwork, pond liner (when required), and pipe. Statewide average
unit prices for the standard pay items are provided in the publication Item Average Report,
which is available for download at:
https://fdotwp1.dot.state.fl.us/wTWebgateReports/login.aspx (note: the user must have a
login and password for a specific project). For alternates that are similar, estimating
construction cost differences rather than total construction costs may be reasonable. If
different alternates are expected to have substantially different maintenance costs,
estimate these as well. Since maintenance costs will be spread over time, it will be

Chapter 9: Stormwater Management Facility 9-15


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

necessary to equate these to initial costs using a life cycle analysis. Contact the District
Maintenance Office to obtain the latest unit prices for routine maintenance activities.
Each alternate should have, at a minimum, cost estimates for right of way. When the
estimates and assessments are complete, the various offices should furnish their findings
to you via the project manager.

Step Six Summarize findings

For each basin, combine the findings of the other offices with your construction cost
estimates. Use a summary table similar to Figure 9.1-4 to compare the alternates. The
Drainage Manual lists the minimum documentation requirements.

Check with the project manager to see if the district staff wants to review the summary
before proceeding to the next step.

Step Seven Select site

The team should meet to discuss all alternates and select the most appropriate site. Cost,
maintainability, constructability, public opinion, aesthetics, and environmental (social,
cultural, natural, and physical) impacts will affect the selection of a pond site. The team
should weigh and balance all factors in their decision. Include documentation of the
decision with the summarized findings of the previous step.

9.1.6.1 Start Final Design


For most projects, the actual right-of-way requirements will be determined during the final
design of the pond. The acquisition of the pond site occurs during the process of acquiring
any additional right of way for the roadway corridor. You should revisit the site evaluation
process if the final right-of-way requirements are substantially different from those
originally estimated. Pond locations frequently change as the final design progresses.
Sometimes additional sites are evaluated, and occasionally the originally selected site is
not used. Any additional evaluations of pond sites should be documented as required by
the District Drainage Engineer. All changes in right-of-way requirements must be
coordinated with the right-of-way office.

Chapter 9: Stormwater Management Facility 9-16


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Figure 9.1-4

Chapter 9: Stormwater Management Facility 9-17


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Figure 9.1-5

Chapter 9: Stormwater Management Facility 9-18


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Figure 9.1-5 (continued)

Chapter 9: Stormwater Management Facility 9-19


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

9.2 MAINTENANCE, CONSTRUCTION, AESTHETICS, AND OTHER


CONCERNS
9.2.1 Maintenance
Maintenance must be a consideration throughout the process of designing a stormwater
facility. Long-term maintenance costs are inevitable, but they can be minimized by
appropriate consideration during the design of a facility. The difference between a
maintainable design and a design that is difficult and expensive to maintain often will be
the difference between an attractive operating facility and a neglected, non-functioning
facility generating frequent public complaints.

9.2.1.1 Pond Configurations


Side slopes:

Use a slope of 1 (vertical) to 4 (horizontal) or flatter. Steep slopes are harder to mow and
are more susceptible to erosion than flat slopes. Slopes steeper than 1:3 must be mowed
with special equipment. This is generally more expensive than using regular mowers.

Where possible, conserve established slope vegetation to increase stability and add an
aesthetic feature to the pond.

Maintenance berms:

The Drainage Manual gives the minimum widths and slopes. These are acceptable for most
situations, but discuss site-specific concerns with the local maintenance staff.

For ponds that will maintain a permanent or normal pool, keep the lowest point of the
maintenance berm at least one foot above the top of the treatment volume. This will
minimize saturation of the maintenance berm.

Corners:

Use a radius of 30 feet or larger for the inside edge of the maintenance berm. This is based
on the largest piece of normal maintenance equipment. Several maintenance vehicles
were modeled using the AUTOTURN program (Transoft Solution, Inc.). The GRADALL
880 requires the largest turning radius and gate opening.

Chapter 9: Stormwater Management Facility 9-20


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Benchmark:

Have a benchmark constructed in or near all ponds. It will be used to check critical
elevations of the pond and outlet control structure. Avoid installing benchmarks in areas
subject to settlement such as high fill sections and areas subject to vehicle loads. An
outside corner of the maintenance berm in a minimal fill section would be an appropriate
location.

Sediment buildup:

Design the pond with a three-foot deep sediment sump near the inlet. In retention ponds
(described in Section 9.3.4.2) where the groundwater is close to the pond bottom, the
depth of the sump may need to be reduced to avoid exposing the groundwater. The area
of the sump should be approximately 20 percent of the pond bottom area.

In retention ponds, the sediment is visible, but often it accumulates so slowly that it is
difficult to see how much exists. A staff gage placed near the inlet allows the buildup to
be measured.

Permanent (Normal) Pool Depth:

The main body (not the littoral shelf) of the permanent or normal pool should be deep
enough to minimize aquatic growth, but shallow enough to maintain an aerobic
environment throughout the water column. The regulatory agencies usually will specify
the maximum depth for water quality credit, but this depth may be exceeded for harvesting
fill needed for the project or to preclude future maintenance cleaning; in such cases, the
extra pond depth will not be credited toward the regulatory permanent pool requirement.
If the minimum depth is not specified, use five feet to minimize aquatic growth.

Side Bank Underdrain Filters:

Do not construct these around the entire pond. Design the pond to have at least 20 feet
of the side slope without underdrains so that maintenance vehicles can get to the pond
bottom without running over the underdrain.

9.2.1.2 Diversion Structures


Diversion structures of off-line systems must have a manhole for access on each side of
the weir (refer to Section 3.10 of the Drainage Manual). Furthermore, the manholes
should be located out of the roadway pavement to allow access without blocking traffic.
Off-line systems are discussed in Section 9.3.

Chapter 9: Stormwater Management Facility 9-21


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

9.2.1.3 Conveyance to and from the Pond


The right of way obtained for conveyance to and from the pond must be sufficient to
maintain the conveyance. This is true for either piped or open-ditch conveyance systems.
Figure 9.2-2 provides typical sections for establishing the width of the right-of-way
requirements.

Where the pond discharges to something other than an existing storm drain system,
obtain right of way from the pond to a receiving surface water body (lake, wetland, ditch,
canal, etc.) even if there are no physical changes proposed to the conveyance path. This
assures that the Department will have the right to maintain the flow path.

Figure 9.2-2: Required Right of Way Widths

Chapter 9: Stormwater Management Facility 9-22


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

9.2.1.4 Vehicle Access


Roads:

Sometimes, you can use the right of way available for conveyance to provide
maintenance access to the pond. For pond sites located far from the project, it may be
more reasonable to reach the pond from a local road. In flat terrain, an ideal width of right
of way for access only (not including conveyance) is 15 feet. Larger widths may be
necessary for turns. In irregular terrain, consider the distance to tie into natural ground.
Concentrated flows crossing the access road may require a culvert crossing. If the vertical
clearance is restricted, discuss it with maintenance personnel.

The roadway designer should design and incorporate curb cuts and driveways in the
plans where the access road joins the public road.

Gates:

If you plan to fence the pond, use a 24-foot or two 12-foot sliding cantilever gates
(Standard Plans, Index 550-003). This will allow the largest piece of normal maintenance
equipment to enter and exit without having to back out along the access road. If you must
use a swinging gate, pave the area under the arc of the gate swing. Show the gate type,
location, and size in the plans.

9.2.1.5 NPDES Permits


Active National Pollutant Discharge Elimination System (NPDES) permits may cover the
limits of proposed construction. The District NPDES Coordinator needs to review the
proposed project to ensure compliance with any active permits.

9.2.2 Construction
Consider the right of way needed to construct the facility. The right of way needed to
maintain the facility, i.e., the permanent right of way, may be, but is not always, sufficient
to construct the facility. If the construction area is outside the permanent right of way, you
should use temporary construction easement documents to obtain sufficient area for the
contractor to construct the facility.

Some water management districts require a professional land surveyor to lay out final
placement of drainage structures. Some of the Department’s districts are directing the
contractor to do this. Discuss this with the project manager or district construction
personnel. If they want to have the contractor survey the final placement, include the
requirement in the contract documents, as directed by the district.

Chapter 9: Stormwater Management Facility 9-23


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Often, the regulatory agencies place special requirements on the Department’s projects
as “conditions of the permit.” Requirements that will affect the contractor’s work must be
incorporated into the plans or specifications for bidding and payment purposes. It is not
sufficient that the permits will become part of the contract documents.

9.2.2.1 Structure Tolerances


Unless otherwise dictated, the tolerance for drainage structures is controlled by Section
5-3 of the Standard Specifications, which reads: “reasonably close conformity with the
lines, grades, . . . specified in the contract documents.” The tolerance is particularly
important for weirs, orifices, and other flow control openings of outlet control structures.
You can calculate weir dimensions quite precisely, but it is not reasonable to construct
concrete structures to that same precision. Complicating this in the past, the regulatory
agencies’ inspectors sometimes have expected the dimensions to be exactly as shown
in the plans.

During design, if you realize that the designed discharge is sensitive to small changes in
weir dimensions, you should conservatively account for the tolerance in the calculations.
For example, to maintain the discharge rate at or below the allowable rate, specify a weir
width that is 0.05 feet smaller than the width required to discharge at the allowable rate.
And include the tolerance mentioned above. If the contractor constructs the weir 0.05 feet
wider than specified, it will match the designed width. If the weir is constructed 0.05 feet
narrower than specified, the discharge rate still will be less than the 0.05 feet maximum
allowed. In the last condition, you should check that stage has not increased to a point
where the pond is now discharging through the overflow point.

Although not often used, another option is to use “bolt on weir plates” with slotted bolt
holes. The plate elevation then can be adjusted to exact elevations after the structure is
set.

9.2.2.2 Earthwork Tolerances


By Standard Specifications, the tolerance for earthwork within a stormwater management
facility is 0.3 feet above or below plan cross section (Section 120-12). For some retention
ponds, having a bottom 0.3 feet higher than anticipated may substantially reduce the
treatment volume and somewhat affect the attenuation capacity. Conversely, having a
bottom 0.3 feet lower than anticipated may substantially increase the retention (or treated)
volume and affect the recovery time. This tolerance will not affect wet detention facilities.

Do not specify a tolerance that may conflict with the Standard Specifications. If the
standard tolerance will substantially reduce the retention or treatment volume—as in a
shallow retention pond—design the pond to allow for the bottom being 0.3 feet higher or
lower than shown in the plans. In other words, specify a pond bottom that is 0.3 feet lower

Chapter 9: Stormwater Management Facility 9-24


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

than necessary to retain the minimum volume. For example, the pond bottom may need
to be 0.7 feet below the weir to provide the treatment volume. Specify the bottom to be
1.0 foot below the weir to allow for the earthwork tolerance. Determine the recovery time
assuming that the pond bottom is 1.25 feet below the weir, i.e., 0.3 feet below the
specified bottom elevation.

You should reserve this extra effort for facilities where the earthwork tolerance could
substantially reduce the retention or treatment volume.

9.2.3 Aesthetics
The Florida Department of Transportation has adopted a Highway Beautification Policy
to include aesthetic considerations in the design aspects of highways. Chapter 5 of the
Project Development & Environmental Manual summarizes the requirements and
provides direction in applying them to Department projects. Aesthetic considerations are
cited in Section 5.4.4.2 of the Drainage Manual as an integral part of sound pond design.
Often, programmatic or aesthetic commitments are made during the project development
phase. If so, the environmental document will contain a discussion of visual impacts and
aesthetic requirements for stormwater ponds. Discuss this with the Landscape Architect
and Environmental Management Office project manager.

The location, size, shape, side slopes, fencing, and landscaping all affect the aesthetic
quality of a pond. In general, irregular shapes, gradual slopes, and no fence are more
aesthetically pleasing and have less visual impact than rectangular shapes and steep
slopes with a chain link fence. For this reason, the Drainage Manual mandates that the
default pond design should not include fencing, and that fencing must be justified within
the design documentation. You can use irregular side slopes for permanently wet ponds
to create an undulating water edge even when the perimeter of the site is rectangular.
Preservation of existing vegetation and inclusion of native and wetland vegetation can
greatly improve the visual appearance of a pond. Typically, this will require that you
design and construct physical barriers to protect the existing vegetation from construction
equipment.

In urban areas, ponds designed with a park-like appearance will encourage the local
government to undertake the maintenance. If you design a pond site to be landscaped, a
memorandum of agreement (MOA) for maintenance may be executed with the local
government. In the absence of an MOA, the Department may undertake the landscape
maintenance of a pond. The District Landscape Architect is familiar with the MOA
procedure. Any landscape projects should be coordinated by the project manager with
support from the District Landscape Architect.

Chapter 9: Stormwater Management Facility 9-25


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

The shape, depth, and side slopes will affect how much right of way is required for a pond.
Therefore, you must evaluate and weigh aesthetics among the other factors during the
site selection process (see Section 9.1). The Department has determined that pond
aesthetics is an acceptable design objective that would justify acquisition of additional
right of way, including eminent domain acquisition, when appropriate. Seek out the District
Landscape Architect to coordinate and develop appropriate aesthetic features. Your
responsibility is to ensure that the design constraints (volumes, depths, littoral shelves)
are met while accommodating the aesthetic features. Coordinate with the District
Landscape Architect to establish the quantity of right of way needed to meet aesthetic
and design constraints.

9.2.3.1 Fence
The Drainage Manual mandates that the default pond design does not include fencing
and that use of fencing must be justified within the design documentation. Design
stormwater ponds to avoid the need for fence, if feasible. Typically, the flow velocities
within a stormwater pond are low and, therefore, the velocities do not create a hazard.
Unexpected deep standing water—such as an immediate 1:2 drop off at the water’s
edge—should be avoided or fenced. Under the Statewide Environmental Resource
Permit (ERP Ch. 62-330) Rule, the Drainage Manual, Florida Department of
Environmental Protection (FDEP), and all the water management districts allow for
unfenced facilities if the slopes are 1 (vertical) to 4 (horizontal) or flatter. Refer to Section
2.6.1 of the Drainage Manual for further discussion of protective treatment.
When it is necessary to provide a fence, one that fits the surrounding community is ideal.
The style (wood, block, chain link, wrought iron, etc.) will vary from community to
community. Pay item 0550-10 series covers special fencing; however, special details and
specifications will need to be included in the contract documents. Because of the extra
work, special fencing has not been commonly used. Another complication with special
fencing is that the Department’s maintenance units do not normally have the materials to
repair them; therefore, confer with the local maintenance engineer anytime you are
considering special fencing.
If it is not feasible to provide a special fence, the next option is to use standard FDOT
fence. In rural areas, the Type A fence, Standard Plans, Index 550-001 usually is
appropriate. In urban areas, Type B fence (chain link), Standard Plans, Index 550-002
usually is appropriate.
Fence Color:
One of the simplest things you can do to reduce the visual impact of chain link fence is to
specify that it be color coated. Standard Plans, Index 550-002 offers an option for PVC
(vinyl) coated fence fabric that is a soft gray color; however, you can specify the color to

Chapter 9: Stormwater Management Facility 9-26


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

be medium green, dark green, or black as allowed by AASHTO M 181. The posts, rails,
and fittings also can be color coated. To specify color-coated fence, use a pay item
footnote (0550-102x2 thru 0550-102x2, as applicable to the fence height required) similar
to the following:

Color coat the fence fabric, posts, rails, and fittings around the stormwater facilities
with xxx (state the desired color) PVC. Apply the PVC coating of the posts and
rails in addition to the standard metallic coating and ensure that it meets the
requirements of ASTM F 1043. The PVC coating of the fittings must meet the
requirements of ASTM F 626. Include the cost of the coating in the cost of these
items.

Fence Height and Barbed Wire Attachments:


The Department has no requirement for the height of the fence surrounding a stormwater
facility, nor does it require the use of barbed wire attachments on a fence surrounding a
stormwater facility. Other regulatory agencies may have applicable requirements
regarding fence height and barbed wire attachments.

9.2.3.2 Debris Collection


Discuss with maintenance personnel and the District Landscape Architect the need to
collect debris near the inflow pipe to the pond to prevent the debris from spreading. If it is
possible to collect the debris, direct it to one location where maintenance personnel can
easily remove it. Figure 9.2-3 shows some possible configurations.
Do not locate inflows and outlets near preserved existing vegetation or planted landscape
areas that have the potential to shed leaves, limbs, etc., that may clog pipes and
structures.

9.2.4 Aviation Safety Requirements


Per the Drainage Manual, when a prospective pond—wet or dry—is located within five
miles of an airport, the drainage designer must contact the District Aviation Coordinator
to ascertain any relevant airport design restrictions. The FAA requirements are targeted
to minimize the potential for bird strikes and are specific to the types of aircraft using the
airport and to the layout of the airport’s runways. The district aviation coordinators are
familiar with these requirements and will provide guidance to the drainage designer.

The best choice, in responding to FAA requirements, is to move the proposed pond
outside the glide paths of the air traffic. If this is imprudent, dry ponds are less attractive
to birds than wet ponds. Additionally, several design approaches are routinely used in wet
ponds to minimize attracting birds:

Chapter 9: Stormwater Management Facility 9-27


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

 Use steep, rocked slopes, typically 1:2, without littoral zones, to discourage the
presence of food sources for birds.

 Suspend nets over the surface of the pond to make the area less hospitable for
ducks, geese, and other water fowl.

 Ask districts for other techniques

Figure 9.2-3

Chapter 9: Stormwater Management Facility 9-28


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

9.3 STORMWATER QUALITY


9.3.1 Design Criteria
FDEP, the WMDs, and the delegated local governments have established design criteria
for the operations of stormwater management facilities. There are two main categories of
criteria: (1) water quality, and (2) water quantity (see Section 9.4). The criteria related to
water quality are based on research of rainfall and runoff in Florida and were established
to meet state water quality standards. See Appendix N for a discussion of the
development of the typical criteria.

Although the criteria are similar around the state, there is some variation. It is essential
that you become familiar with the applicable agency’s criteria. Read their manuals and
coordinate as necessary. Arrange a pre-application meeting to review the status of
applicable rules and to identify potential problems and concerns to be addressed during
design. Agencies usually have checklists and standard forms to be completed for a
stormwater permit. Review these forms and address the items relating to stormwater
management.

9.3.1.1 Treatment Volumes


Pollutants in stormwater runoff from urbanized areas generally exhibit a "first flush" effect.
This is a phenomenon where the concentrations of pollutants in stormwater runoff are
highest during the early part of the storm with concentrations declining as the runoff
continues. Substantial reductions in pollutant loads to the state’s waters will occur when
this first flush is captured and treated. Therefore, each method of treatment requires that
a volume of runoff be captured and treated before discharging to surface or ground water.
This volume is called the treatment volume.

In general, the treatment volume will vary depending on the classification of the receiving
water body and whether the volume is captured on-line or off-line. Sensitive water bodies
such as shellfish harvesting waters (Class II) and Outstanding Florida Waters require a
larger treatment volume. The classification of the receiving water body should be
identified in the Project Development phase as a part of the water quality impact
evaluation. FDEP includes a list of sensitive water bodies in Rule 62-302, F.A.C.

9.3.1.2 Special Conditions


Some of the Department’s districts have agreements with regulatory agencies regarding
treatment requirements for certain types of highway improvements, such as bridge
widening and intersection improvements. Check with the District Drainage Engineer to
see if your project is covered by an agreement.

Chapter 9: Stormwater Management Facility 9-29


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Compensatory treatment may be an option when trying to meet water quality regulations.
Sometimes, limited or very expensive right of way creates hardship conditions in which it
is unrealistic to provide the standard treatment. Sometimes, the Department can arrange
to provide compensatory treatment for an area that currently does not receive any
treatment. Providing this treatment compensates for not providing the standard treatment
in the area where the hardship condition exists. Treating a larger volume of runoff at
another location (drainage area) on the project usually is not considered replacement
treatment.

Nutrient-Impaired Basins

When designing stormwater systems that discharge to basins verified for nutrient
impairment, state law requires the applicant to demonstrate that there will be no net
increase of the pollutant of concern. To satisfy this requirement, all WMDs require a pre-
vs. post-development comparison of annual nutrient loading, using the Harper (2007)
Methodology, to demonstrate that the post-development annual loading is not greater
than the pre-development loading for the pollutant of concern. Guidance on this analysis
is contained in Section 9.3.4.6 of this document.

The BMPTRAINS software, developed by the UCF Stormwater Academy


(https://stormwater.ucf.edu/), can be used to analyze best management practice (BMP)
nutrient removal from different land uses. See Section 4.5.1 for an example. Software
results are readily accepted by permitting agencies around the state.

9.3.2 Concerns of Off-Line Systems


Although off-line treatment systems are preferred from a water quality standpoint and
sometimes require less treatment volume, they can complicate the design. You would
design off-line systems to bypass essentially all additional stormwater runoff volumes
greater than the treatment volume to the receiving water or an attenuation basin. The
bypass flow must pass over the weir of the diversion structure. This can present design
problems in that the weir may need to be very long to keep the hydraulic gradient at an
acceptable level. Skimmers need to be constructed in front of these weirs, further
complicating the practicality of long weirs.

Another concern is that there will be some additional attenuation storage in the off-line
basin associated with the hydraulic gradient of the peak flow passing over the weir. When
there is significant attenuation storage above the treatment volume, there is a concern
that the system will function more as an on-line system than as an off-line system due to
mixing. You could use metal or rubberized flap gates to address this concern, but they
can be a maintenance problem and a noncompliance issue if not carefully designed.

Chapter 9: Stormwater Management Facility 9-30


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

The outlet control structures of off-line systems are difficult to maintain simply because
they normally are placed in junction boxes. They are neither seen nor reached as easily
as the outlet control structures of on-line systems.

9.3.3 Seasonal High Water Table


Frequently, the first parameter considered in the design of a retention or detentionBMP)
is the location of the water table. Define the depth to the normal high water table and the
seasonal high water table (SHWT) to establish the appropriate type of BMP and the
needed treatment volume. The SHWT is critical to the operation of all of the treatment
methods described below. The control (or normal) water elevation of wet detention
systems is related to, and sometimes set at, the SHWT. The SHWT is a critical factor in
calculating the recovery time of the treatment volume in a retention system. For filtration
systems, the lowest point of the underdrain pipe should be at least one foot above the
SHWT.

Use the NRCS soil surveys, project-specific soil investigations, and field observations
(vegetative indicators, observation wells, etc.) to estimate the SHWT. Recognize,
however, that soil staining may denote a relic or historic water table that has since been
lowered by other drainage features in the region.

9.3.4 Treatment Methods


The treatment methods most commonly used by the Department are wet detention,
retention, filtration, and exfiltration. Refer to Chapter 7 of this document for exfiltration
system BMPs. The type of soil and the SHWT control the selection of the treatment
method. The following figure provides qualitative guidance for the selection.

 Surface   Surface 

WET DETENTION ONLY

FILTRATION
SHWT

OR
RETENTION
WET DETENTION

Clay soils Sandy Soils

Low ------Hydraulic Conductivity (K)-------- High

Chapter 9: Stormwater Management Facility 9-31


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

As shown, wet detention is the only option in areas where the SHWT is near the surface.
However, wet detention also may be appropriate in areas where the SHWT is far from the
surface and clay soils exist. The use of retention requires that the SHWT be far from the
surface and that sandy soils exist. Filtration requires that the SHWT be far from the
surface unless impermeable liners are used. The Department does not encourage the
use of liners, although their use is justifiable sometimes. Consult the District Drainage
Engineer before proposing liners.

You cannot apply specific values to this figure because site-specific factors—such as
pond shape, groundwater boundary conditions, and drainage basin characteristics—need
to be considered. Situations exist where both filtration and wet detention are suitable. In
these cases, the Department should weigh and balance other factors—such as right-of-
way costs, property owner preference, and long-term maintenance costs—to select the
most appropriate treatment method.

9.3.4.1 Wet Detention Systems


These systems are permanently wet ponds that are designed to slowly release the
treatment volume through the outlet control structure. The pollutants are removed by
physical, biological, and chemical assimilation. Specifically, pollutant removal processes
that occur within the permanent pool include uptake of nutrients by algae and wetland
vegetation, adsorption of nutrients and heavy metals onto bottom sediments, biological
oxidation of organic materials, and sedimentation.

Advantages Disadvantages
1. Very effective at removing dissolved and 1. Treatment requirements are typically
suspended pollutants. double the requirements for retention and
filtration.
2. High probability to function as designed. 2. Depth of the treatment volume is
sometimes limited to 1.5 feet.
3. Recovery of treatment volume is easily 3. Because of the above items, right-of-way
predicted. requirements are greater than other
methods.
4. Easy and low-cost long-term 4. Sometimes requires planting of the littoral
maintenance. zone.
5. Produces on-site fill material for project 5. Creates a potential mosquito habitat.
needs

Despite the disadvantages, the Department encourages the use of wet detention.

Chapter 9: Stormwater Management Facility 9-32


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

The average length-to-width ratio of the pond should be at least 2:1. Maximize the flow
path of water from the inlet to the outlet to promote good mixing and avoid “dead” storage
areas. If you cannot avoid short flow paths, use the littoral shelf to increase the effective
flow path, provided this is acceptable to the regulatory agency. Figure 9.3-1 shows
examples of pond configurations.

Per the regulatory agency requirements, you may need to plant the littoral shelf. If so,
consult with the District Landscape Architect.

Chapter 9: Stormwater Management Facility 9-33


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Figure 9.3-1: Wet Detention Configurations

Chapter 9: Stormwater Management Facility 9-34


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

9.3.4.2 Retention Systems


A retention system is designed to store the treatment volume, allowing it to infiltrate into
the soil. Soil permeability, water table conditions, and the depth to any confining layer
must be such that the retention system can infiltrate the treatment volume within a
specified time following a storm event. After the pond completes the drawdown, the basin
does not hold water; thus, the system is normally “dry.” Unlike wet detention systems and
filtration systems, the retention system will discharge the treatment volume into the
ground, not to surface waters.

Most regulatory agencies require that the treatment volume be available within 72 hours
after a storm. See Section 9.4.6.1 on the subject of groundwater flow from retention
systems and a recommended approach to modeling recovery of the treatment volume.

9.3.4.3 Filtration/Underdrain Systems


A filtration system is designed to treat the water quality volume, allowing it to pass through
a sand filter. It differs from a retention system in that the treatment volume is not infiltrated
into the soil, but instead discharges to surface water. After passing through the sand filter,
the water collects in perforated pipes that discharge to surface water. The Department’s
standard underdrain is shown in Standard Plans, Index 440-001.

Compared with the previous two treatment methods discussed, underdrains are the least
reliable. They are subject to clogging during and after construction and are difficult to
maintain. Vehicle loads can crush the underdrain pipes. Filtration systems also do not
remove dissolved constituents, such as nitrogen and phosphorus, and therefore do not
count toward load reduction credit in impaired basins. The Department realizes that using
underdrains is sometimes necessary due to clayey soils, but encourages a thorough
evaluation of other treatment methods first.

Configuration:

When you use side bank underdrain (Standard Plans, Index 440-001, Type Va), slope
the pond bottom up from the underdrain. This will reduce the saturated soil condition and
localized ponding associated with a flat pond bottom. It also increases the chances of
sustaining a stand of grass on the bottom. See Figure 9.3-2.

If feasible, construct underdrains out of the primary flow path to avoid directing debris and
sediments there.

To account for construction tolerances, the underdrain pipe should be placed on a slope.
Specify flow lines for the pipe at the beginning, at bends, and at the end of the underdrain.

Chapter 9: Stormwater Management Facility 9-35


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

In all but very short runs of underdrain, the flow line should drop six inches or more to
account for construction tolerances.

Figure 9.3-2: Bottom Configurations with Side Bank Underdrains

Design Technique

Hydraulic Conductivity of the Fine Aggregate Media:

For design purposes, use K = 0.5 ft/hr as the hydraulic conductivity of the fine aggregate
media. This does not include the factor of safety of two required by the regulatory
agencies. You do not have to apply that factor of safety to the hydraulic conductivity. It is
sometimes applied to the length of the underdrain or to the time to draw down the
treatment volume. You could refine the above value by experience from permeability
testing of locally available fine aggregate media meeting the requirements of the standard
specifications for underdrain filter material.

Determining the length of underdrain required is a trial-and-error process and can be


accomplished by using the following procedure with Table 9.3-2.

1. Develop incremental storage volumes from the maximum elevation of retention


storage (i.e., lowest elevation of the outlet control structure) down to the pond
bottom. Record these in Column 1 through Column 3 of Table 9.3-2.

Chapter 9: Stormwater Management Facility 9-36


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

2. Determine the effective head (HE), the average flow length (LAVG), and the average
width (WAVG) for flow paths through the underdrain. Determine these for each water
surface elevation considered in Step 1. See the discussion following Step 10 for a
suggested approach to determining these values. Record these in Column 4
through Column 6.
3. Calculate the hydraulic gradient (i) for each water surface elevation considered in
Step 1 using the values determined in Step 3, and record the results in Column 7.
Hydraulic gradient (i) = HE/LAVG.
4. Assume an underdrain pipe length (L) and calculate the area of filter (A) for each
water surface elevation considered in Step 1. Record results in Column 8.
5. Calculate the Darcy flow (Q) using the hydraulic conductivity (K), the hydraulic
gradient, and the filter area for each water surface elevation considered in Step 1.
Record results in Column 9.
6. Calculate the average flow rate for each depth interval and record results in
Column 10.
7. Divide the incremental storage volume (ΔV) from Column 3 by the average flow
rate from Column 10 to obtain the incremental time (ΔT) to draw down that storage
volume. Record results in Column 11.
8. Sum the incremental drawdown times recorded in Column 11 to obtain the
drawdown time (ΣT). Record results in Column 12.
9. If the total computed drawdown is longer than required, increase the underdrain
length and return to Step 5.
10. Size the underdrain pipe to handle the design flow rate.
Determining the Effective Head, Average Flow Length, and Average Width:

Bottom Underdrain (Type Vb):

To determine the effective head (HE) at a given water surface, use the vertical distance
from the water surface to the bottom of the fine aggregate material. For the average flow
length (LAVG) through the filter, use the depth of fine aggregate, 2.0 feet. For the average
width (WAVG) of filter normal to flow, use the standard width of 1.5 feet unless you use
non-standard geometry.

Side Bank Underdrains (Type Va):

The standard plans index shows the upper and lower limit to side bank underdrain. Try to
avoid using the upper limit configuration because of its limited flow capacity in low head

Chapter 9: Stormwater Management Facility 9-37


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

conditions. There is very little head and the length of the filter material through which the
water must pass is long, resulting in a very small hydraulic gradient.

1. Make a scaled drawing of the average cross section geometry. One is shown in
Figure 9.3-3. The average should represent the midpoint between the high end
and low end of the underdrain.
2. For the effective head (HE) at a given water surface elevation, use the vertical
distance from the water surface to the pipe centerline. At high heads, this is non-
conservative because the free draining effect of the course aggregate reduces the
head. At low heads, this is a reasonable assumption.

The combined effect of using HE & LAVG as described here should result in
conservative flow rates in low head conditions and reasonable rates in high
head conditions. At high heads, the non-conservatism of using the effective
head (HE) to the center line of the pipe is offset by using an average length
(LAVG) that is longer than the actual distance through the fine aggregate. At
low heads, the conservatism of using a longer-than-actual average length
(LAVG) is justified because this zone of the filter is most likely to receive
sediment and clog.

3. For the average flow length (LAVG) through the filter at a given water surface, use
the average of several straight-line distances from the outside of the pipe to the
top of the fine aggregate. This is conservative because it ignores the course
aggregate, which is relatively free draining. Refer to Figure 9.3-3 and Table 9.3-1
for an example.
4. For the average width (WAVG) of filter normal to flow, use the average of the
saturated fine aggregate area. Due to the complex transition between vertical and
horizontal flow, this is best determined by “visually” estimating the average width
based on your scaled drawing. Refer to Figure 9.3-3 and Table 9.3-1 for an
example.

Chapter 9: Stormwater Management Facility 9-38


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Table 9.3-1: Average Flow Length and Average Width through Side
Bank Underdrain

Water L AVG W AVG


Surface Avg. Flow Length through Avg. Width of Filter
Elevation Filter Normal to Flow
WSE-5 or
(L5 + L4 + L3 + L2 + L1 + L0) / 6 W to W5
above
WSE-4 (L4 + L3 + L2 + L1 + L0) / 5 W to W4
WSE-3 (L3 + L2 + L1 + L0) / 4 W to W3
WSE-2 (L2 +L1 + L0) / 3 W2A to W2B
WSE-1 (L1 + L0) / 2 W1A to W1B
Refer to Figure 9.3-3.

Figure 9.3-3: Side Bank Underdrain (Shown 6” Below Upper Limit)

Chapter 9: Stormwater Management Facility 9-39


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Table 9.3-2: Drawdown Worksheet for Underdrain

Chapter 9: Stormwater Management Facility 9-40


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

9.3.4.4 Stormwater Re-Use Systems


These systems represent wet ponds that provide water for re-use—either for irrigation,
alternative water supply, or supplemental water for reclaimed wastewater lines. A wet
detention pond can be converted to a re-use pond simply by plugging the bleeder and
pumping the treatment volume to its designated re-use. Since there is less discharge
volume compared to a standard wet detention pond, there is less mass of pollutant being
released from the stormwater re-use pond.

9.3.4.5 Regional Stormwater Pond Systems


Regional stormwater ponds, by definition, provide water quality treatment for a significant
portion of the upstream basin, not just the onsite FDOT project. These ponds often are
located downstream of the FDOT project, avoiding the more expensive land adjacent to
the state highway. Typically, this approach includes the cooperation of a local government
that assumes ownership and perpetual maintenance of the pond. FDOT holds a storage
easement, prescribing a needed storage volume below a certain design elevation.
Multiple gains result from this cooperative approach:

1. FDOT is relieved of ongoing property liability and maintenance responsibility.

2. The downstream waterway enjoys improved water quality.

3. Property adjacent to the state highway, previously targeted for usage as a pond,
is available for development.

4. Oftentimes, stormwater re-use is facilitated by a single, larger stormwater pond.

Regional treatment facilities can be difficult to permit because of the Class III treatment
requirements of conveyance facilities between the FDOT site and the location of the
regional pond. These intermediate waterway requirements sometimes can be eliminated
by classifying the manmade, intermediate conveyance waterways as part of the
stormwater system, thereby severing jurisdiction. The cooperation of the WMD will be
essential in such efforts.

9.3.4.6 Harper (2007) Methodology for Nutrient Loadings


Computation
The 2007 Harper Methodology was the computational foundation for the 2009 Statewide
Stormwater Rule. The rule was not implemented, but the Harper Methodology has been
accepted by the WMDs and FDOT as a best practice for estimating annual nutrient
loadings. Details of the methodology are outlined in the draft March 2010 Applicant’s
Handbook posted on the state drainage website, under Design Aids:
http://www.fdot.gov/roadway/Drainage/ManualsandHandbooks.shtm

Chapter 9: Stormwater Management Facility 9-41


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

The above draft publication is referenced ONLY for its helpful outline of the background
rationale and computational steps involved in the Harper Methodology, NOT for
regulatory requirements.

Since 2010, event mean concentrations (EMCs) for different land uses have changed as
additional data have become available. Current general EMCs are tabulated below:

Table 9.3-3: Example EMC Values for Different Land Uses


Event Mean Concentration (EMC)
mg/L
LAND USE
CATEGORY*
Total Nitrogen Total Phosphorus

Single-Family 2.07 0.327


Multi-Family 2.32 0.520
High-Intensity Commercial 2.40 0.345
Light Industrial 1.20 0.260
Highway 1.64 0.220
Agricultural—Citrus 2.24 0.183
Agricultural—Row Crops 2.65 0.593
Agricultural—General
Agriculture 2.79 0.431
Undeveloped 1.15 0.055
*Numbers may vary as more information becomes available or for specific locations.

The BMPTRAINS computer program was developed to employ the latest policy and
methodology for assessing nutrient loadings and BMP performance related to nutrient
removal. The program is available on-line at the UCF Stormwater Management Academy
website: http://stormwater.ucf.edu/

The program includes helpful tutorials and a user’s manual.

Additional helpful tools sponsored by the Academy are available under the title
Stormwater Management and Design Aids (www.SMADAonline.com). A partial list of
programs in the SMADA online package is below:

1. BMPTRAINS “Light”—Used to select one BMP with an estimate of nutrient


pollutant removal and in the selection of BMPs for net improvement or pre/post
analysis.
2. BMP performance evaluation

Chapter 9: Stormwater Management Facility 9-42


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

3. Rainfall distributions and IDF curves


4. Statistical analyses such as regression and frequency distributions
5. Time of concentration
6. Hydrograph generation
7. Unit hydrograph generation
8. Transport pipe and channel flow and sizing
9. Pollutant load calculations
10. Storm sewer design and analysis

9.3.4.7 Protection of Springsheds from Nitrates


The Harper Methodology targets annual loadings of nutrients to surface waters, making
the assumption that nutrients infiltrated into the ground via retention systems are
“removed.” For springsheds, nitrates infiltrated into the ground are the critical transport
mechanism for springshed impairment. Nitrate-removing retention BMPs currently are
under development using bio-activated media (BAM). Until design methodology is
released, contact your local District Drainage Engineer for guidance when designing
retention ponds within Karst springshed geology.

Examples Illustrating the Use of BMPTRAINS for Nutrient Loading Analysis

FDOT has extracted relevant design criteria and combined them into one reference
publication and computer program, named BMPTRAINS. The design engineer should
verify the design criteria at a pre-meeting with the WMD or FDEP, since newer regulations
may exist. The BMPTRAINS model provides the option to over-ride existing criteria and
assumptions. An example of an assumption is the event mean concentration (EMC) data.

Example Problem 9.3-1: Wet detention, net improvement

A wet detention pond serves a section of a two-lane highway that is about 1,100 feet long
and the right-of-way width is about 200 feet. The catchment area is five acres and is part
of a larger watershed that may impact the design. The existing portion of highway was
not served by any treatment system. The existing and proposed portion of the highway
will be treated in the post-development condition. The site is located in West Palm Beach,
Florida, on Hydrologic Soil Group D. The existing land use condition is assumed to be a
highway with a non-DCIA Curve Number of 80 and 40 percent DCIA. The post-
development land use condition is assumed to be a highway with a non-DCIA Curve
Number of 80 and 85 percent DCIA. The area needs net improvement using a wet
detention pond with a littoral zone (assumed 10 percent removal efficiency credit for the

Chapter 9: Stormwater Management Facility 9-43


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

littoral zone) in the design. The area and depth for the pond allowed an average annual
pond residence time of 50 days.

First, identify the meteorological zone, which is Zone 5, and the mean annual rainfall,
which is 61 inches, as shown in Table 9.3-3.

General Site Information for Example Problem 9.3-1

Next, the catchment site information data are summarized below.

Watershed Characteristics for Example Problem 9.3-1

Chapter 9: Stormwater Management Facility 9-44


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Using the BMPTRAINS program, the net improvement expected with the wet detention
design is 71.5 percent removal of total phosphorus and 46.2 percent removal of total
nitrogen, as shown in the BMPTRAINS program screenshot below. Note that, for the wet
detention option, a residence time greater than 50 days will only marginally improve
removal. Thus, a design criterion of 50 days annual residence time is above the minimum
required by the water management district (21 days x 1.5 = 31.5 days) and can fit within
the existing right of way.

Wet Detention Pond Effectiveness for Example Problem 9.3-1

Chapter 9: Stormwater Management Facility 9-45


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Example Problem 9.3-2: A highway receiving runoff from an industrial park that will
have retention systems in a series—a pre- vs. post-development loading analysis
is required

The water table conditions in this area are suitable for retention systems. An exfiltration
trench in series with a retention basin can serve a five-acre light-intensity commercial site.
The catchment also can contain 10 tree wells along the road. The tree wells are designed
to be three feet deep with a six-inch depth above soil column. The length and width of the
tree wells are designed to be four feet each. Use a 0.2 sustainable water storage capacity
of the soil. Treat the tree wells as retention systems. The site is located in Orlando,
Florida, on Hydrologic Soil Group C. The existing land use condition is assumed to be
undeveloped-dry prairie with a non-DCIA Curve Number of 79 and 0.0 percent DCIA. The
post-development land use condition is a low-intensity commercial area with a non-DCIA
Curve Number of 85 and 65 percent DCIA. The combination of treatment systems is to
provide treatment sufficient to match the post-development annual nutrient loads with the
pre-development annual nutrient loads.

First identify the meteorological zone and annual rainfall for the project.

General Site Information for Example Problem 9.3-2

A summary of the catchment characteristic data is shown below. Note that the catchment
is highly developed, leaving no feasible space for a retention basin or other land-intensive
BMP. You need to consider BMPs that are more useful for ultra-urban environments.

To meet pre/post conditions, the required phosphorus removal is calculated as 70.3


percent and the required nitrogen removal is 89.3 percent. These are calculated knowing
the Event Mean Concentrations and the pre- and post-runoff volumes.

Chapter 9: Stormwater Management Facility 9-46


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Catchment Characteristics Data for Example Problem 9.3-2

The tree wells receive runoff water first and, thus, the effectiveness associated with a
design is examined first. The capture effectiveness is low (1.3 percent) because of the
number and size of the catchment. The results are shown below.

Effectiveness of 10 Tree Wells in the Catchment or Watershed of Example


Problem 9.3-2

Note: As the BMPTRAINS model is improved, output screens may change.


VEGETATED AREAS (Example Tree Wells):
Vegetated Areas (tree wells or similar) for: Facility handbook example 2
Catchment 1 Catchment 2 Catchment3 Catchment 4
Contributing catchment area: 5.000 0.000 0.000 0.000 ac
Required treatment efficiency (Nitrogen): 70.300 %
Required treatment efficiency (Phosphorus): 89.257 %
Vegetated Area (Tree Well) depth 3.00 ft
Vegetated Area (Tree Well) water depth above soil column: 0.50 ft
Vegetated Area (Tree Well) length: 4.00 ft
Vegetated Area (Tree Well) width: 4.00 ft
Sustainable water storage capacity of the soil: 0.20
Number of similar Areas within watershed: 10.00
Retention depth for provided hydraulic capture efficiency: 0.010 0.000 0.000 0.000 in
Is this a retention or detention system? Retention
Type of soil augmentation: View Media Mixes
Provided treatment efficiency (Nitrogen): 1.307 0.000 0.000 0.000 %
Provided treatment efficiency (Phosphorus): 1.307 0.000 0.000 0.000 %
Is/are the vegetated areas sufficient? NO

Chapter 9: Stormwater Management Facility 9-47


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Next, you can use an exfiltration system to collect some of the runoff water. Using the
geometric design for the exfiltration, the retention storage volume is calculated as 0.55
inches. The effectiveness of using the exfiltration design is shown below.

Exfiltration Trench Design and Effectiveness for Example Problem 9.3-2

EXFILTRATION TRENCH:
EXFILTRATION TRENCH SERVING ENTIRE CONTRIBUTING WATERSHED:

Contributing watershed area: 5.000 ac


Required Treatment Eff (Nitrogen): 70.300 %
Required Treatment Eff (Phosphorus): 89.257 %
Required retention for the entire watershed to meet required efficiency: 1.756 in
Required water quality retention volume: 0.732 ac-ft

EXFILTRATION TRENCH FOR MULTIPLE TREATMENT SYSTEMS (use only if other BMP
method is oversized or undersized) :

Provided retention depth: 0.550 in


Provided treatment efficiency (Nitrogen): 56.620 %
Provided treatment efficiency (Phosphorus): 56.620 %
Remaining treatment efficiency needed (Nitrogen): 31.535 %
Remaining treatment efficiency needed (Phosphorus): 31.535 %
Remaining retention depth needed if retention: 1.206 in

Finally, you can use a retention basin at the discharge from the watershed. The land for
the retention basin within the watershed is part of the industrial park but did not provide
sufficient removal by itself to meet post-equal-pre average annual mass loading. The
retention basin can hold 0.50 inches of runoff and, thus, was limited to the removal
effectiveness associated with that volume of storage. Using a combination of retention
basin, tree wells, and an exfiltration trench provided by the roadway was sufficient to meet
the post-equal-pre requirements.

The retention options are in a series. The first flush of water is captured by the tree wells
and what is not captured is routed to the inlets for the exfiltration trench. The exfiltration
trench can handle a fraction of that runoff, so the bypassed water is routed to the retention
basin. All of these retention BMPs are designed to be off-line BMPs. Summary results of
the tree well, exfiltration, and retention basin designs with the overall effectiveness
removal are shown below.

Chapter 9: Stormwater Management Facility 9-48


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Summary Loadings and Removal Effectiveness for Example 9.3-2

Summary Performance
Catchment
A - Single Catchment
Configuration 3/25/2014
Nitrogen Pre Load (kg/yr) 5.35 BMPTRAINS MODEL
Phosphorus Pre Load (kg/yr) 0.29
Nitrogen Post Load (kg/yr) 18.00
Phosphorus Post Load (kg/yr) 2.73
Target Load Reduction (N) % 70
Target Load Reduction (P) % 89
Target Discharge Load, N (kg/yr) 5.35
Target Discharge Load, P (kg/yr) 0.29 1
Provided Overall Efficiency, N (%): 57
Provided Overall Efficiency, P (%): 57
Discharged Load, N (kg/yr & lb/yr): 7.72 17.00
Discharged Load, P (kg/yr & lb/yr): 1.17 2.58
Load Removed, N (kg/yr & Ib/yr): 10.28 22.65
Load Removed, P (kg/yr & Ib/yr): 1.56 3.44

For additional example problems, see the User’s Manual to be used with the BMPTRAINS
model (www.stormwater.ucf.edu) located at:

BMPTRAINS Stormwater Best Management Practices Analysis Model (Latest


Version) Registration, Model, and User's Manual.

Chapter 9: Stormwater Management Facility 9-49


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

9.4 STORMWATER QUANTITY CONTROL


9.4.1 The Department’s Design Storms
A problem with developing a design storm distribution is that actual storms have an
unlimited combination of durations and intensity patterns. What should the duration of the
design storm be? Should the peak rainfall occur near the beginning, in the middle, or near
the end of the storm? Should there be multiple peaks?

Most of the current widely used rainfall distributions address this by nesting short-
duration, high-intensity storms in the middle of a long duration storm, although very
intense peaks do not usually occur in long storms. You usually would place the largest
intensity value in the middle of the storm pattern, then place the remaining values
alternately before and after this point, in order of decreasing intensity. The various NRCS
distributions, the South Florida Water Management District (SFWMD) three-day
distributions, and the St. Johns River Water Management District (SJRWMD) four-day
distributions are examples of design storm distributions created using this approach.
These “nested” distributions are not indicative of actual rainfall patterns and subsequently
may produce inaccurate representations of actual runoff characteristics.

You may have used these distributions in the past for the design of conveyance systems
because they give conservatively high runoff estimates. But, when you use these
distributions to determine the pre-developed discharge, they can overestimate it. In the
developed condition, the outlet control structure would be designed to pass the
“overestimated pre-developed discharge,” thereby discharging more in the post-
developed condition.

Another problem with these distributions is that different drainage areas will react
differently to the same rainfall pattern. Small basins with short times of concentration and
little storage will have higher runoff rates from short, intense storms than from long-
duration, low-intensity storms. Long-duration, low-intensity storms usually do not
generate peak discharges from small basins. The opposite is true for large basins. Very
large basins with large amounts of storage will have less runoff from short, intense storms
than from long-duration, low-intensity storms. Large river systems and static water bodies
such as lakes reach peak stages when extreme antecedent conditions exist and
variations in intensity usually do not affect their stages.

To overcome the concerns of a single design storm distribution, the Suwannee River
Water Management District (SRWMD) developed a series of design distributions to better
reflect actual rainfall patterns. They developed distributions for 1-, 2-, 4-, and 8-hour
storms and for 1-, 3-, 7-, and 10-day storms using National Oceanic and Atmospheric
Administration (NOAA) hourly and sub-hourly data. SRWMD requires the use of these
distributions for projects within the district.

Chapter 9: Stormwater Management Facility 9-50


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Chapter 14-86, Florida Administrative Code

In 1986, the Department established Chapter 14-86 of the F.A.C., requiring adjacent
developments to maintain discharges at or below pre-developed discharges using a
multiple storm approach. In the Department’s Drainage Connection Handbook (February
1987), the SRWMD design distributions mentioned above were accepted as appropriate
for the entire state. These distributions can be found at the Department’s website, listed
below: http://www.fdot.gov/roadway/Drainage/files/IDFCurves.pdf

In a July 1988 memorandum, the State Roadway Design Engineer directed the districts
to design the Department’s stormwater management systems to Chapter 14-86. In
October 1992, the Drainage Manual was revised to require the design of the Department’s
stormwater management systems to comply with Chapter 14-86. In 2013, the Drainage
Manual was amended to require the application of Chapter 14-86 on FDOT stormwater
designs only for closed basins and areas where downstream historical flooding is
documented.

9.4.1.1 Critical Duration


Since the time the Department developed Chapter 14-86, there have been two
interpretations of the critical duration and how to apply the multiple storm concept. The
definition of critical duration (shown below), as defined in Chapter 14-86, lends itself to
two interpretations.

“Critical Duration” means the duration of a specific storm event (i.e., 100-year storm) that
creates the largest volume or highest rate of net stormwater runoff (post-development runoff
less pre-development runoff) for typical durations up through and including the 10-day
duration event. The critical duration is determined by comparing various durations of the
specified storm and calculating the peak rate and volume of runoff from each. The duration
resulting in the highest peak rate or largest total volume is the “critical duration” storm.

Chapter 9: Stormwater Management Facility 9-51


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

(A) Peak Discharge Approach


This interpretation of critical duration and the multiple storm concept allows a post-
developed runoff rate, for a given frequency, that is equal to or less than the highest pre-
developed runoff rate of any duration. For example, given the pre-developed runoff rates
shown in the table below, the allowable runoff rate would be 70, regardless of the duration
associated with the peak post-developed runoff rate. The post-developed runoff rates
shown are acceptable because none are greater than 70. You need only run enough
durations in the post-developed condition to be assured that runoff rates of the other
durations do not exceed the allowable.

Acceptable Post-
Pre-Dev Runoff
Duration Dev Runoff XX
XX Year Event
Year Event
1-hour 65
2-hour 70 60
4-hour 66 70
8-hour 60 65
24-hour 30 35
3-day 25
7-day 24
10-day 21

This approach is consistent with the last sentence of the definition of critical duration. “The
duration resulting in the highest peak rate . . . is the critical duration.” With this approach,
the pre-developed critical duration can be different from the post-developed critical
duration, as shown in the values above. Also, the pre-developed runoff rate could be
calculated with the rational method (Q = CIA) for small basins; therefore, it would not be
directly associated with any of the eight durations. The examples in the Drainage
Connection Handbook follow this interpretation.

The above discussion pertains to discharges to open basins only with historical flooding
documented. For discharges to closed basins, a similar approach is used with an
additional constraint on the runoff volume. For a given frequency, the allowable post-
developed runoff volume is the largest pre-developed runoff volume of any duration.
When using the NRCS technique for computing runoff, the 10-day duration event will
always produce the largest runoff volume and, therefore, be the critical duration. But, for
other more-refined approaches to modeling infiltration, the critical duration could be
something other than the 10-day duration.

Chapter 9: Stormwater Management Facility 9-52


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

(B) Storm for Storm Approach (Preferred)


This interpretation of critical duration and the multiple storm concept requires, for a given
frequency, that the post-developed runoff rate for each duration be less than or equal to
the pre-developed runoff rate of corresponding duration. For example, in the table below,
the allowable runoff rate for each duration is the pre-developed runoff rate. The post-
developed runoff rates shown are acceptable because they are all less than or equal to
the pre-developed runoff rate of corresponding duration. The 4-hour duration is critical
because it most closely matches the pre-developed runoff rate.

Acceptable Post-
Pre-Dev Runoff
Duration Dev Runoff XX
XX Year Event
Year Event
1-hour 65 60
2-hour 70 68
4-hour 66 66
8-hour 60 57
24-hour 30 26
3-day 25 23
7-day 24 22
10-day 21 20

This approach is consistent with the first sentence of the definition of critical duration.
“Critical Duration means the duration . . . that creates the . . . highest rate of net
stormwater runoff (post-development runoff less pre-development runoff). . . .” In the
example above, when you subtract the pre-development runoff rate from the
corresponding post-development runoff rate, all the “net stormwater runoff” values are
negative except the 4-hour duration, which has zero “net stormwater runoff.” So, the 4-
hour duration has the highest rate of net stormwater runoff; therefore, it is the critical
duration. This approach is better than the peak discharge approach, where the release
timing of the facility is critical. FHWA’s Hydraulic Engineering Circular No. 22 (HEC 22)
contains a discussion of the concern for release timing.

The above discussion pertains to discharges to open basins only with historical flooding
documented. For discharges to closed basins, a similar approach is used with an
additional constraint on the runoff volume. For a given frequency, the post-developed
runoff volumes for each duration cannot exceed the pre-developed runoff volumes of
corresponding duration.

Chapter 9: Stormwater Management Facility 9-53


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Although both the Peak Discharge Approach and the Storm for Storm Approach have
been applied to FDOT projects in the past, the Department prefers that you use the Storm
for Storm Approach on its projects. The examples in Section 9.4 are based on the Storm
for Storm Approach.

9.4.1.2 Storm Frequencies


The previous sections primarily discuss durations and the multiple storm concept.
Chapter 14-86 [14-86.003 (3)(c) 2 & 3] requires that we consider various rainfall event
frequencies up to and including the 100-year event. The rule does not say that all
frequencies must be evaluated.

The more frequent FDOT design storms (2-year to 50-year) do not usually control the
size of the pond because the runoff from these storms is less than the runoff for the 100-
year storm. The purpose of evaluating the less frequent storms is to ensure that the pre-
developed discharges are not exceeded. And so it becomes a check of the operation of
the outlet control structure under various rainfall event frequencies.

Where the discharge is controlled by a simple rectangular weir (one with a constant
width), it may be reasonable to run only the 2-year, 25-year, and 100-year events. Where
the discharge is controlled by a complex weir (width varies with elevation), an orifice, a
pipe, tailwater conditions, or any combination of these, evaluate all frequencies (2-year,
5-year, 10-year, 25-year, 50-year, and 100-year). Some software programs can run all
the frequencies at once. If these programs are available to you, run all the frequencies,
regardless of the outlet control structure configuration.

9.4.2 Estimating Attenuation Volume


A first step in estimating attenuation volume is identifying outfalls and their associated
drainage basins. At this stage, consider if it will be necessary to divert runoff from one
basin to another. Although the Department does not encourage diverting runoff, doing so
sometimes allows the Department to provide stormwater management (treatment and
attenuation) in more economical locations. For example, an economical parcel for a pond
site may be available in one drainage basin while the parcels in an adjacent basin are
very expensive. Diverting some roadway runoff to the economical parcel basin from the
expensive parcel basin may be more economical even when other costs, such as
construction and maintenance, are considered. Before you propose diverting runoff, be
sure it is acceptable to the regulatory agency.

When diverting runoff, be careful how you calculate the allowable discharge. Base your
allowable (pre-developed) discharge calculations on the pre-developed drainage area
that discharges to the proposed outfall. If an area does not drain to the proposed outfall
in the pre-developed condition, do not include that area in the allowable (pre-developed)

Chapter 9: Stormwater Management Facility 9-54


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

discharge calculations. Therefore, in a basin you divert runoff to, the pre-developed
drainage area is smaller than the post-developed drainage area. Conversely, in a basin
you divert runoff from, the pre-developed drainage area is larger than the post-developed
drainage area.

The actual attenuation volume cannot be determined until you “route” the design storms
and design the pond. There are several methods for estimating the attenuation volume.
The methods more commonly used on the Department’s projects are discussed below.

9.4.2.1 Pre Versus Post Runoff Volume


A common technique for estimating attenuation volume is to calculate the difference in
runoff volume between the post-developed conditions and the pre-developed conditions
using the NRCS equation for runoff.

 P - 0.2S )
2

QR =
P + 0.8S

As written, this assumes the initial abstraction (Ia ) = 0.2S & S =(1000/CN) – 10

where:

QR = Runoff depth (in inches)

P= Rainfall depth (in inches); Use the 100-year, 24-hour depth for evaluating
alternate drainage schemes or pond sites

S= Maximum retention or soil storage (in inches)

CN = Watershed curve number

The runoff volume is determined from: VOL = QR Drainage Area

A similar approach can be taken using the Rational Equation Method.

VOL = (CPOST – CPRE) P Drainage Area

An advantage of this technique is that it does not involve any design storm distributions.
So there is no concern for which storm duration is critical. On the other hand, this
technique ignores the timing differences between the pre-developed and post-developed
hydrographs. As a result, it may underestimate the attenuation volume.

Chapter 9: Stormwater Management Facility 9-55


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Example 9.4-1: Estimating attenuation volume using differences in runoff volume

Given:
 Pre-developed roadway pavement = 2 10-foot lanes
 Drainage area: includes roadway right of way &
off-site drainage to the roadway = 10.4 ac

For preliminary pond sizing, use the information from the old drainage map unless
you have reason not to.

 Offsite land use = Residential lots averaging 1/2 ac


 Proposed typical section = 5-lane urban section; Combined roadway, curb, and
sidewalk width = 83 ft
 Proposed right-of-way width = 100 ft
 Length of roadway within drainage area = 1,706 ft
 Offsite runoff draining to the project will be taken through the pond, not bypassed
around.
 Project located in Somewhere City, Florida, flat terrain <1 percent grade, Hydrologic
Soil Group B/D, project drains to open basin.

Example 9.4-1

Chapter 9: Stormwater Management Facility 9-56


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Find: The estimated attenuation volume.

1. Pre-developed area & curve number:


Roadway pavement: = 0.79 ac @ CN = 98 (20 ft x 1,706 ft)
Pervious area = 9.61 ac @ CN = 85 (10.4 ac – 0.79 ac)
Proposed pond area = 0.77 ac @ CN = 85
Total = 11.2 ac @ CN = 85.9

Assume the pond area is 20 percent of the roadway right of way (0.2 x 1,706 ft x
98 ft = 0.77 ac). For this example, the proposed pond is located outside the area
draining to the roadway; thus, the pond must be added to the other areas.

For this example, the roadway right of way to be acquired is within the area
draining to the roadway. For your project, the acquired right of way may be
outside the area draining to the road, thereby requiring that the additional right of
way be added to the other areas.
2. Post-developed area and curve number:
Roadway, curb, and sidewalk: = 3.24 ac @ CN = 98 (82.7 ft x 1,706 ft)
Pervious area = 7.17 ac @ CN = 85 (10.4 ac – 3.24 ac)
Pond area = 0.77 ac @ CN = 98
Total = 11.2 ac @ CN = 89.7
3. Calculate the difference in runoff volume between the pre-developed conditions
and post-developed conditions for the 100-year, 24-hour storm using the NRCS
equation for runoff.

Refer to the NOAA website link in Section 1.4 of the Drainage Manual to obtain
location-specific precipitation data for the 100-year, 24-hour volume. For this
example, the 100-year, 24-hour volume for Somewhere City, Florida, is 10.7
inches.

Q = (P – 0.2S)2  (P + 0.8S) where: S = (1,000  CN) – 10

Pre Post
Potential abstraction (S) = 1.64 1.15
Runoff depth (Q) in inches 8.95 9.44
Runoff volume (ac-ft) = 8.36 8.81

Volume difference = 0.45 ac-ft

The estimated attenuation volume is this volume difference of 0.45 ac-ft.

Chapter 9: Stormwater Management Facility 9-57


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

9.4.2.2 Simple Pond Model Procedure


Another technique for estimating attenuation volume is to route a design storm through a
simple pond model. It works best with a routing program that allows a rating curve for the
stage-discharge relationship and a stage-storage (not area) relationship for the pond
configuration. The model should be set up as follows:

 Arbitrarily select pond bottom and top elevations.

 Use two points for the stage-discharge relationship:


(1) Zero discharge @ pond bottom, and (2) Allowable discharge @ pond top

 Use two points for the stage-storage relationship:


(1) Zero storage @ pond bottom, and (2) Estimated storage @ pond top

As with any routing, this is an iterative process. During each iteration, the estimated
storage volume is changed to bring the routed peak stage close to the top of the pond.
The storage volume that causes the peak pond stage to match the top of the pond is the
estimated attenuation storage.

This approach is useful when the discharge rate is limited to something other than the
pre-developed rate. It is complicated when working with the Department’s multiple design
storms. Which design storm do you route? The following suggestions will help to simplify
working with the multiple design storms:

 Determine the pre-developed discharges for the 100-year, 1-hour design storm
through the 100-year, 8-hour design storm. Use the smallest of these calculations as
the allowable discharge rate. For the Storm for Storm Approach to critical duration,
the post-developed discharge rate will be limited to all of the corresponding pre-
developed rates, so using the rate for estimating purposes is reasonable. The basis
for running only the 1-hour through the 8-hour design storm is that one of these design
storms usually is critical to sizing ponds discharging to open basins.

 Route the post-developed conditions using a “nested” design storm such as the NRCS
Type 2 Florida modified or the applicable WMD design storm. These distributions often
are as severe as or more severe than the Department’s distributions.

Chapter 9: Stormwater Management Facility 9-58


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Example 9.4-2: Estimating attenuation volume using a simple pond model

Given:
 The same conditions as in Example 9.4-1
 Pre-developed time of concentration = 29 min.
 Post-developed time of concentration = 21 min.

Find: The estimated attenuation volume

1. Pre-developed runoff:

Determine the pre-developed discharge rates for the 100-year FDOT 1-hour and
8-hour design storms. Using a typical program based on the NRCS unit
hydrograph approach, you should obtain values similar to these when using a
peak shape factor of 256. The rainfall volumes for Somewhere City, Florida, are
tabulated in Step 1 of Example 9.4-3.

Pre-Developed Peak Runoff Rates (cfs)


1-hour, 100-yr 2-hour, 100-yr 4-hour, 100-yr 8-hour, 100-yr
33.2 30.1 25.5 27.8

The discharge associated with the 4-hour, 100-year design storm is the smallest
and will be used as the allowable discharge.

2. Develop a simplified pond model as follows.


Elevation Discharge (cfs) Storage
Pond Bottom 0 0 0
Top of Pond 10 25.5 Trial and Error

3. Route a nested design storm through the pond using post-developed conditions.
For this example, we will route the 25-year, SFWMD 72-hour storm. Adjust the
storage as necessary to have the routed peak stage match the top of pond. After
numerous iterations, a storage value of 1.3 ac-ft was found acceptable, so:

The estimated attenuation volume is 1.3 ac-ft.

Chapter 9: Stormwater Management Facility 9-59


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

9.4.2.3 Other Techniques


FHWA’s HEC 22 provides several methods to estimate attenuation volume, including
examples and comparisons. Although most of these techniques are reasonably accurate,
they—like the previous method—are complicated when working with the Department’s
multiple design storms.

9.4.3 Tailwater Conditions


Tailwater conditions can affect the design of the outfall structure, the size of the pond,
and even the evaluation of alternate pond sites. The pond must meet the attenuation
requirements during the tailwater conditions expected to occur coincident with the design
storms. Predicting the tailwater condition sometimes can be difficult. Our facilities usually
discharge to points associated with watersheds that are much larger than the drainage
area of our facility. It may be appropriate to model the larger watershed and apply design
storms to both the road project and the larger watershed simultaneously. This method will
help to address any timing-related effects.

Tailwater conditions can become more challenging when discharging at or close to the
confluence of two streams, as shown in the figure below. Depending on the relative size
of each basin, it may be overly conservative to use the combined (or coincident) 100-year
probability for each stream. National Cooperative Highway Research Program (NCHRP)
conducted Project 15-36: Estimating Joint Probabilities of Design Coincident Flows at
Stream Confluences (http://www.trb.org/Publications/Blurbs/169456.aspx) to develop
practical procedures for estimating joint probabilities of coincident flows at stream
confluences. This paper focuses on two practical design methods and provides a step-
by-step application guide for designers in Appendix H of the document.

Pond Site
Large River

A simpler approach is to estimate the worst-case tailwater condition and see if it


submerges the control point of the outlet control structure. If it does not, the tailwater
condition can be ignored in the design of the weir/orifice of the outlet control structure.

Placing a pond in a 100-year riverine floodplain can complicate the design due to high
tailwater conditions that may be coincident with the design storm. Other complications

Chapter 9: Stormwater Management Facility 9-60


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

such as flood plain compensation and changes to floodway conveyances may exist as
well. Chapter 5 (Bridge Hydraulics) addresses impacts to floodway conveyances.

9.4.4 Routing Calculations


Most engineers currently use computer programs to route hydrographs through
stormwater facilities. The majority of computer programs use the storage indication
method for this process. HEC 22 contains a discussion of the storage indication method,
with an example. The Drainage Connection Handbook also discusses this method.

Although the computer reduces the effort, it does not eliminate the iterative process of
modifying the pond and outlet control structure after each run. To design an acceptable
pond and outlet control structure, you usually will run numerous iterations. You can adjust
six items to meet the discharge requirements: (1) weir width (or orifice size), (2) weir crest
(or orifice invert) elevation, (3) pond surface area, (4) pond depth, (5) pond length to width
ratio, and (6) outlet pipe size. Although some of these items may be constrained by
regulatory requirements, the following provides general guidance for making adjustments
during the iterative process.

If the only change made is: The results are:


Increasing weir width (or orifice
Increases discharge and lowers stage.
size)
Lowering weir crest (or orifice Increases discharge (volume more than rate) and lowers
invert)1 peak stage.
Increasing pond surface area Decreases discharge and lowers peak stage. For
(increases storage above and retention systems, increases infiltration and shortens
below weir crest) recovery time.
Lowering pond depth1 Decreases discharge and lowers peak stage. For
(increases storage below the retention systems, decreases infiltration and lengthens
weir only) recovery time when saturated groundwater flow
conditions exist.
Increasing length to width ratio Increases discharge and raises peak stage, due to slight
reduction in storage area for the same surface area. For
retention systems, increases infiltration and shortens
recovery time when saturated groundwater flow
conditions exist.
Decreasing outlet pipe size Increases discharge and raises peak stage due to
additional friction losses in the pipe. Increases outlet
velocity in discharge pipe.
1. Normally applicable to only retention systems.

Chapter 9: Stormwater Management Facility 9-61


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

9.4.5 Discharges to Watersheds with Positive Outlet (Open Basins)


Using Chapter 14-86
Using the Storm for Storm Approach, the Department’s criterion for discharges to open
basins requires that—for a given frequency—the post-developed discharge rate for each
duration must be less than or equal to the pre-developed discharge rate of corresponding
duration. Most of the regulatory agencies also have requirements for post-developed
discharge rates. You must meet these requirements and the Department’s criterion.

Example 9.4-3: Discharge to watershed with positive outlet (open basin)


This example uses information developed in Examples 9.1-1, 9.4-1, and 9.4-2.

Given:
The following information has been verified since the time of the pond site evaluation.

 SHWT elevation at pond site: = 56.1 ft Agreed to by regulatory agency


 Lowest ground elev. around pond site = 59.1 ft From design survey

Find:
The required pond configuration to meet the FDOT criterion. For this example, the pond
also will be designed to meet SWFWMD and SFWMD criteria.

1. Determine the location-specific rainfall volumes using the NOAA website link in
Section 1.4 of the Drainage Manual.
Rainfall Volumes: Somewhere City, Florida
2-yr 5-yr 10-yr 25-yr 50-yr 100-yr
1-hr 2.4 2.95 3.25 3.75 4.1 4.5
2-hr 2.8 3.5 3.9 4.5 5.0 5.5
4-hr 3.3 4.0 4.6 5.4 6.0 6.6
8-hr 3.8 4.9 5.6 6.5 7.3 8.0
1-day 4.8 6.3 7.7 8.7 9.7 10.7
3-day 6.1 7.9 9.1 10.8 12.2 14.1
7-day 7.5 9.4 11.5 13 14.8 16.8
10-day 8.5 11 13 15 17 19

Chapter 9: Stormwater Management Facility 9-62


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

First Round of Iterations


2. Determine the pre-developed runoff rates: This will establish the allowable
discharge rates.

Time of concentration = 29 min (from Ex. 9.4-2)


Pre-developed CN:
Roadway pavement: = 0.79 ac @ CN = 98 (from Ex. 9.4-1)
Pervious area: = 9.61 ac @ CN = 85 (from Ex. 9.4-1)
Proposed pond area: = 0.94 ac @ CN = 85
Total: = 11.3 ac @ CN = 85.9
The proposed pond size is from Example 9.1-1, Pond Siting Stage. This is a reasonable
assumption for the first iteration.

To simplify this problem, we have used the time of concentration, roadway pavement
area, and offsite land use from prior examples. Actually, you should use the latest
information from the design surveys and field reviews of the proposed project to establish
the pre-developed conditions. Using a typical program, which uses the NRCS unit
hydrograph approach, you should obtain values similar to these when using a peak shape
factor of 256. This peak shape factor is used throughout this example.

For the first round of iterations for a pond discharging to an open basin (with documented
flooding history), it is usually sufficient to run the 100-year FDOT 1-hour to 8-hour duration
design storms and the regulatory agency design storm.

Pre-Developed Runoff (cfs)


DOT 1-hr, DOT 2-hr, DOT 4-hr, DOT 8-hr, FLT2M, 25-
SF72, 25-yr
100-yr 100-yr 100-yr 100-yr yr
33.6 30.4 25.8 28.1 30.6 36.3

3. Post-developed runoff:

In urban sections, the time of concentration is best determined from the storm
sewer design tabulations. For this example, assume the storm sewer tabs have a
Tc = 21 min.

Time of concentration: = 21 min


Post-developed area & CN:
Roadway, curb, and sidewalk: = 3.24 ac @ CN = 98 (from Ex. 9.4.1)
Pervious area: = 7.17 ac @ CN = 85 (from Ex. 9.4.1)

Chapter 9: Stormwater Management Facility 9-63


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Pond: = 0.94 ac @ CN = 98
Total: = 11.3 ac @ CN = 89.7
4. Develop a stage-storage relation (pond configuration) for the first round of
iterations.

Dimensions at peak stage = 210.5 ft by 109 ft (from Ex. 9.1-1)

For the first iteration, use the configuration estimated in the pond siting evaluation unless you have
reasons not to.

Peak stage = 58.1 ft to maintain freeboard between ground line of 59 ft.

Although some WMDs allow treatment below SHWT, this example assumes that
treatment is above SHWT. Then, the pond length and width at SHWT elevation (for
routing purposes, the SHWT elevation is considered pond bottom) are:

Bottom length = Top length – 2 [side slope (peak stage – elevSHWT)]


= 210.5 m – 2 [5 ( 58.1 ft – 56.1 ft)] (1:5 side slopes)
= 191 ft

Similarly, Bottom width = 90 ft

Using these dimensions and side slopes, develop a stage-storage relationship. The
values below were obtained using the equation for the volume of a frustum of a pyramid.

Stage (ft) Storage (ac-ft)


56.10 0.00
56.50 0.16
56.90 0.34
57.30 0.52
57.70 0.72
58.10 0.92

5. Develop an outfall structure for the first round of iterations. Do so using the
maximum allowable stage and discharge. For this example, the maximum
allowable stage is the ground elevation minus the freeboard [59.1 ft – 1.0 ft =
58.1 feet]. The maximum allowable discharge is the largest pre-developed

Chapter 9: Stormwater Management Facility 9-64


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

discharge, which, for this example, is the SFWMD 72-hour, 25-year design storm
(see Step 2).

Weir crest elevation = 56.7 ft The treatment volume (10,950 ft3, given in Ex.
9.1-1) stacks 0.59 ft high.

Weir width (L) = Q  (C x H1.5) from Q = C x L x H1.5


= 36.3 cfs  (3.1 x 1.371.5)
The max head = 58.1 ft – 56.7 ft = 1.37 ft = 7.3 ft

For this example, we have assumed no tailwater effects. For your projects, you will
need to consider the effects of the tailwater conditions on the outfall control
structure.

During this round of iterations, ignore the effects of the water quality bleed down
orifice and start the routings at the top of the treatment volume.

6. Route the selected design storms. Using a typical routing program, you should
obtain values similar to the following.

Table 9.4-1
Discharge Peak Pond
Design Storm
(cfs) Stage (ft)
FDOT 1-hr, 100-yr
Pre 33.6
Post 38.2 58.1
Pond configuration: FDOT 2-hr, 100-yr
Pond dimensions at SHWT = 191 ft x Pre 30.4
90 ft Post 35.0 58.1
SHWT elev. = 56.1 ft FDOT 4-hr, 100-yr
Avg side slope = 1:5 Pre 25.8
Weir crest elev. = 56.7 ft Post 28.8 57.9
Weir width = 7.3 ft
FDOT 8-hr, 100-yr
Starting WS = 56.7 ft
Pre 28.1
Post 30.6 57.9
Allowable Stage = 58.1ft
SCS-T2FLM, 25-yr
Pre 30.6
Post 33.0 58.0
SFWMD 72-hr, 25-yr
Pre 36.3
Post 38.0 58.1

Chapter 9: Stormwater Management Facility 9-65


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

From this table, it appears that the 100-year, 1-hour or 2-hour may be critical because
they exceed the pre-developed discharge more than the others. Overall, the configuration
used in the first iteration is close to meeting the requirements. Shorten the weir length to
decrease the peak discharge. Doing so will cause the stage of the 1-hour, the 2-hour, and
the SFWMD design storm to exceed the allowable stage, so the pond size needs to be
increased also.

After making several runs, the stage-storage relationship shown below and a weir width
of 6.0 ft is close to meeting the requirements of the design storms modeled. The second
row in the table is the weir crest elevation sufficient to store the treatment volume.

Stage (ft) Storage (ac-ft)


56.10 0.00
56.40 0.27
56.50 0.36
56.90 0.73
57.30 1.11
57.70 1.52
58.10 1.94

Using this configuration, you should obtain values as shown below.


Table 9.4-2
Peak Pond
Discharge
Design Storm Stage
cfs
ft
FDOT 1-hr, 100-yr
Pond configuration: Pre 33.6
Pond dimensions at SHWT = Post 27.1 57.7
288.7 ft x 131.2 ft FDOT 2-hr, 100-yr
SHWT elev. = 56.1 ft Pre 30.4
Average side slope = 1:5 Post 27.3 57.7
Weir crest elev. = 56.4 ft FDOT 4-hr, 100-yr
Weir width = 6.0 ft Pre 25.8
Starting WS = 56.4 ft Post 26.4 57.7
FDOT 8-hr, 100-yr
Allowable Stage = 58.1 Pre 28.1
Post 27.4 57.7
SCS-T2FLM, 25-yr
Pre 30.6
Post 27.5 57.7
SFWMD, 72-hr, 25-yr
Pre 36.3
Post 30.7 57.8

Chapter 9: Stormwater Management Facility 9-66


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

From this table, it appears that the FDOT 4-hour is critical since it is the only duration for
which the post-developed discharge is not less than the pre-developed discharge. The
SFWMD design storm creates the highest stage of the storms modeled.

Second Round of Iterations


7. Adjust the drainage basin characteristics due to the pond size being increased in
the previous step. Remember that, for this example, the pond is outside the area
draining to the pond so increasing the pond size also increases the total area. See
Example 9.4-1. During the first iteration, we assumed the entire pond area had a
CN = 98. A more refined estimate of the pond area curve number can be made at
this time.
Pond Area:

Water surf dims at peak stage = 308 ft. x 150 ft.


Water surface area at peak stage = 1.06 ac @ CN = 98
Total pond area (incl maint berms) = 1.53 ac
Grassed area within total pond area = 0.47 ac @ CN = 85

Total Project Area and Curve Number:


Pre-developed CN:
Roadway pavement: = 0.79 ac @ CN = 98 (same as Step 2)
Pervious area: = 9.61 ac @ CN = 85 (same as Step 2)
Proposed pond area: = 1.53 ac @ CN = 85
Total: = 11.9 ac @ CN = 85.9

Post-developed CN:
Roadway, curb, and sidewalk: = 3.24 ac @ CN = 98 (same as Step 3)
Pervious area: = 7.64 ac @ CN = 85 [7.17 ac (Step 3) + 0.47
ac]
Pond: = 1.06 ac @ CN = 100
Total = 11.9 ac @ CN = 89.9

8. Calculate the pre-developed runoff and then route the design storms. For this
example, we will add the FDOT 24-hour, 100-year design storm at this time. The
results are shown in the following table.

Chapter 9: Stormwater Management Facility 9-67


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Table 9.4-3
Peak Pond
Discharge
Design Storm Stage
(cfs)
(ft)
FDOT 1-hr, 100-yr
Pond configuration: (Same as Pre 35.2
previous table) Post 28.7 57.8
Pond Dimensions at SHWT = FDOT 2-hr, 100-yr
288.7 ft x Pre 31.9
131.2 ft Post 28.9 57.8
SHWT El. =56.1ft FDOT 4-hr, 100-yr
Avg Side Slope = 1: 5 Pre 27.0
Weir Crest El. = 56.40 ft Post 27.9 57.7
Weir Width = 6.0 ft FDOT 8-hr, 100-yr
Starting WS = 56.4 ft Pre 29.5
Post 28.9 57.8
Allowable Stage = 58.1ft FDOT 24-hr, 100-yr
Pre 11.2
Post 11.1 57.2
SCS-T2FLM, 250-yr
Pre 32.1
Post 29.0 57.8
SFWMD, 72-hr, 25-yr
Pre 38.1
Post 32.5 57.9

From this table, we can see the discharge for the 4-hour needs to be reduced and
the stage of the SFWMD storm can still be increased, so the weir width can be
reduced. After several iterations, a weir 4.5 ft wide works. The results are as
follows.

Chapter 9: Stormwater Management Facility 9-68


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Table 9.4-4
Peak Pond
Design Storm Discharge Stage
(cfs) (ft)
FDOT 1-hr, 100-yr
Pond configuration: Pre 35.2
Pond dimensions at SHWT = Post 25.8 57.9
288.7 ft x 131.2 ft FDOT 2-hr, 100-yr
SHWT elev. = 56.1 ft Pre 31.9
Avg. side slope = 1:5 Post 26.7 57.9
Weir crest elev. = 56.4 ft FDOT 4-hr, 100-yr
Weir width = 4.5 ft Pre 27.0
Starting WS = 56.4 ft Post 26.8 57.9
FDOT 8-hr, 100-yr
Allowable stage = 58.1 Pre 29.5
Post 27.6 58.0
FDOT 24-hr, 100-yr
Pre 11.2
Post 10.9 57.3
SCS-T2FLM, 25-yr
Pre 32.1
Post 27.5 58.0
SFWMD, 72-hr, 25-yr
Pre 38.1
Post 30.3 58.0

Since this configuration meets the requirements for these design storms, the pond
size is probably adequate. We need to make sure that the discharges are not
exceeded for the less frequent (2-year through 50-year) DOT design storms. We
also will check the longer-duration storms, though it appears that the long duration
storms (24-hour to 240-hour) will not control the size of the pond, since the stages
and discharges of the 24-hour are much less than the 1-hour through 8-hour
duration storms.

9. Check the size of the bleed down orifice. For this example, you will need a 1.5-
inch diameter orifice or less to meet the typical wet detention criteria [discharge no
more than half of the treatment volume in 60 hours and discharge the total
treatment volume in no less than 120 hours]. At maximum pond stage, the
discharge through this orifice is less than 0.1 cfs. This is insignificant for this
problem. The orifice flow will be ignored and the routing calculations will be started
at the weir crest, as done in previous iterations.

Chapter 9: Stormwater Management Facility 9-69


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

If the discharge through the bleed down orifice at peak stage is small, ignore it. If not, model
the orifice in the routing. If the orifice is modeled, the starting water surface should reflect
some amount of drawdown. The average inter-event period between storms is 72 hours.
Most wet detention systems hold at least half of the treatment volume for 60 hours.
Therefore, for most wet ponds, starting the water surface at an elevation associated with
half of the treatment volume would be reasonable. If the regulatory requirements allow for a
quicker drawdown, it may be reasonable to start the water surface at the bleed down orifice.

10. Run the other design storms. The other design storms were routed through the
above pond configuration and all the post-developed rates were less than the pre-
developed rates, except one. A summary of these is shown below.
Table 9.4-5 (Example 9.4-3)
Pond Config. 100-yr 50-yr 25-yr 10-yr 5-yr 2-yr
as in Table Discharge Discharge Discharge Discharge Discharge Discharge
5.3-4 (cfs) (cfs) (cfs) (cfs) (cfs) (cfs)
1-hr
Pre 35.2 31.0 27.4 22.3 19.3 14.0
Post 25.8 22.3 19.4 15.3 13.0 9.1
2-hr
Pre 31.9 28.0 24.2 19.8 16.9 11.9
Post 26.7 23.3 20.0 16.1 13.7 9.6
4-hr
Pre 27.0 24.0 21.1 17.1 14.2 10.8
Post 26.8 23.7 20.7 16.8 13.8 10.5
8-hr
Pre 29.5 26.5 23.0 19.0 16.0 11.3
Post 27.6 24.5 21.1 17.3 14.3 9.9
24-hr
Pre 11.2 10.0 8.9 7.7 6.0 4.3
Post 10.9 9.7 8.6 7.4 5.8 4.1
3-day
Pre 8.2 7.1 6.2 5.2 4.5 3.4
Post 8.2 7.1 6.2 5.2 4.4 3.3
7-day
Pre 5.9 5.2 4.5 4.0 3.2 2.5
Post 5.9 5.2 4.5 4.0 3.2 2.6
10-day
Pre 7.8 6.9 6.1 5.3 4.4 3.4
Post 7.8 6.9 6.1 5.3 4.4 3.4

The 7-day, 2-year post-developed discharge rate is greater than the pre-developed
rate. If carried to three significant digits, the increase is 0.02 cfs (2.56-2.54). This
is within the accuracy of these calculations and would be acceptable for most
projects. If you or your project reviewers are concerned about an increase like this,

Chapter 9: Stormwater Management Facility 9-70


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

you could modify the weir configuration slightly, as is done in Step 8 of Example
9.4-4.

11. Fine tune pond dimensions.

The stage-storage values used in this example are based on length and width
dimensions applied to a frustum of a pyramid. When you apply the radii to the
corners, you would reduce the storage using the same pond dimensions, so use
an equivalent stage-area relationship when working with the contours within the
CADD file. Doing so also will allow you to configure the pond for aesthetic
purposes while maintaining the necessary stage-storage relationship.

9.4.6 Discharges to Watersheds without Positive Outlet (Closed


Basins) Using Chapter 14-86
Using the Storm for Storm Approach, the Department’s criterion for projects discharging
to a closed basin is that—for a given frequency—the post-developed discharge (rate and
volume) for each duration must be less than or equal to the pre-developed discharge (rate
and volume) of corresponding duration.

Ensure the retention volume is large enough that the post-developed discharge volumes
do not exceed the pre-developed discharge volumes. The retention volume is the volume
between the pond bottom and lowest discharge elevation of outlet control structure.

When using the NRCS runoff methodology, you can conservatively calculate the retention
volume as the difference between the pre-developed and post-developed discharge
volume for the 100-year, 10-day event. Some of this volume is infiltrated into the soil
during the storm, so the actual retention volume is sometimes less than this. During long-
duration design storms, such as the 3-day through 10-day durations, the volume infiltrated
during the storm can be substantial. You can account for this by using a program that
models the infiltration while routing the storm hydrograph. When you do this, you will not
know the required retention volume until you have routed the storms and know how much
volume infiltrates during the storm event.

The retention volume must recover at a rate such that half of the volume is available in
seven days and the total volume is available in 30 days. When measuring the recovered
volume, the pond is instantly (or over a very short time) filled with a runoff volume equal
to the retention volume. Then, the water can infiltrate with no inflow to the pond.

Chapter 9: Stormwater Management Facility 9-71


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

9.4.6.1 Retention System Groundwater Flow Analysis


The approach described below is based on the current approach to modeling the recovery
of the treatment volume in retention systems. You can apply the same techniques to the
infiltration of retention systems discharging to closed basins.

The next several pages summarize the critical information contained in the following
documents. We suggest that you read these documents before designing a retention
system.

a) Stormwater Retention Pond Infiltration Analyses in Unconfined Aquifers.


Prepared by Jammal and Associates, Inc., 1989 (Revised 1991), for the
SWFWMD, Brooksville, Florida. See web link below:
http://publicfiles.dep.state.fl.us/dwrm/stormwater/stormwater_rule_development/docs/retpond_infil_analys.pdf

b) Full-Scale Hydrologic Monitoring of Stormwater Retention Ponds and


Recommended Hydro-Geotechnical Design Methodologies. Prepared by
PSI, Jammal and Associates Division, for the SJRWMD, August 1993,
Special Publication SJ93-SP10. See web link below:

http://www.sjrwmd.com/technicalreports/pdfs/SP/SJ93-SP10.pdf

During a storm event, runoff from the drainage basin enters the pond and infiltrates the
pond bottom. At the beginning of a storm, the groundwater beneath the pond moves
primarily vertically downward through unsaturated soil. If runoff to the pond exceeds the
infiltration, the water depth in the pond increases as the wetting front continues to move
down. Although the soil between the wetting front and the pond bottom is wet, it is not
totally saturated due to entrapped air. After the wetting front reaches the water table, the
vertical infiltration adds water to the water table aquifer. At this time, the groundwater
moves primarily horizontally within the saturated aquifer while the water table begins to
mound and saturate the soil beneath the pond. If infiltration continues, the mound rises
to and above the pond bottom. When the mound reaches the pond bottom, the area
beneath the bottom is fully saturated and flow moves primarily horizontally. See Figure
9.4-1.

Chapter 9: Stormwater Management Facility 9-72


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Figure 9.4-1: Groundwater Flow Characteristics during Infiltration

Chapter 9: Stormwater Management Facility 9-73


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Determining the drawdown characteristics and the recovery time may involve modeling
the downward vertical flow through unsaturated soil, or the horizontal saturated flow of
the groundwater mound, or both.

(A) Unsaturated Flow

The design infiltration rate is: K VU


ID =
FS
f
The time necessary to saturate the soil below the pond is: T = HB
ID

The source for the above equations is the modified Green and Ampt infiltration equation.
Their derivation is presented in Stormwater Retention Pond Infiltration Analyses in
Unconfined Aquifers (Jammal and Associates, 1991).

The total volume of water required to saturate the voids in the soil below the pond
bottom is: VOLVOIDS = APB  HB  f

where:
HB = Height of pond bottom above groundwater (see Figure 9.4-1)
ID = Design infiltration rate
KVU = Unsaturated vertical hydraulic conductivity
You can obtain this typically from a Double Ring Infiltration (DRI) test.
Although infiltration is occurring during the test, the soil is not fully
saturated due to entrapped air. The unsaturated K is less than the
saturated K. Unsaturated K ranges from one-half to two-thirds saturated K
(Bouwer 1978, ASTM D 5126, & Jammal and Assoc., 1991).
f= Fillable porosity. See description in following pages.
APB = Area of pond bottom
FS = Factor of safety, usually 2.0.

You can use this factor of safety to account for the variability of the measurements and
for the sediment that inevitably will enter the pond and clog the bottom surface.

(B) Saturated Flow


In most areas of the state, except the high sandy ridges, the groundwater mound likely
will rise to the pond bottom, forcing the groundwater into a saturated horizontal flow. The
most common approach to analyzing saturated flow conditions is to assume flow to be
purely horizontal and uniformly distributed across the thickness of the receiving aquifer.

Chapter 9: Stormwater Management Facility 9-74


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Model the aquifer as having a single homogeneous layer of uniform thickness and a
horizontal initial water table.

Several computer models are available to analyze saturated flow. Most use a form of the
USGS program MODFLOW. A simplified approach was developed by Jammal and is
discussed in the SJRWMD’s, “Applicant’s Handbook for Regulation of Stormwater
Management Systems.” Regardless of which program or technique is used, four
parameters are needed to model saturated flow: (1) the thickness of the aquifer, (2) the
groundwater table elevation, (3) the fillable porosity, and (4) the horizontal saturated
hydraulic conductivity of the aquifer.

 Thickness (or Elevation of the Base) of Mobilized Aquifer:

This is the thickness of soil through which the horizontal flow will occur. You usually
will measure this depth to the top of a confining or very dense layer, such as
hardpan, that will restrict the downward vertical movement of groundwater. Use
the lesser of the depth of the soil boring or the width of the pond as the maximum
value in the analysis. (The maximum depth of the mobilized aquifer is about equal
to the width of the pond. [Bouwer, 1978]).

 Groundwater Elevation:

For modeling the recovery of the treatment volume, you usually will use the SHWT
elevation. For modeling the infiltration of a pond discharging to a closed basin, this
groundwater elevation should represent the groundwater elevation during an
extreme event like the 100-year, 10-day design storm. Currently, there is no
standard procedure for determining this elevation; nevertheless, it could be
substantially higher than the SHWT. For example, where the pond is located near
the low in the watershed (lake or flood plain at the low), it may be reasonable to
use the 100-year lake or floodplain elevation as the extreme event groundwater
elevation. Where the pond is located higher in the watershed, the extreme event
groundwater elevation may be closer to the SHWT. Use your judgment and handle
these situations on a case-by-case basis.

 Fillable Porosity:

This is sometimes called specific yield, storage coefficient, effective storage


coefficient, or effective porosity. It is the difference between volumetric water
content of soil before and after wetting. The total porosity of a soil is the percentage
of the total volume of the material occupied by pores or interstices. The fillable
porosity is less than the total porosity because some water exists in soils above
the water table; therefore, not all of the unsaturated void space is available for
filling. In the zone immediately above the groundwater, capillary rise causes the

Chapter 9: Stormwater Management Facility 9-75


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

voids to be substantially filled with water. In fine sands, the distance saturated due
to capillary rise is roughly six inches. Therefore, the fillable porosity varies with the
depth to the water table.

Specific field or laboratory testing usually is not required for determining the fillable
porosity. For most calculations associated with fine sands, the fillable porosity will
vary from 0.1 to 0.3 (10 percent to 30 percent). The SFWMD has produced soil
storage curves that you can use to estimate the fillable porosity. For fine sand
aquifers, the SJRWMD recommends using a fillable porosity in the range of 20
percent to 30 percent in infiltration calculations. The higher values of fillable
porosity will apply to the well-to-excessively-drained, hydrologic group “A” fine
sands, which generally are deep and contain less than 5 percent by weight passing
the No. 200 sieve.

With all other dimensional and aquifer factors being the same, the predicted
recovery time decreases as the assumed value of fillable porosity increases.
Increasing the fillable porosity from 0.2 (20 percent) to 0.3 (30 percent) decreases
the recovery time by 15 percent to 30 percent.

 Horizontal Saturated Hydraulic Conductivity of the Aquifer:

Since you assume horizontal flow for the saturated analysis, the hydraulic
conductivity should represent that direction. This should represent the weighted
value of the soil above the confining layer. There are numerous techniques for
measuring this value, and they are briefly described below. The Department
recommends applying a safety factor of 1.5 to 2.0 to the measured values. You
can apply this factor of safety to account for the variability in the elevation of the
impermeable layer, measurement of the conductivity, and the estimate of fillable
porosity.

Cased hole tests:


Generally, measure horizontal hydraulic conductivity if the casing bottom is below
the water table during the test. Generally, measure vertical hydraulic conductivity
if the casing bottom is above the water table during the test. Use the results with
caution if the bottom of the casing is near an impermeable or confining layer.

Uncased hole tests:


This also applies to cased holes that use screen, perforated pipe, or rock bottom
to maintain borehole shape. These generally measure horizontal hydraulic
conductivity K.

Chapter 9: Stormwater Management Facility 9-76


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Double Ring Infiltration (DRI) test:

Generally, the DRI measures the vertical unsaturated hydraulic conductivity.


Although the Department does not encourage the use of the DRI to obtain the
weighted horizontal saturated hydraulic conductivity, if it is the only test information
you have, the saturated horizontal hydraulic conductivity could be estimated by
applying two adjustment factors as follows:

KVS = 1.5 KVU

KHS = 1.5 KVS (conservative SWFWMD guideline)


or
KHS = 1.5 x 1.5 KVU

Pumping tests:
These tests generally are expensive and should be reserved for highly sensitive
projects. They can overestimate hydraulic conductivity if the bore holes extend into
a highly permeable layer which is below a confining layer and the proposed pond
bottom is above the confining layer.

(C) Special Saturated Analysis


If you cannot model the aquifer as having the characteristics discussed above, you may
need to use a more complicated, fully three-dimensional model with multiple layers, such
as MODFLOW.

(D) Coordination with the Geotechnical Engineer


When requesting the soils investigation, provide the Geotechnical Engineer with the
following information:

 Pond location

 Approximate pond shape (length, width, plan area configuration)

 Estimated pond bottom elevation

 Your estimate of SHWT

 The ideal functional characteristics of the pond, such as: “This pond will be
designed to retain a volume of stormwater for flood control purposes. It should

Chapter 9: Stormwater Management Facility 9-77


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

infiltrate half the retention volume in no more than seven days and all of the volume
in no more than 30 days.”

 The anticipated groundwater flow conditions/analysis you expect to model based


on your preliminary review of the soil and groundwater conditions.

The Geotechnical Engineer needs to know the information listed above because the soils
investigation can vary depending on the groundwater flow condition anticipated during
your design conditions. Refer to Table 9.4-6

Table 9.4-6 – Typical Soil Investigations

Anticipated Groundwater
Soil Investigation1
Flow Conditions/Analysis
1) Thickness of mobilized aquifer.
2) Determine SHWT elevation.
Saturated
3) Determine weighted saturated horizontal
hydraulic conductivity of mobilized aquifer.
1) Obtain unsaturated vertical hydraulic
conductivity at or near pond bottom.
Unsaturated
2) Determine SHWT or confirm that SHWT is at
(Probably limited to high,
least as low as drainage engineer estimated.
sandy ridges)
3) Confirm that no confining layer exists between
pond bottom and SHWT.
Karst Areas See discussion in this section.
1 Preliminary results of the soil investigation may dictate that a different soil
investigation is necessary. For example, you may have estimated sandy
conditions down to a deep water table, planned on doing an unsaturated
analysis, and requested appropriate soil information. Then the initial soil borings
could indicate a confining layer close enough to the pond bottom to warrant a
saturated analysis.

If the groundwater elevation is within two feet of the pond bottom, you can assume that
horizontal saturated flow will occur. If the groundwater is farther from the pond bottom,
you should compare the volume of the voids under the pond to the volume of runoff that
must be infiltrated.

For estimating the groundwater flow conditions, the volume to be infiltrated should be the
treatment volume for retention systems discharging to open basins with known historical
flooding. It should be the difference between the 100-year, 10-day runoff volume for
ponds discharging to closed basins. If the volume to be infiltrated is larger than the volume
of the voids under the pond, the groundwater mound will rise to the pond bottom, thus
forcing saturated horizontal flow.

Chapter 9: Stormwater Management Facility 9-78


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Karst Areas:

The WMDs and FDEP have identified known Karst areas and usually have special
requirements for stormwater facilities in these areas to assure that water quality of the
aquifer is maintained. Sink holes can present problems during or after construction, so it
is important that you are aware of potential sink hole locations.

Some sink holes can be only a meter or two in diameter, thus making it difficult to identify
their potential as a hazard. Sometimes potential sink holes can be identified in the field
by localized depressions in the ground surface. You may find it useful to try ground
penetrating radar in some situations, but this tool has a disadvantage in that it does not
penetrate clay layers. Work closely with the Geotechnical Engineer to identify potential
sink holes.

As a preventive measure, you could place a permeable geotextile strong enough to span
a small opening several feet below the pond bottom. This would allow small sink holes to
develop without requiring any maintenance work. Doing this will add substantial costs and
may not be warranted for all facilities in Karst areas. You and the Geotechnical Engineer
should make a joint decision to follow this approach.

Example 9.4-4: Discharge to watershed without positive outlet (closed basin)

Given:
 Pre-developed roadway pavement = 2 10-foot lanes
 Drainage area: includes roadway right of way and
offsite draining to the roadway: =13.0 ac
 Offsite land use = Residential lots averaging 1/2 ac
 Proposed typical section = 4-lane urban section
 Combined roadway, curb, and
sidewalk width = 73 ft
 Length of roadway within drainage
area = 2,313 ft
 Treatment volume = 17,600 ft3
 Maximum allowable pond stage = 104 ft
 Offsite runoff draining to the project will be taken through the pond, not bypassed
around.

Chapter 9: Stormwater Management Facility 9-79


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

 Project located near Somewhere City, Florida; rolling terrain, approx. 2 percent
grades, Hydrologic Soil Group B
 A confining or impermeable layer exists at approximately elevation 92 ft
 The saturated horizontal hydraulic conductivity was estimated to be 8 ft/day
 The SHWT was estimated at approximately elevation 93 ft
Find: Pond size and outlet control structure configuration.

1. Pre-developed runoff:
Time of concentration = 21 min (given)
Area & curve number:
Roadway pavement: = 1.06 ac @ CN = 98 (20 ft x 2,313 ft)
Pervious area: = 11.94 ac @ CN = 70 (13 ac – 1.06 ac)
Proposed pond area: = 1.50 ac @ CN = 70 (preliminary size)
Total: = 14.5 ac @ CN = 72.1
As in Example 9.4-1, the proposed pond is outside the area draining to the
roadway; thus, the pond area must be added to the other areas.

Also, as in Example 9.4-1, the roadway right of way to be acquired is within the
area draining to the roadway. For your project, the acquired right of way may be
outside the area draining to the road, thereby requiring that the additional right of
way be added to the other areas.

2. Post-developed runoff:
Time of concentration = 16 min. (given)
Area and curve number:
Roadway, curb, and
sidewalk: = 3.88 ac @ CN = 98 (72.8 ft x 2,313 ft)
Pervious area: = 9.12 ac @ CN = 70 (13 ac – 3.88 ac)
Pond: = 1.50 ac @ CN = 98
Total: = 14.5 ac @ CN = 80.4

Chapter 9: Stormwater Management Facility 9-80


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

3. Determine the location-specific rainfall volumes using the NOAA website link in
Section 1.4 of the Drainage Manual.
Rainfall Volumes (inches): Somewhere City, FL
2-yr 5-yr 10-yr 25-yr 50-yr 100-yr
1-hr 2.2 2.7 3.0 3.5 3.8 4.1
2-hr 2.7 3.3 3.7 4.2 4.6 5.1
4-hr 3.1 3.9 4.4 5.0 5.6 6.1
8-hr 3.6 4.6 5.1 5.9 6.6 7.3
1-day 4.4 5.8 6.8 7.8 8.7 9.6
3-day 5.6 7.2 8.3 9.9 11 12.4
7-day 7.0 8.9 10 12 13.4 15
10-day 7.6 9.5 11.2 13.7 15.2 16

For this example, we will use peak shape factor = 323 for all NRCS hydrograph runs.

4. Assumptions:
a) Unsaturated vertical hydraulic conductivity: A DRI could not be performed
because of the depth of the pond bottom. The unsaturated vertical hydraulic
conductivity was estimated from the saturated horizontal conductivity (KHS = 8
ft/day)
8 ft/day  (1.5 – 1.5) = 3.6 ft/day (see discussion of DRI)
A factor of safety of 2 was applied to both values; thus, the modeled values
are KHS = 4 ft/day, and KVU = 1.8 ft/day
b) Groundwater elevation: The extreme event groundwater elevation is assumed
to be 3 feet above the SHWT. Then, extreme event groundwater elevation =
96.0 feet.
c) Fillable porosity is assumed = 0.1 (10 percent), worst case for fine sands

First Round of Iterations


5. Develop a starting-size pond.
You can take any approach to develop the starting trial size for the pond.
Perhaps you found a preliminary estimate in the pond siting stage, or you can
make an educated guess or a guess based on experience from a similar project.
The following approach could be used:

 Assume the retention volume will be the difference in runoff volume for the
100-year, 10-day design storm. Using the approach of Example 9.4-1, the
volume difference is 66,588 ft3 for this example.

Chapter 9: Stormwater Management Facility 9-81


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

 Assume a height of the peak stage over the weir crest. For this example,
we will use 1 foot. With a peak pond stage of 104 feet, this puts the weir
crest at approximately 103 feet.

 Assume a pond bottom elevation, staying several feet above the


estimated extreme event groundwater elevation. For this example, we will
start 4 feet above the groundwater elevation with a pond bottom of 100
feet maintaining 4 feet between estimated peak groundwater and pond
bottom.

 Determine a pond size and shape that will fit the retention volume
between the pond bottom and the weir crest. For this example, a pond
with a 200 foot x 100 foot bottom and 1:4 side slopes meets these
constraints and will be used as a starting size.

6. Calculate the pre-developed discharge rates and volumes, and route the post-
developed runoff through the pond. The weir width was arbitrarily selected for this
iteration. Using a typical routing program that models infiltration during the storm,
you should obtain values similar to the following.

For the first round of iterations for a pond discharging to a closed basin, it is usually
sufficient to run the 100-year, FDOT 8-hour through 10-day duration storms.

Chapter 9: Stormwater Management Facility 9-82


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Table 9.4-6
Disch.
Disch. Peak Pond
Volume
Design Storm Rate Stage
(ft3 x
(cfs) (ft)
Pond Configuration: 103)
Pond bottom dimensions = 200 ft x Pre
100 ft 216 29.1
FDOT 8-hr, 100-year Pos 104.2
Pond bottom elevation = 100 ft 182 27.3
t
Avg side slope = 1: 4 Pre
Weir crest elevation = 102.9 ft FDOT 24-hr, 100- 323 10.4
Pos 103.7
Weir width = 5 ft year 289 10.9
t
Volume below weir crest = 68,585 ft 3 Pre
Allowable stage = 104 ft FDOT 3-day, 100- 459 8.2
Pos 103.6
year 422 8.4
t
Modeled Soil Conditions: Pre
Aquifer base elevation = 92 ft FDOT 7-day, 100- 589 6.1
Pos 103.5
Saturated horizontal condition (K HS) year 543 6.2
t
= 4 ft/day Pre
Water table elevation = 96 ft FDOT 10-day, 100- 639 7.5
Pos 103.5
Fillable porosity = 0.1(10%) year 589 7.7
t
Unsat Vert Cond. (K VU) = 1.8 ft/day
Quantity Control Retention Volume
7-Day 30-Day
Total recovered in 28 days (ft3) (ft3)
Vol. Reqd. to be Recovered = 33,300 66,590
Vol. Recovered (infiltrated) = 30,600 53,400

All the post-developed discharge volumes are substantially less than the pre-
developed discharge volumes of corresponding duration, so the pond retains more
volume than needed. That is, the post-developed discharge volumes could be
increased. This is done by lowering the weir. Although most of the post-developed
discharge rates exceed the pre-developed rates, they are close to the pre-
developed rates. To maintain similar post-developed rates, we will need to reduce
the weir width as it is lowered. After making several iterations of weir adjustments,
the following configuration produces the results in the following table.

For this example, we will add the 1-hour, 2-hour, and 4-hour duration storms.

Chapter 9: Stormwater Management Facility 9-83


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Table 9.4-7
Peak
Disch. Disch.
Pond
Design Storm Volume Rate
Stage
(ft3 x 103) (cfs)
Pond Configuration: (ft)
Pond bottom dimension = 200 ft x Pre 83.3 33.8
100 ft FDOT 1-hr, 100-year 103.5
Post 64.3 16.8
Pond bottom elevation = 100 ft Pre 122 31.1
Avg side slope = 1:4 FDOT 2-hr, 100-year 103.7
Post 109 19.9
Weir crest elevation = 101.5 ft Pre 164 25.4
Weir width = 1.5 ft FDOT 4-hr, 100-year 103.9
Post 155 22.6
Volume below weir crest = Pre 217 29.1
32,768 ft 3 FDOT 8-hr, 100-year 104.0
Post 212 24.0
Allowable stage = 104 ft FDOT 24-hr, 100- Pre 323 10.5
103.0
year Post 322 10.2
Modeled Soil Conditions:
FDOT 3-day, 100- Pre 459 8.2
Aquifer base elevation = 92 ft 102.8
year Post 458 8.3
Saturated horizontal cond.(KHS) =
FDOT 7-day, 100- Pre 588 6.1
4 ft/day 102.6
year Post 581 6.2
Water table elevation = 96 ft
Fillable porosity = 0.1(10%) FDOT 10-day, 100- Pre 638 7.6
102.8
Unsat vertical cond. (KVU) = 1.8 year Post 631 7.7
ft/day Retention Volume
7-Day 30-Day
Total recovered in 17 days
(ft3) (ft3)
Vol. Reqd. to be Recovered = 16,390 32,770
Vol. Recovered (infiltrated) = 24,200 32,770

This pond configuration meets the drawdown and discharge volume requirements. The
rate requirements are close to being met as the 3-day through 10-day storms at only 0.1
cfs above the pre-developed discharge rates.

Second Round of Iterations

7. Adjust the drainage basin characteristics due to the pond size being smaller than
estimated in Step 1. Remember that, for this example, the pond is located
outside the area draining to the road, so changing the pond size also changes
the total area. In Step 2, we assumed the entire pond area had a CN = 98. A
more-refined estimate of the pond area curve number can be made at this time.

Pond Area:

Water surface dimension at peak stage = 232 ft x 132 ft


Water surface area at peak stage = 0.70 ac
Total pond area (including maintenance berms & slopes) = 1.1 ac
Grassed area within total pond area = 0.40 ac

Chapter 9: Stormwater Management Facility 9-84


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Total Project Area and CN:

Pre-developed area & curve number:


Roadway pavement: = 1.06 ac @ CN = 98 (from Step 1)
Pervious area: = 11.94 ac @ CN = 70 (from Step 1)
Proposed pond area: = 1.11 ac @ CN = 70
Total: = 14.1 ac @ CN = 72.1

Post-developed area and curve number:


Roadway, curb, and sidewalk: = 3.88 ac @ CN = 98 (from Step 2)
Pervious area: = 9.54 ac @ CN = 70 (9.12 ac (from Step 2) +
0.42 ac)
Pond: = 0.70 ac @ CN = 100
Total: = 14.1 ac @ CN = 79.2
8. Calculate the pre-developed discharge rates and volumes and route the post-
developed runoff through the pond. Using the same pond/weir configuration as in
the previous table produces the following results.
Table 9.4-8
Disch.
Disch. Peak Pond
Volume
Design Storm Rate Stage
(ft x
3
cfs ft
Pond Configuration: 10 )3

Pond bottom dimensions = 200 ft x Pre 81.2 32.9


100 ft FDOT 1-hr, 100-year 103.3
Post 56.0 14.5
Pond bottom elevation = 100 ft Pre 119 30.3
Avg side slope = 1: 4 FDOT 2-hr, 100-year 103.6
Post 98.7 17.9
Weir crest elevation = 101.5 ft Pre 160 24.7
Weir width = 1.5 ft FDOT 4-hr, 100-year 103.8
Post 142 21.1
Volume below weir crest = 32,768 Pre 211 28.4
ft 3 FDOT 8-hr, 100-year 103.9
Post 197 22.1
Pre 315 10.1
Allowable stage = 104 ft FDOT 24-hr, 100-year 103.0
Post 303 9.7
FDOT 3-day, 100- Pre 447 7.9
Modeled Soil Conditions: 102.8
year Post 435 8.0
Aquifer base elevation = 92 ft
FDOT 7-day, 100- Pre 572 5.9
Saturated horizontal cond.(KHS) = 4 102.6
year Post 554 6.0
ft/day
Water table elevation = 96 ft FDOT 10-day, 100- Pre 621 7.3
102.8
Fillable porosity = 0.1(10%) year Post 602 7.4
Unsaturated vertical cond. (KVU) = Retention Volume
0.8 ft/day 7-Day 30-Day
Total recovered in 17 days
(ft3) (ft3)
Vol. Reqd. to be Recovered = 16,390 32,770
Vol. Recovered (infiltrated) = 24,,200 32,770

Chapter 9: Stormwater Management Facility 9-85


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

This essentially meets all the requirements. The 24-hour and 3-day are critical durations
for discharge volume. The 8-hour duration creates the highest stage. The 3-day through
7-day are critical durations for discharge rate and they exceed the pre-developed
discharge rates by less than 2 percent. This may be acceptable. For this example, several
more iterations could be made to bring these rates down without increasing the pond size.

Notice that the retention volume recovered in 7 days was more than necessary and the
total volume was recovered in only 17 days. This indicates that we can lower the pond
bottom. We can lower the weir crest the same amount that the pond bottom is lowered
and maintain similar discharge volumes, which we need to do. As we lower the weir crest,
we can reduce the weir width to reduce the discharge rate, which is the primary intent. So
after several iterations, the following configuration using two weirs seems to do the trick.
Notice it involves a compound weir.

Table 9.4-9
Peak
Disch. Disch.
Pond
Design Storm Volume Rate
Pond Configuration: Stage
(ft3 x 103) (cfs)
Pond bottom dimension: = 192 ft x (ft)
92 ft Pre 81.2 32.9
FDOT 1-hr, 100-year 103.0
Post 38.9 8.2
Pond bottom elevation = 99 ft Pre 119 30.3
Avg side slope = 1:4 103.5
FDOT 2-hr, 100-year Post 78.5 13.8
#1 weir crest elevation = 100.5 ft Pre 160 24.7
#1 weir width = 0.5 ft FDOT 4-hr, 100-year 103.7
Post 124 21.1
Volume below #1 weir crest = Pre 211 28.4
29,120 ft3 FDOT 8-hr, 100-year 103.7
Post 185 21.7
#2 weir crest elevation = 103.3 ft FDOT 24-hr, 100- Pre 315 10.1
#2 weir width = 12 ftAllowable stage year 103.0
Post 299 8.1
= 104 ft
FDOT 3-day, 100- Pre 447 7.9
102.9
year Post 436 7.4
Modeled Soil Conditions:
FDOT 7-day, 100- Pre 572 5.9
Aquifer base elevation = 92 ft 102.6
year Post 556 5.9
Saturated horizontal cond. (KHS) = 4
ft/day FDOT 10-day, 100- Pre 621 7.3
102.9
Water table elevation = 96 ft year Post 610 7.3
Fillable porosity = 0.1(10%) Quantity Control Retention Volume
Unsat vertical cond. (KVU) = 1.8 7-Day 30-Day
Total recovered in 28 days
ft/day (ft3) (ft3)
Vol. Reqd. to be Recovered = 14,560 29,120
Vol. Recovered (infiltrated) = 17,590 29,120

Chapter 9: Stormwater Management Facility 9-86


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

This configuration meets all the requirements for the storms modeled. The 7-day and 10-
day durations are critical for discharge rate. The 3-day and 10-day durations are critical
for discharge volume, and the 4-hour and 8-hour durations create the highest stage. The
total retention volume is recovered in 28 days, just under the 30-day requirement.
Although it appears that the pond size could be reduced slightly, remember that the
earthwork tolerance will slightly affect characteristics of this pond. A slightly lower pond
bottom will reduce the aquifer thickness, thus reducing the recovery time. A slightly higher
pond bottom will reduce the retention volume and increase the discharge. So, when
considering the construction tolerance, this configuration looks good.

9. Run the other design storms.

The other storm frequencies should be calculated to check that the pre-
developed discharges are not exceeded. The results are in Table 9.4-10.
10. The stage-storage values used in this example have been based on length and
width dimensions applied to a frustum of a pyramid. When you apply the radii to
the corners, you would reduce the storage using the same pond dimensions, so
use an equivalent stage-area relationship when working with the contours within a
CADD file. Doing so will allow you to configure the pond for aesthetic purposes
while maintaining the necessary stage-storage relationship.

Chapter 9: Stormwater Management Facility 9-87


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Table 9.4-10 (Example 9.4-4, Closed basin)


100-year 50-year 25-year
Same
Config. as in Disch. Vol. Disch. Rate Disch. Vol. Disch. Rate Disch. Vol. Disch. Rate
Table 9.4-9 (cfs x 103) (cfs) (cfs) x 103 (cfs) (cfs) x 103 (cfs)
1-hour
Pre 81.2 32.9 70.6 28.7 60.4 24.7
Post 38.9 8.2 31.5 6.7 24.5 5.2
2-hour
Pre 119 30.3 100 25.2 84.8 21.3
Post 78.5 13.8 60.7 9.1 48.3 7.3
4-hour
Pre 160 24.7 139 21.7 115 18.1
Post 124 21.1 103 16.4 78.9 9.9
8-hour
Pre 211 28.4 181 24.4 151 20.4
Post 185 21.7 155 14.9 125 9.8
24-hour
Pre 315 10.1 273 8.8 233 7.5
Post 299 8.1 258 7.0 217 5.8
3-day
Pre 447 7.9 380 6.9 329 6.1
Post 436 7.4 369 6.4 317 5.6
7-day
Pre 572 5.9 495 5.2 427
Post 556 5.9 478 5.2 411 4.64.6
10-day
Pre 621 7.3 582 6.9 509 6.2
Post 610 7.3 570 6.9 497 6.2

10-year 5-year 2-year


1-hour
Pre 44.6 18.3 35.8 14.8 22.8 9.4
Post 14.1 3.0 8.9 1.9 2.4 0.6
2-hour
Pre 67.2 16.6 53.9 13.1 35.8 8.4
Post 33.9 5.2 23.5 3.7 10.2 1.7
4-hour
Pre 92.1 14.6 74.1 11.9 47.6 7.7
Post 58.3 7.3 42.3 5.4 19.8 2.6
8-hour
Pre 119 16.1 100 13.4 63.8 8.5
Post 93.0 7.1 73.9 5.5 38.6 2.7
24-hour
Pre 189 6.1 147 4.7 92.2 2.9
Post 173 4.6 130 3.5 72.9 2.0
3-day
Pre 255 4.9 207 4.1 139 2.9
Post 242 4.5 193 3.7 123 2.6
7-day
Pre 333 3.8 283 3.3 198 2.4
Post 316 3.8 265 3.3 177 2.4
10-day
Pre 389 4.9 310 4.0 224 3.0
Post 376 4.9 296 4.0 208 3.0

Chapter 9: Stormwater Management Facility 9-88


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

9.4.7 Off-Site Inflows


In 2013, House Bill 599 (2012), enacted as Chapter 2012 174, Laws of Florida, amended
Chapter 373, F.S. to create provision Section 373.413(6). This provision states that
“FDOT is responsible for treating stormwater generated from state transportation projects
but is not responsible for the abatement of pollutants and flows entering its stormwater
management systems from off-site sources; however, this subsection does not prohibit
the Department of Transportation from receiving and managing such pollutants and flows
when cost effective and prudent. Further, in association with right-of-way acquisition for
state transportation projects, the Department of Transportation is responsible for
providing stormwater treatment and attenuation for the acquired right-of-way but is not
responsible for modifying permits for adjacent lands affected by right-of-way acquisition
when it is not the permittee.”

FDOT generally has four options when dealing with offsite flows that would be
intercepted by a linear transportation project:

1) Bypass offsite flows around the project's treatment system

2) Accept offsite flows and direct them to a treatment system that is designed
to treat the transportation project and the offsite flow

3) Accept offsite flows and direct them to a treatment system that is designed
to treat only the project

4) Accept offsite flows and direct them to a treatment system that is designed
to treat the project and partially treat the off-site property

Empirical nutrient loading model results (Harper Methodology) show that—in all cases
involving wet detention treatment, even when the treatment facility is designed for only
the project area—there is an overall environmental benefit achieved by commingling
(i.e., the net pollutant reduction is greater).

The same modeling shows that—for retention-type treatment systems, when the offsite
lands provide equal or greater nutrient loading when compared to the FDOT project—
there is also an overall environmental benefit achieved by commingling, even when the
treatment facility is designed for only the project area. Thus, in these cases, the water
quality at downstream points of discharge from the commingled system will be equal to
or better than those systems that bypass offsite flows. Based on these results, FDEP
and the WMDs support allowing commingling in these cases without requiring further
analysis as long as the proposed treatment pond meets the ERP design requirements
for the runoff from the project area and results in an overall environmental benefit.

The same empirical nutrient loading model results (Harper Methodology) show that—
where undeveloped or unimproved offsite lands flow into onsite FDOT dry retention

Chapter 9: Stormwater Management Facility 9-89


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

ponds—the water quality at downstream points of discharge from the commingled


system may, in some cases, be worse than those systems that bypass offsite flows. As
such, these designs should be evaluated on a case-by-case basis to ensure that
environmental protection is not diminished.

In summary:
 For wet detention:
 Commingle offsite inflows unless cost or hydraulic issues lead to
bypassing
 For dry retention:
 Commingle developed offsite inflows unless cost or hydraulic issues lead
to bypassing
 For inflows from lower EMC areas, consult the District Drainage Engineer
- Calculate change in nutrient removal
- If reduction in treatment, evaluate B/C

9.4.8 Commingling of Untreated Onsite Runoff


When you are adding new lanes to an existing roadway that has no formal water quality
treatment, if you leave the drainage system for the existing roadway untouched, water
quality treatment does not need to be provided for the existing unchanged lanes.
Regardless, as a matter of good environmental stewardship, attempt, if economically
prudent, to bring the runoff from the existing roadway into the treatment system for the
new lanes. Just as in the section above for offsite inflows, commingling of existing onsite
runoff will always result in improved downstream water quality, even if the stormwater
management system is sized only for the new lanes. If economically prudent, consider
increasing the pond sizes to treat the old system, even though not required.

9.5 OUTLET CONTROL STRUCTURES


9.5.1 Weirs
The most common form of flow control is a weir notched into the side of a concrete
structure. To maximize the predictability of the flow, the weir should be smaller than the
distance between the inside edges of the walls. This smaller size will allow air to get under
the nappe. Using a weir size equal to the inside edges of the walls would create an
unstable condition when the flow is attempting to spring free from the leading edge of the
weir.

Chapter 9: Stormwater Management Facility 9-90


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Sometimes outlet control structures contain multiple (or staged) weirs, such as a small
weir at a low elevation with a larger weir at a higher elevation. These compound weirs
can be handled in one of two ways. SWFWMD recommends treating the lower slot as an
orifice, with head (H) measured to the centroid when the opening is submerged. Then
you can model the upper portion with standard weir formulas and the two flows are added.
Alternatively, you can extend the lower slot computations to the water surface. Then you
model the flows from the sides of the upper slot as a separate weir and add the flows. In
either case, a totally smooth transition in the performance curve at the stage of the upper
weir crest cannot be expected. Some amount of manipulation of the curve should be
made to smooth it at the transition.

9.5.2 Discharge Coefficients


The following coefficients are recommended for the typical concrete box outlet control
structure. You will find these values documented in a report titled “Performance and
Design Standards for Control Weirs, An Investigation of Discharge Through Slotted
Weirs,” based on a study by the University of South Florida, March 1993; WPI nos.
0510610, & 0510522. Contact the FDOT Research Center at (850) 414-4615 to obtain a
copy.

The first two tables apply to control devices formed into the wall of the outlet control
structure. As a result, the thickness of the structure wall will affect the discharge
coefficient. The discharge coefficient first rises with increasing head and then remains
constant. This behavior is observed for both orifices and weirs and is caused by
attachment of the flow at the sides of the opening. The wall thickness of the typical FDOT
structure can vary depending on whether the structure is precast or “cast in place.” Unless
you specify “cast in place,” assume that the structure will be precast. The Roadway and
Traffic Design Standards specify the wall thickness.

Chapter 9: Stormwater Management Facility 9-91


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

Table 9.5-1: Orifice Discharge Coefficients


ORIFICE Discharge Coefficient, CD
Condition of upstream edge H/b<0.6 H/b>0.7
Concrete edge1 0.276 (H/b) + 0.491 0.709
90 elbow fitting
o 0.620 (H/b) + 0.284 0.645
1 These values account for edge imperfections, chipping, wear, and some amount of bevel.

CD is dimensionless, to be used with the equation: Q = CD AO (2gH)1/2


AO = area of opening
H = distance of water surface above orifice center
b = thickness of the structure wall

Table 9.5-2: Weir Discharge Coefficients


RECTANGULAR WEIR Weir Coefficient, CW
0.25<H/b<2.01 H/b>2.01
Condition of upstream edge
0.468(H/b)
Concrete edge2 3.45
+2.45
1 A typographical error exists in the original report, which shows this value to be 2.5 instead of 2.0.
2 These values account for edge imperfections, chipping, wear, and some amount of bevel.

CW is dimensional and calculated from CW = (2g)½ CD


CW is to be used in the equation: Q = CW L H 1.5
L = width of weir
H = distance of water surface above the weir crest
b = thickness of the structure wall

Thin plate weirs fabricated from metal and bolted over a larger opening in the wall provide
a more-uniform, predictable performance. Install the metal weir plate over an opening of
sufficient size to ensure that the flow passing over the weir encounters no interference
from the headwall. The plate’s thickness should be 0.25 inch or less to approximate a
sharp edge. If you construct it as discussed here, the weir coefficient is as follows and is
independent of height.

Metric US Customary
Weir Coefficient CW for Thin Plates 1.73 3.13

Chapter 9: Stormwater Management Facility 9-92


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

9.5.2.1 Submerged Control Devices


For weirs, use the Villemonte relationship to compute the ratio of flow under submerged
conditions to flow under free discharge.

QS
= (1 - S n )0.385
QF

where
QS = Flow under submerged conditions
QF = Flow under free discharge
S = H2/H1 = Submergence ratio
H1 = Upstream headwater
H2 = Downstream headwater
n = 1.5 for rectangular weirs, & 2.5 for triangular weirs

Use the following similar relationship for orifices.

QS
= (1 - S )0.5
QF

Chapter 9: Stormwater Management Facility 9-93


January 2019

Drainage Design Guide


Chapter 9: Stormwater Management Facility

9.5.3 Skimmers
Regulatory agencies commonly require skimmers to prevent oil and grease from leaving
the pond. The head loss due to skimmers is minimized if the flow area under the skimmer
is three times larger than the flow area of the weir. If this area is provided, you need not
calculate the head loss associated with the skimmer.

If it is impossible to provide the flow area mentioned above, the head loss across the
skimmer can be calculated using this formula:

HL = k V2/2g

where:
k= Loss coefficient
V = Velocity under the skimmer

A loss coefficient, k, of 0.2 is recommended based on a May 25, 1988, SWFWMD


Technical Memorandum by R.E. Benson Jr., P.E., Ph.D.

9.5.4 Miscellaneous
To minimize plant growth, construct a concrete apron around the outlet control structure.
You should extend it five feet from the structure.

In wet detention facilities, the outlet control structure generally includes a drawdown
device, such as an orifice or a v-notch weir, to establish the normal water level and to
slowly release the treatment volume. If the drawdown device is smaller than three inches
wide or less than 20 degrees for v-notches, include a device to eliminate clogging.
Examples of such devices include baffles, grates, screens, and pipe elbows.

It is not necessary to use the ditch bottom inlet type grates on outlet control structures
unless needed for safety. If the structure is accessible to the public or maintenance
vehicles will traverse it, grates are recommended.

Always consider the effects of storms that are more severe than what was designed for.
Sometimes an overflow spillway can be built into the berm. Or additional flow can
sometimes pass through the top of the outlet control structure while using the freeboard
to store more volume and create additional head.

Chapter 9: Stormwater Management Facility 9-94


January 2019

Drainage Design Guide


Chapter 10: Temporary Drainage Design

CHAPTER 10: TEMPORARY DRAINAGE DESIGN

10. TEMPORARY DRAINAGE DESIGN .............................................................10-1

10.1 Purpose.........................................................................................................10-1

10.2 Criteria ..........................................................................................................10-1

10.3 Method ..........................................................................................................10-2

10.4 Detours .........................................................................................................10-2


10.4.1 On Existing Streets .................................................................................10-2
10.4.2 Constructed Detours...............................................................................10-3

10.5 Construction of the Proposed Facility.......................................................10-3


10.5.1 Milling and Drop-Offs ..............................................................................10-3
10.5.2 Driveways ...............................................................................................10-4
10.5.3 Temporary Drains and Curb Inlets..........................................................10-4
10.5.4 Box-Culvert Extensions ..........................................................................10-5
10.5.5 Temporary Barrier Wall...........................................................................10-5
10.5.5.1 Flow under Temporary Barrier Walls ..................................................10-6

Chapter 10: Temporary Drainage Design 10-i


January 2019

Drainage Design Guide


Chapter 10: Temporary Drainage Design

10. TEMPORARY DRAINAGE DESIGN

10.1 PURPOSE
The primary purposes for designing temporary drainage are to:

 Minimize travel lane flooding


 Prevent damage to property adjacent to a project during construction
 Facilitate construction activities by temporarily rerouting or altering
drainage conveyances

The information covered in this chapter offers practical considerations and


solutions to physical conditions that affect drainage efficiency at roadway
construction sites. This is intended for the drainage system engineer and designer
and indirectly for other technical personnel. Proper use of this information includes
ongoing communication and collaboration with engineers responsible for
roadways, structures, and traffic control plans.

10.2 CRITERIA
Consult the Drainage Manual for hydraulic and hydrologic criteria that apply to the
design of temporary drainage systems. Specifically, refer to Section 2.2 in the
Drainage Manual for design frequencies at temporary roadside and median
ditches, swales, and side drains. Refer to Section 3.3 in the Drainage Manual
under “General Design” for design frequencies of temporary storm drain systems.
Refer to Section 4.3.2 in the Drainage Manual for design frequencies at temporary
culverts, bridge culverts, and bridges.

Some drainage situations are much more common in temporary conditions than in
permanent conditions. For example, in temporary MOT scenarios, a temporary
traffic lane ultimately will become a paved shoulder; thus, the shoulder area,
proposed to carry spread in the permanent condition, now has traffic flowing with
little room before being confined by a barrier wall. Thus, stormwater may pool
along the temporary barrier wall or curbing near an inside high-speed lane.

Be aware that criteria can be different for permanent and temporary conditions on
the same section of roadway, primarily because the two conditions can have
differing design speeds. Consult Section 3.9 of the Drainage Manual to determine
spread criteria.

Chapter 10: Temporary Drainage Design 10-1


January 2019

Drainage Design Guide


Chapter 10: Temporary Drainage Design

10.3 METHOD
Carefully consider the temporary conditions that could arise during the construction
or rehabilitation of a permanent transportation facility. Safety of travelers and
workers, cost of construction, and the time required to complete construction tasks
all are affected adversely when temporary drainage is not adequately addressed
during design. You can reduce construction delays resulting from inclement
weather conditions by creating a well-designed temporary drainage system.
Further, unsafe traveling conditions and construction delays increase the cost of
projects. Provide temporary drainage features where and when they are needed.

Design temporary drainage for construction sites with emphasis on the following:

(1) Drain detours efficiently, whether on existing streets or temporary lanes.


(2) Prevent drainage problems caused by construction staging. Examine
detour designs in the light of construction staging to determine whether
construction activities might divert or trap water and compromise safety
and efficiency.
(3) Provide details for box culvert extensions that require a temporary
rerouting of water away from work areas.
(4) Provide emergency relief that will convey storm events without substantial
risk of flooding travel lanes.

10.4 DETOURS
The term “detour” is defined in FDM 240. Detours may be either located on existing
streets or constructed with temporary paved or graded lanes. Design the drainage
for detours on existing streets by using the existing street drainage system while
preventing overtaxing of the system on those streets. When temporary lanes or
roads must be constructed, provide design for temporary drainage systems.

10.4.1 On Existing Streets


Concentrate on construction site ingress and egress points when designing
drainage for detours routed over existing streets or roads. Ensure that the
construction site does not divert drainage onto the detour route in excess of the
capacity of the existing street drainage systems. Conversely, ingress and egress
locations cannot be allowed to divert excessive water onto the travel lanes or to
accommodate surface drainage that causes erosion of the construction site.

Chapter 10: Temporary Drainage Design 10-2


January 2019

Drainage Design Guide


Chapter 10: Temporary Drainage Design

10.4.2 Constructed Detours


Refer to Figures 10.5-1 and 10.5-2. When constructing detours, provide designs
for temporary drainage systems that prevent stormwater from pooling or backing
up on lanes where traffic will travel. Detour lanes, whether constructed in a median
or off an outside edge of pavement, often are built on fill that can disrupt the flow
of stormwater unless you design temporary measures to carry the water through
or around the fill area.

Include directional flow arrows in the plans when a swale is used between fill for a
temporary road and fill for a permanent road. See Figure 10.5-1.

Consider every temporary low area created when a detour road interrupts a
proposed ditch gradient as a possible location for temporary drainage structures.

Temporary detours sometimes have vertical curvature or gradients that are


independent of the main project. Be careful not to overlook these areas when
locating temporary drainage structures. See the temporary pipe shown in Figure
10.5-1.

10.5 CONSTRUCTION OF THE PROPOSED FACILITY


Provide a design for temporary drainage of construction features, such as milled
lanes, drop-offs between lanes, turnout construction, and construction operations
for new side drains, cross drains, and box culverts. Examine areas where traffic
control items, especially temporary barrier walls, might cause water to pond.

10.5.1 Milling and Drop-Offs


A drainage problem can occur where the natural sheet flow across the roadway
surface is curbed by an adjacent lane; this problem will be most evident in sag
vertical curves where curbed water is directed to the low point and then flows in
concentrated form across an adjacent lane. The best way to avoid this is to
schedule construction phasing so that an adjacent lane does not curb natural sheet
flow across the roadway surface. It is difficult to avoid this curbing effect when
adding lanes to the median side of a divided highway, or to the high side of a
section in super-elevation.

Chapter 10: Temporary Drainage Design 10-3


January 2019

Drainage Design Guide


Chapter 10: Temporary Drainage Design

Where you cannot avoid this curbing effect with construction phasing, provide
temporary measures to prevent flooding of adjacent travel lanes. Sandbags or
temporary asphalt curb can be effective in directing runoff away from travel lanes.
Prevent overtopping flow at drop-offs by calling for sandbags or temporary asphalt
curb to be placed along the drop-off to force water away from travel lanes. Refer
to Figures 10.5-3, 10.5-4 and 10.5-5.

If the speed limit in a work zone is 45 mph or less, you can use intermittent
transverse saw-cuts in travel lanes to allow water to flow through the travel area
without overtopping.

10.5.2 Driveways
Constructing driveways can cause water to pond in the turnout area and
subsequently flood adjacent lanes on the roadway. Include details in the plans for
placing sandbags or temporary asphalt curb along outside edges of pavement
adjacent to turnout construction to prevent water in the turnout site from flowing
across the travel lanes. Provide details for temporary flumes and inlets, where
needed, to direct water at turnout sites into the storm drain system, thus preventing
water from collecting in low areas and/or causing erosion.

10.5.3 Temporary Drains and Curb Inlets


In accordance with Standard Plans, Index 425-001, provide a note in the plans
requiring “temporary drains for subgrade and base” at inlets, or include a similar
detail in the plans. Either detail will require construction of temporary drains for
water that is trapped on base and subsequent paving layers around inlets during
construction.

Chapter 10: Temporary Drainage Design 10-4


January 2019

Drainage Design Guide


Chapter 10: Temporary Drainage Design

10.5.4 Box-Culvert Extensions


Furnish a temporary drainage design that will provide dry work areas for box
culvert extensions during common storms and provide flood protection during
severe storms. When standing or flowing water occupies box culverts that are to
be extended, divert this water away from work areas for the duration of required
work. Refer to Figures 10.5-6 through 10.5-10 for examples of details for inclusion
in plans. Include details and notes in the plans that provide, at a minimum, the
following information:

1. Provide sizes for diversion pipes that are to be inserted into existing box
culverts.
2. Show the configuration requirements for sandbagging.
3. Include measures for stabilization and erosion control.
4. List any items that must be removed prior to final grading.
5. List descriptions and quantities for items not included in the cost of the
structure.

10.5.5 Temporary Barrier Wall


The temporary barrier wall most commonly used on Department projects is the
Type K Temporary Concrete Barrier System detailed in Standard Plans, Index 102-
110. The concrete units are configured with two 27-inch drainage slots.

When needed, perform spread calculations for temporary precast concrete barrier
wall, based on rainfall of four inches per hour.

Chapter 10: Temporary Drainage Design 10-5


January 2019

Drainage Design Guide


Chapter 10: Temporary Drainage Design

10.5.5.1 Flow under Temporary Barrier Walls

For barrier walls placed on a longitudinal grade, an approach to calculating spread


that is similar to the approach used for curb inlets is summarized as follows.

1. Determine the flow approaching the slot.


2. Assume normal depth of flow at the slot and use the modified Manning’s
Equation for shallow channel flow to determine the spread and associated
depth of flow (y) at the edge of the barrier wall.

 0.56  5 3 1 2 8 3
Q Sx SL T
 n 
where:
Q = Gutter flow rate, in ft3/sec
n= Manning’s roughness coefficient (see Table B-2, Appendix B)
Sx = Pavement cross slope, in ft/ft
SL = Longitudinal slope, in ft/ft
T= Spread, in ft

3. Using the depth of flow (y) at the edge of the barrier wall, determine the
flow through the slot using the capture equations in HEC 22, and
assuming that the two 27-inch slots operate independently. The
Department suggests that the slot flow be reduced to 75 percent of the
equation value to account for 25 percent blockage.
4. Subtract the flow through the slot from the flow approaching the slot to
determine the flow bypassing the slot.
5. Add the bypass flow to the surface runoff for the next slot.
6. Repeat steps 1 through 5 for the length of the barrier wall or until
equilibrium is achieved.

Table 10.5-1 provides the spread values for several pavement widths and slopes
using the approach described above.

For sag vertical curves, you will likely need a more complicated approach. Several
items change with changing (y) values. As the depth of ponding increases, the
length of roadway draining directly (not including the bypass from approach
grades) to the ponded area increases, as does the number of slots that operate in
sump condition.

Chapter 10: Temporary Drainage Design 10-6


January 2019

Drainage Design Guide


Chapter 10: Temporary Drainage Design

Table 10.5-1: Spread at Temporary Barrier Walls (27”) Slots


Cross
Pavement Longitudinal Longitudinal Longitudinal
Slope
Width Slope = 0.3% Slope = 1% Slope = 3%

ft/ft Spread (ft) Spread (ft) Spread (ft)

12 feet 0.01 3.00 2.39 1.95


0.02 1.94 1.55 1.26
0.03 1.51 1.20 0.98
0.01 3.74 2.98 2.43
24 feet 0.02 2.42 1.93 1.57
0.03 1.88 1.5 1.22
0.01 4.33 3.45 2.81
36 feet 0.02 2.8 2.24 1.82
0.03 2.18 1.74 1.41

Chapter 10: Temporary Drainage Design 10-7


January 2019

Drainage Design Guide


Chapter 10: Temporary Drainage Design

TEMPORARY DETOUR AND BRIDGE

1. Prevent water from being trapped in the area where the limits of temporary fill overlap the limits of permanent fill
by providing a design for temporary ditches that drain positively to the stream, as shown on the left side of this
detail, or provide a design for temporary drains under the detour, as shown on the right.
2. When the grade of a detour road is lower than ditch elevations, take care to avoid any sag in the ditch grade
that could collect water until it pops over and spills onto the detour. Find and solve this problem during design.
Do not force the contractor to handle it. The temporary pipe shown is a possible scenario.

Figure 10.5-1: Temporary Detour and Bridge

Chapter 10: Temporary Drainage Design 10-8


January 2019

Drainage Design Guide


Chapter 10: Temporary Drainage Design

TEMPORARY DETOUR IN MEDIAN

1. Conditions shown here may differ as to location, amount, configuration, direction of flow, etc., with regard to
temporary barrier wall, temporary drainage structures, and work area.
2. Calculate spread using the method defined in Section 10.5.5. Prepare a design for temporary conditions that
meets the requirements of this chapter and the Drainage Manual.
3. Require installation of temporary drainage structures where needed to maintain normal flow through the work
area.
4. Where two or more runs of temporary barrier wall are parallel to each other, on or adjacent to a common width
of pavement, a temporary slotted drain may be required to hold spread within acceptable limits.

Figure 10.5-2: Temporary Detour in Median

Chapter 10: Temporary Drainage Design 10-9


January 2019

Drainage Design Guide


Chapter 10: Temporary Drainage Design

SECTION SHOWING SANDBAGGING TO PREVENT OVERTOPPING

1. Place one row of sandbags, two bags deep, adjacent to drop-offs where
overtopping spread may occur and where the down-slope lanes must be
used to maintain traffic.
2. When possible, avoid this situation by phasing milling and pavement lifts
so that the exposed sides of drop-offs face down-slope.
3. Consider using this detail wherever new pavement lifts must be placed
or existing pavement must be milled, and where the exposed edge of a
drop-off faces upslope.

Figure 10.5-3: Section Showing Sandbagging to Prevent Overtopping

Chapter 10: Temporary Drainage Design 10-10


January 2019

Drainage Design Guide


Chapter 10: Temporary Drainage Design

Figure 10.5-4: Typical Terminus Figure 10.5-5: Limits of Sandbagging

Chapter 10: Temporary Drainage Design 10-11


January 2019

Drainage Design Guide


Chapter 10: Temporary Drainage Design

Figure 10.5-6: Temporary Drainage at Box Culvert Extension

Chapter 10: Temporary Drainage Design 10-12


January 2019

Drainage Design Guide


Chapter 10: Temporary Drainage Design

Figure 10.5-7: Temporary Drainage for Extension of Double Barrel Box-


Culverts

Figure 10.5-8: Section Normal to Figure 10.5-9: Section AA


Culvert (Single Barrel) (Single Barrel)

Figure 10.5-10: Section BB (Single Barrel)

Chapter 10: Temporary Drainage Design 10-13


January 2019

Drainage Design Guide


Appendix A: Data Collection/Published Data

APPENDIX

A. DATA COLLECTION/PUBLISHED DATA

Appendix A: Data Collection/Published Data


January 2019

Drainage Design Guide


Appendix A: Data Collection/Published Data

TABLE OF CONTENTS

A. Data Collection/Published Data ........................................................................ A-1


A.1 Data Collection...............................................................................................A-1
A.2 Published Data...............................................................................................A-5
A.2.1 Soils ...........................................................................................................A-5
A.2.2 Land Use ....................................................................................................A-5
A.2.3 Precipitation................................................................................................A-6
A.2.4 Topography and Contour Information .........................................................A-7
A.2.5 Streamflow and Flood History ....................................................................A-7
A.3 Field Investigations and Surveys ...................................................................A-8
A.3.1 Drainage Areas ..........................................................................................A-8
A.3.2 High-Water Information ..............................................................................A-9

Appendix A: Data Collection/Published Data A-i


January 2019

Drainage Design Guide


Appendix A: Data Collection/Published Data

A. DATA COLLECTION/PUBLISHED DATA


A.1 DATA COLLECTION
All of the information presented in this section may not be required to address the needs
of each project.

Table A-1, below, lists examples of data, along with typical sources and uses, for three
data categories, including:

 Completed or ongoing studies


 Natural resource base
 Manmade features

Drainage projects typically require numerous potential sources of data. Identifying these
sources can be difficult, and making the subsequent necessary contacts can be time-
consuming. To assist in this process, Table A-1, below, includes typical data sources. In
many cases, the local community or Water Management District in which the drainage
project is being conducted is either the best source of data or the most logical starting
point.

The primary use of drainage data is to quantify the hydrologic/hydraulic characteristics


of the watershed to evaluate stormwater runoff discharge and volume. Quantification of
watershed characteristics is a must for both existing and future conditions. Table A-1,
below, presents examples of data uses.

Before initiating calculations, collect drainage data using the following general
guidelines:

1. Identify data needs, sources, and uses, using Table A-1 as a checklist. Much of this
information will have to be provided in the environmental document and supporting
files.

2. Collect published data, based on sources identified in Step 1 and information


presented in Section A.2.

3. Compile and document the results of Step 2, and compare data needs and uses
with the availability of published data. Identify any additional field data needs.

4. Collect field data based on needs identified in Steps 1 and 3, using information
presented in Section A.3.

5. Compile and document the results of Step 4.

Appendix A: Data Collection/Published Data A-1


January 2019

Drainage Design Guide


Appendix A: Data Collection/Published Data

Table A-1: Data Needs, Sources, and Uses


Data Needs Examples Typical Sources Examples of Uses
Storm Master Plan County, City, or Water Management Establish type and configuration of future
District stormwater control facilities

208 Plan  U.S. Environmental Protection Delineate watersheds and subbasins


Agency
 Regional Planning Agency
SCS Pl 566 Plan U.S. Environmental Protection Agency Establish flood flows, stages, and area of
inundation on principle streams
Flood Plain Information U.S. Army Corps of Engineers Establish flood flows, stages, and area of
inundation on principle streams
Special Studies  City or County Varies with Study
 U.S. Geological Survey
1. Completed or ongoing studies
 Regional Planning Agency

Flood Insurance Study  U.S. Federal Emergency Establish flood flows, stages, and area of
Management Agency/Department of inundation of principle streams
Housing and Urban Development
 City or County

Topographic Map  U.S. Geological Survey  Delineate watersheds and subbasins


 Regional Planning Agency  Identity potential detention sites
 Water Management District
 Determine land slope
 Field Survey
 FL. Dept. of Environmental
Protection

Appendix A: Data Collection/Published Data A-2


January 2019

Drainage Design Guide


Appendix A: Data Collection/Published Data

Data Needs Examples Typical Sources Examples of Uses


Soils  U.S. Natural Resources  Determine the runoff coefficients,
Conservation Service curve numbers, and other runoff
 Construction Logs factors
 Evaluate erosion potential
 Project construction condition

Historic Inundation Areas and High  U.S. Geological Survey  Document location and severity of
Waters  City or County historic inundation and other
 Water Management District problems
 Regional Planning Agency
2. Natural Resource Base  News Media – Newspapers, Radio,
T.V.
 Museums, Historical Societies
 Residents
 Field Survey

Precipitation Intensity-Soils Duration-  National Weather Service Develop design storms


Frequency Data  Water Management District

Historic Stage and Discharge  National Weather Service Assess severity of historic floods
 Water Management District
Stream Stage and Discharge  U.S. Geological Survey  Develop discharge-probability
 Water Management District relationships
 Asses severity of historic floods

Existing Land Use Areas and High Waters  Regional Planning Agency Determine runoff coefficients, curve
 Field Survey numbers, and other factors

Land Use Plan  Regional Planning Agency Determine runoff coefficients, curve
3. Manmade Features  City or County numbers, and other factors

Zoning Map and Ordinance City or County Project future land use

Subdivision Plats City or County  Project future land use


 Established type and configuration of
future stormwater control facilities

Appendix A: Data Collection/Published Data A-3


January 2019

Drainage Design Guide


Appendix A: Data Collection/Published Data

Data Needs Examples Typical Sources Examples of Uses


Agricultural and Other Land Management  U.S. Natural Resources Conversation Determine runoff coefficients, curve
Measures Services numbers, and other factors
 Regional Planning Agency
 Field Survey

Transportation, Sewage and Other Public  Regional Planning Agency  Establish future watershed and sub-
Facility-Systems and Plans  City or County basin divides
 Department of Transportation  Project future land use

 Delineate existing/future watershed


Stormwater Systems Maps, Plans,  Regional Planning Agency and sub-basin divides
Profiles; As-Builts  City or County  Develop hydraulic characteristics

 Delineate existing/future watershed


Bridge, Culverts, Channels, and Other  Water Management District
and sub-basin divides
Hydraulic Structure As-Builts or Plans  Department of Transportation
 Develop hydraulic characteristics
Subdivisions Plats  Field Survey

Identify potential sites for detention and


Land Ownership—Public vs. Private City or County
other facilities

Appendix A: Data Collection/Published Data A-4


January 2019

Drainage Design Guide


Appendix A: Data Collection/Published Data

A.2 PUBLISHED DATA

Published data include soils, land use, precipitation, topography and contour,
streamflow and flood history, and groundwater. A good basic reference for water
resources data in Florida is the Water Resources Atlas of Florida (Florida State
University, 1984). Of particular relevance to drainage projects are data on weather and
climate, surface water, groundwater, water quality, drainage, flood control, navigation,
and ecosystems.

A.2.1 Soils
Collect published soils data by following this procedure:

1. Identify soils data needed to evaluate runoff, soil erosion, slope and foundation
stabilities, and hydraulic conductivity.

2. Obtain soils data from the U.S. Natural Resources Conservation Service (formerly
the U.S. Soil Conservation Service [SCS]) in the form of detailed soils reports for
the county area being considered. Old plans, construction logs, and soil boring
results can provide additional site-specific data. Specific project information usually
is available during the final design stage.

When a project involves a channel in which storm tide surge conditions may be
expected to result in erosion of the channel, the geology in the area of the channel is
important in analyzing the nature of the potential erosion enlargement. More detailed
and extensive borings may be important, which would not be the case where channel
stability is reasonably assured. You may need to make a preliminary assessment of the
potential for enlargement to specify the extent of the geotechnical study required.

A.2.2 Land Use


Collect published land use data by following this procedure:

1. Determine historical land use from older land use maps or aerials.

2. Determine current land use from sources such as land use maps, aerial
photographs, and field reconnaissance. Contact appropriate county and municipal
governments. Regional Planning Councils and Water Management Districts also
may have existing land use data. Compare historical (from Step 1) and current
land use to identify areas undergoing rapid growth and an approximate rate of
change. Establishing land use at the time of design can be crucial to project
success.

Appendix A: Data Collection/Published Data A-5


January 2019

Drainage Design Guide


Appendix A: Data Collection/Published Data

3. Determine future land use based on projections of existing land use, land use
plans, and site-specific layouts of proposed development, zoning maps, and
discussions with public officials. County and municipal governments as well as
Regional Planning Councils and Water Management Districts also may be good
sources of future land use data.

4. Ascertain the existence of master drainage plans, stormwater management plans,


and similar plans that may designate or restrict land use.

A.2.3 Precipitation
Collect published precipitation data by following this procedure:

1. Select an appropriate procedure for hydrologic calculations using information


presented in this handbook.

2. Determine the type of precipitation data needed. Generally, either intensity-


duration-frequency (IDF) curves or hyetographs for historic or design storm
conditions are used.

3. Collect published precipitation data. The primary source is the National Weather
Service. Additional data may be available from Water Management Districts.
Sources of published precipitation data are briefly discussed below.

A series of publications by the National Weather Service (formerly the U.S. Weather
Bureau) presents precipitation depth-duration-frequency data developed from observed
precipitation data across the United States. HYDRO-35, by Frederick et al. (1977), is
particularly useful for small drainage areas, since rainfall depths for durations of 5, 10,
15, 30, and 60 minutes are presented for return periods of 2, 5, 10, 25, 50, and 100
years. Technical Paper No. 40, by Hershfield (1961), commonly known as TP-40, is a
standard reference for obtaining hydrologic design rainfall depths for durations of 30
minutes and one, 2, 3, 6, 12, or 24 hours, and for return periods of one, 2, 5, 10, 25, 50,
and 100 years. Technical Paper No. 49, by Miller (1964), extends the depth-duration-
frequency data presented by Hershfield (1961) to include rainfall depths for durations of
2, 4, 7, and 10 days at return periods of 2, 5, 10, 25, 50, and 100 years. The
Department has developed rainfall curves based on these references.

Statistical rainfall depth data for Florida is found in the National Oceanic and
Atmospheric Administration (NOAA) Atlas 14 Rainfall Data. These data are available at
http://hdsc.nws.noaa.gov/hdsc/pfds/pfds_map_cont.html?bkmrk=fl. FDOT rainfall
distributions and intensity-duration-frequency curves may be found at
http://www.fdot.gov/roadway/Drainage/files/IDFCurves.pdf.

Appendix A: Data Collection/Published Data A-6


January 2019

Drainage Design Guide


Appendix A: Data Collection/Published Data

A.2.4 Topography and Contour Information

Collect topographic data by using the following procedure:

1. Obtain published topographic data. The Florida Department of Environmental


Protection (FDEP) may have contour information at one- or two-foot contours
developed from LIDAR, Water Management Districts, and municipal or county
government agencies. The U.S. Geological Survey (USGS) has maps with five-
foot or 10-foot contour intervals, which often are not detailed enough for design.

2. If published data are either unavailable or inadequate for project needs, the
Department can develop contours from aerial photographs for large-scale projects
or by survey for small areas.

A.2.5 Streamflow and Flood History

Collect streamflow and flood history data by using the following procedure:

1. Obtain published data. The principal source of published streamflow data is the
USGS. Additional sources include Water Management Districts and municipal or
county government agencies.

2. Because published streamflow data may not be available for a specific project site,
an evaluation of flood history may require researching news media sources,
making field survey observations, and interviewing local residents and other
knowledgeable persons.

Groundwater

Data on groundwater levels and movements can be obtained from information on


existing detention ponds and other ponds in the area; existing non-pumping wells or
wells that could be temporarily shut off to determine the static groundwater level;
observations made by inspectors and others during construction of sanitary sewers,
storm drains, and major buildings; and regional or area-wide reports prepared by the
USGS or similar state agencies. If existing data sources are not sufficient to define the
position of the groundwater table, it may be necessary to construct special observation
wells, particularly at potential sites of detention facilities. These wells could be installed
in the boreholes used to take soil samples during a site-specific subsurface exploration.

Appendix A: Data Collection/Published Data A-7


January 2019

Drainage Design Guide


Appendix A: Data Collection/Published Data

A.3 FIELD INVESTIGATIONS AND SURVEYS

A.3.1 Drainage Areas


If sufficient topographic information for a project site is readily available, a field
determination of drainage area may not be necessary, but it is always advisable to spot-
check selected control elevations. For those project sites for which detailed information
is not available, perform field survey work. In all cases, a site visit is highly
recommended to confirm drainage area conditions.

Depending on District preference, drainage areas may be outlined by field survey or


drainage personnel on county maps, aerial photographs, USGS contour maps, or
specially prepared maps. Drainage area boundaries should connect with the job
centerline, typically at high points in grade or at other locations where there is a definite
division in the direction of stormflow runoff. After the overall areas are plotted, the
Drainage Engineer should subdivide the drainage area to show how the various
sections contribute to the structures in the proposed drainage or storm drain system.

Follow all drainage area boundaries from the project centerline, around the area being
covered, and close again at the centerline. There is no need to show ridges that do not
drain to the project unless this information is pertinent to determine runoff concentration
points or flow path segments. Clearly indicate by notation on the map all exceptions to
the rule for closing all drainage area boundaries. These notes should show location and
elevation of break-over or diversion to or from the drainage area.

Typically, a drainage area should close to each existing culvert along the project and for
each probable cross drain location. As an exception, note flow distribution information
where two or more structures operate together to drain a single area.

For municipal-type construction surveys, mark appropriate city maps or specially


prepared maps to show the boundaries of total areas contributing to the project. Mark
streets or other drainage facilities in these areas with flow arrows. In many instances,
elevations may have to be determined to accurately delineate direction of flow in
gutters.

Show all areas contributing to existing storm drains, which drain to or across the project.
In very flat terrain, as is found in South Florida, it often is necessary to develop profiles
for cross streets and parallel streets to make a definite determination of drainage areas.
In flat terrain, consider collecting additional field data about agricultural ditches to
confirm flow patterns.

Appendix A: Data Collection/Published Data A-8


January 2019

Drainage Design Guide


Appendix A: Data Collection/Published Data

Specially flown aerial photography is available for most new construction projects.
Ridge lines usually can be indicated on the photographs. When using photographs, the
field survey party should verify questionable points and supplement the information with
structure sizes, elevations, and high water marks as required. Determine drainage
areas by stereo interpretation with spot field survey work as appropriate.

A.3.2 High-Water Information


To evaluate flood elevations and establish roadway grades, you will need reliable high-
water information. Show high-water elevation locations upstream of the proposed
project, upstream of significant existing structures, and at some point along or at the
end of outfall ditch surveys. Clearly record the location at which a high-water elevation
is taken in the field notes, along with the date and time if available.

At many locations, it is not possible to obtain documented information on high water. In


such cases, estimate elevation by observing natural growth or by other means; the
survey crew should provide complete information on the methods used. The crew chief
should attempt to obtain information from local residents, maintenance personnel (both
state and county), and rural mail carriers, school bus drivers, police officers, and school
board officials.

The soils crew usually supplies water table information within the right of way; however,
the survey crew should note information pertaining to standing water, areas of heavy
seepage, or springs within the basin area.

Appendix A: Data Collection/Published Data A-9


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

APPENDIX

B. HYDROLOGY DESIGN AIDS

Appendix B: Hydrology Design Aids


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

TABLE OF CONTENTS

Tables

Table B-1: Overland Flow Manning’s n Values ............................................................B-1


Table B-2: Manning’s n Values for Street and Pavement Gutters ................................B-2
Table B-3: Recommended Manning’s n Values for Artificial Channels.........................B-3
Table B-4: Runoff Coefficients for Storm Return Period ≤ 10 Years ............................B-4
Table B-5: Design Storm Frequency Factors for Pervious Area Runoff Coefficients ...B-5
Table B-6: Definitions of Four SCS Hydrologic Soil Groups .........................................B-6
Table B-7: SCS Runoff Curve Numbers – Agricultural, Suburban, and Urban Land ....B-7
Table B-8: SCS Runoff Curve Numbers for Agricultural Use .......................................B-8
Table B-9: SCS Classifications of Vegetative Covers by Hydrologic Properties ..........B-9
Table B-10: USGS Regression Equations – Natural Flow Conditions - Region 1 ......B-10
Table B-11: USGS Regression Equations – Natural Flow Conditions - Region 2 ......B-11
Table B-12: USGS Regression Equations – Natural Flow Conditions - Region 3 ......B-12
Table B-13: USGS Regression Equations – Natural Flow Conditions - Region 4 ......B-13
Table B-14: USGS Nationwide Regression Equations for Urban Conditions .............B-14
Table B-15: Urban Watershed Regression Equations for Tampa Bay Area ...............B-15
Table B-16: Urban Watershed Regression Equations for Leon County, Florida ........B-16
Table B-17: USGS Watershed Regression Equations for West Central Florida .........B-17
Table B-18: USGS Watershed Regression Equations’ Range of Applicability for West
Central Florida .................................................................................................B-18
Table B-19: Department Intensity Duration Frequency (IDF) Regression Equation
Constants and Coefficients ..............................................................................B-19
Table B-20: Example Application of Department IDF Regression Equations .............B-22

Appendix B: Hydrology Design Aids B-i


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Figures

Figure B-1: Kinematic Wave Formula for Determining Overland Flow Travel Time ...B-23
Figure B-2: Overland Flow Velocities for Various Land Use Types ............................B-24
Figure B-3: Velocity versus slope for Shallow Concentrated Flow .............................B-25
Figure B-4: Regions for USGS Regression Equations – Natural Flow Conditions .....B-26
Figure B-5: Regions for USGS Regression Equations for Natural Flow Conditions in
West Central Florida ........................................................................................B-27

Appendix B: Hydrology Design Aids B-ii


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-1: Overland Flow Manning’s n Values

Recommended
Value Range of Values
Concrete 0.011 0.010 - 0.013
Asphalt 0.012 0.010 - 0.015
Bare sand a 0.010 0.010 - 0.016
Graveled surface a 0.012 0.012 - 0.030
Bare clay-loam (eroded) a 0.012 0.012 - 0.033
Fallow (no residue) b 0.05 0.006 - 0.16
Chisel plow (<1/4 tons/acre residue) 0.07 0.006 - 0.17
Chisel plow (1/4 - 1 tons/acre residue) 0.18 0.070 - 0.34
Chisel plow (1 - 3 tons/acre residue) 0.30 0.190 - 0.47
Chisel plow (>3 tons/acre residue) 0.40 0.340 - 0.46
Disk/Harrow (<1/4 tons/acre residue) 0.08 0.008 - 0.41
Disk/Harrow (1/4 - 1 tons/acre residue) 0.16 0.100 - 0.25
Disk/Harrow (1 - 3 tons/acre residue) 0.25 0.140 - 0.53
Disk/Harrow (>3 tons/acre residue) 0.30 -- --
No till (</4 tons/acre residue) 0.04 0.030 - 0.07
No till (1/4 - 1 tons/acre residue) 0.07 0.010 - 0.13
No till (1 - 3 tons/acre residue) 0.30 0.160 - 0.47
Plow (Fall) 0.06 0.020 - 0.10
Coulter 0.10 0.050 - 0.13
Range (natural) 0.13 0.010 - 0.32
Range (clipped) 0.08 0.020 - 0.24
Grass (bluegrass sod) 0.45 0.390 - 0.63
Short grass prairie a 0.15 0.100 - 0.20
Dense grass c 0.24 0.170 - 0.30
Bermuda grass c 0.41 0.300 - 0.48
Woods 0.45 -- --
__________
All values are from Engman (1983), unless noted otherwise.
a
Woolhiser (1975).
b
Fallow has been idle for one year and is fairly smooth.
c
Palmer (1946). Weeping love grass, bluegrass, buffalo grass, blue gamma grass, native grass
mix (OK), alfalfa, lespedeza.
Note: These values were determined specifically for overland flow conditions and are not
appropriate for conventional open channel flow calculations. See Chapter 3, for open
channel flow procedures.

Appendix B: Hydrology Design Aids B-1


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-2: Manning’s n Values for Street and Pavement Gutters

Range of
Type of Gutter or Pavement Manning’s n

Concrete gutter, troweled finish 0.012

Asphalt pavement:
Smooth texture 0.013
Rough texture [2] 0.016

Concrete gutter with asphalt pavement:


Smooth 0.013
Rough 0.015

Concrete pavement:
Float finish 0.014
Broom finish [3] 0.016

For gutters with small slopes, where sediment may accumulate


increase above values of n by 0.002

_____________
Reference: FHWA HEC-22

Notes:
1) Estimates are by the Federal Highway Administration.

2) The Department’s friction course is rough texture asphalt.

3) The Department’s standard is brush (broom) finish for concrete curb.


[Specification Section 520]

Appendix B: Hydrology Design Aids B-2


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-3: Recommended Manning’s n Values for Artificial Channels


Design Manning’s
Channel Lining Lining Description n Value

Bare Earth or Vegetative Linings

Bare earth, fairly uniform Clean, recently completed 0.022


Bare earth, fairly uniform Short grass and some weeds 0.028
Dragline excavated No vegetation 0.030
Dragline excavated Light brush 0.040
Channels not maintained Dense weeds to flow depth 0.100
Channels not maintained Clear bottom, brush sides 0.080
Maintained grass or sodded ditches Good stand, well maintained 2" - 6" 0.060*
Maintained grass or sodded ditches Fair stand, length 12" - 24" 0.200*

Rigid Linings

Concrete paved Broomed** 0.016


Concrete paved “Roughened” - standard 0.020
Concrete paved Gunite 0.020
Concrete paved Over rubble 0.023
Asphalt concrete Smooth 0.013
Asphalt concrete Rough 0.016

_____________
* Decrease 30% for flows > 0.7' (maximum flow depth 1.5').

** Because this is not the standard finish, it must be specified.

Appendix B: Hydrology Design Aids B-3


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-4: Runoff Coefficients for Storm Return Period ≤ 10 Years a

Sandy Soils Clay Soils


Slope Land Use Min. Max. Min. Max.

Flat Woodlands 0.10 0.15 0.15 0.20


(0-2%) Pasture, grass, and farmland b 0.15 0.20 0.20 0.25
Bare Earth 0.30 0.50 0.50 0.60
Rooftops and pavement 0.95 0.95 0.95 0.95
Pervious pavements c 0.75 0.95 0.90 0.95
SFR: 1/2-acre lots and larger 0.30 0.35 0.35 0.45
Smaller lots 0.35 0.45 0.40 0.50
Duplexes 0.35 0.45 0.40 0.50
MFR: Apartments, townhouses,
and condominiums 0.45 0.60 0.50 0.70
Commercial and Industrial 0.50 0.95 0.50 0.95

Rolling Woodlands 0.15 0.20 0.20 0.25


(2-7%) Pasture, grass, and farmland b 0.20 0.25 0.25 0.30
Bare Earth 0.40 0.60 0.60 0.70
Rooftops and pavement 0.95 0.95 0.95 0.95
Pervious pavements c 0.80 0.95 0.90 0.95
SFR: 1/2-acre lots and larger 0.35 0.50 0.40 0.55
Smaller lots 0.40 0.55 0.45 0.60
Duplexes 0.40 0.55 0.45 0.60
MFR: Apartments, townhouses,
and condominiums 0.50 0.70 0.60 0.80
Commercial and Industrial 0.50 0.95 0.50 0.95

Steep Woodlands 0.20 0.25 0.25 0.30


(7%+) Pasture, grass, and farmland b 0.25 0.35 0.30 0.40
Bare Earth 0.50 0.70 0.70 0.80
Rooftops and pavement 0.95 0.95 0.95 0.95
Pervious pavements c 0.85 0.95 0.90 0.95
SFR: 1/2-acre lots and larger 0.40 0.55 0.50 0.65
Smaller lots 0.45 0.60 0.55 0.70
Duplexes 0.45 0.60 0.55 0.70
MFR: Apartments, townhouses,
and condominiums 0.60 0.75 0.65 0.85
Commercial and Industrial 0.60 0.95 0.65 0.95

aWeighted coefficient based on percentage of impervious surfaces and green areas must be selected for
each site.

b Coefficients assume good ground cover and conservation treatment.

c Depends on depth and degree of permeability of underlying strata.


Note: SFR = Single Family Residential
MFR = Multi-Family Residential

Appendix B: Hydrology Design Aids B-4


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-5: Design Storm Frequency Factors for Pervious Area Runoff
Coefficients*

Design Storm
Return Period (years) Frequency Factor, XT

2 to 10 1.0
25 1.1
50 1.2
100 1.25

_____________
Reference: Wright-McLaughlin Engineers (1969).

* DUE TO THE INCREASE IN THE DURATION TIME THAT THE PEAK OR NEAR PEAK
DISCHARGE RATE IS RELEASED FROM STORMWATER MANAGEMENT SYSTEMS, THE
USE OF THESE SHORT DURATION PEAK RATE DISCHARGE ADJUSTMENT FACTORS IS
NOT APPROPRIATE FOR FLOOD ROUTING COMPUTATIONS.

Appendix B: Hydrology Design Aids B-5


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-6: Definitions of Four SCS Hydrologic Soil Groups

Hydrologic
Soil Group Definition

A Low Runoff Potential


Soils having high infiltration rates even when thoroughly wetted, consisting chiefly of
deep, well-to-excessively-drained sands or gravels. These soils have a high rate of
water transmission.

B Moderately Low Runoff Potential


Soils having moderate infiltration rates when thoroughly wetted and consisting chiefly of
moderately deep, to deep, moderately fine to moderately coarse textures. These soils
have a moderate rate of water transmission.

C Moderately High Runoff Potential


Soils having slow infiltration rates when thoroughly wetted and consisting chiefly of soils
with a layer that impedes downward movement of water, soils with moderate fine to fine
texture, or soils with moderate water tables. These soils have a slow rate of water
transmission.

D High Runoff Potential


Soils having very slow infiltration rates when thoroughly wetted and consisting chiefly of
clay soils with high swelling potential, soils with a permanent high water table, soils with
a clay pan or clay layer at or near the surface, and shallow soils over nearly impervious
material. These soils have a very slow rate of water transmission.
_____________
Reference: USDA, SCS, NEH-4 (1972).

Appendix B: Hydrology Design Aids B-6


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-7: SCS Runoff Curve Numbers – Agricultural, Suburban, and Urban Land

Hydrologic Soil Group


Land Use Description A B C D
Cultivated Landa:
Without conservation treatment 72 81 88 91
With conservation treatment 62 71 78 81
Pasture or range land:
Poor condition 68 79 86 89
Good condition 39 61 74 80
Meadow: good condition 30 58 71 78
Wood or Forest Land:
Thin stand, poor cover, no mulch 45 66 77 83
Good cover b 25 55 70 77
Open Spaces, Lawns, Parks, Golf Courses, Cemeteries:
Good condition: grass cover on 75% or more of the area 39 61 74 80
Fair condition: grass cover on 50% to 75% of the area 49 69 79 84
Poor condition: grass cover on 50% or less of the area 68 79 86 89
Commercial and Business Areas (85% impervious) 89 92 94 95
Industrial Districts (72% impervious) 81 88 91 93
Residential c
Average lot size Average % Impervious d
1/8 acre or less 65 77 85 90 92
1/4 acre 38 61 75 83 87
1/3 acre 30 57 72 81 86
1/2 acre 25 54 70 80 85
1 acre 20 51 68 79 84
Paved Parking Lots, Roofs, Driveways e: 98 98 98 98
Streets and Roads:
Paved with curbs and storm sewers e 98 98 98 98
Gravel 76 85 89 91
Dirt 72 82 87 89
Paved with open ditches 83 89 92 93
Newly graded area (no vegetation established)f 77 86 91 94
a
For a more detailed description of agricultural land use curve numbers, refer to Table B-8. .
b
Good cover is protected from grazing and litter and brush cover soil.
c Curve numbers are computed assuming the runoff from the house and driveway is directed toward

the street with a minimum of roof water directed to lawns where additional infiltration could occur,
which depends on the depth and degree of the permeability of the underlying strata.
d The remaining pervious areas (lawn) are considered to be in good pasture condition for these

curve numbers.
e In some warmer climates of the country, a curve number of 96 may be used.
f Use for temporary conditions during grading and construction.

Note: These values are for Antecedent Moisture Condition II, and Ia = 0.2S.
Reference: USDA, SCS, TR-55 (1984).

Appendix B: Hydrology Design Aids B-7


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-8: SCS Runoff Curve Numbers for Agricultural Use

Treatment Hydrologic Hydrologic Soil Group


Land Use or Practice Condition A B C D

Fallow Straight row ----- 77 86 91 94

Row Crops Straight row Poor 72 81 88 91


Straight row Good 67 78 85 89
Contoured Poor 70 79 84 88
Contoured Good 65 75 82 86
and terraced Poor 66 74 80 82
and terraced Good 62 71 78 81

Small grain Straight row Poor 65 76 84 88


Straight row Good 63 75 83 87
Contoured Poor 63 74 82 85
Contoured Good 61 73 81 84
Contoured Good 55 69 78 83
and terraced Poor 61 72 79 82
and terraced Good 59 70 78 81

Close seeded legumesa Straight row Poor 66 77 85 89


or rotation meadow Straight row Good 58 72 81 85
Contoured Poor 64 75 83 85
Contoured Good 55 69 78 83
and terraced Poor 63 73 80 83
and terraced Good 51 67 76 80

Pasture or range Poor 68 79 86 89


Fair 49 69 79 84
Good 39 61 74 80
Contoured Poor 47 67 81 88
Contoured Fair 25 59 75 83
Contoured Good 6 35 70 79

Meadow Good 30 58 71 78

Woods Poor 45 66 77 83
Fair 36 60 73 79
Good 25 55 70 77

Farmsteads ----- 59 74 82 86
Road (dirt)b ----- 72 82 87 89
(hard surface)b ----- 74 84 90 92

aClosed-drilled or broadcast.
bIncluding right-of-way.
Note: These values are for Antecedent Moisture Condition II, and Ia = 0.2S.
Reference: USDA, SCS, NEH-4 (1972)

Appendix B: Hydrology Design Aids B-8


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-9: SCS Classifications of Vegetative Covers by Hydrologic Properties

Vegetative Cover Hydrologic Condition

Crop rotation Poor: Contains a high proportion of row crops,


small grain, and fallow.

Good: Contains a high proportion of alfalfa and


grasses.

Native pasture or range Poor: Heavily grazed or having plant cover on


less range than 50% of the area.

Fair: Moderately grazed; 50 - 75% plant cover.

Good: Lightly grazed; more than 75% plant cover.

Permanent Meadow: 100% plant cover.

Woodlands Poor: Heavily grazed or regularly burned so that


litter, small trees, and brush are destroyed.

Fair: Grazed but not burned; there may be some


litter.

Good: Protected from grazing so that litter and


shrubs cover the soil.
_____________
Reference: USDA, SCS, NEH-4 (1972).

Appendix B: Hydrology Design Aids B-9


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-10: USGS Regression Equations – Natural Flow Conditions - Region 1

Standard
Error of
Prediction
Peak Runoff Equation (%)
Q2 = 127 A0.656 (ST+1)-0.098 43

Q5 = 248 A0.662 (ST+1)-0.189 40

Q10 = 357 A0.666 (ST+1)-0.239 42

Q25 = 528 A0.671 (ST+1)-0.293 47

Q50 = 684 A0.675 (ST+1)-0.328 52

Q100 = 864 A0.679 (ST+1)-0.362 57

Q200 = 1072 A0.683 (ST+1)-0.392 62

Q500 = 1395 A0.688 (ST+1)-0.430 70

QT = Peak runoff rate for return period of T-years, in cfs


A = Drainage area, in miles2
ST = Basin storage, the percentage of the drainage basin occupied by lakes,
reservoirs, swamps, and wetland. In-channel storage of a temporary nature,
resulting from detention ponds or roadway embankments, is not included in the
computation of ST

Basin Characteristic Range of Applicability

Drainage Area (A) 0.14 miles2 (89.6 acres) to 4,385 miles2


Storage Area (ST) 0% to 44.29%

_____________
Reference: Verdi (2006)

See Figure B-4 for zone delineation.

Appendix B: Hydrology Design Aids B-10


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-11: USGS Regression Equations – Natural Flow Conditions - Region 2

Standard
Error of
Prediction
Peak Runoff Equation (%)

Q2 = 101 A0.617 (ST+1)-0.211 58

Q5 = 184 A0.620 (ST+1)-0.212 53

Q10 = 253 A0.621 (ST+1)-0.215 52

Q25 = 353 A0.621 (ST+1)-0.221 53

Q50 = 435 A0.621 (ST+1)-0.226 54

Q100 = 525 A0.621 (ST+1)-0.231 56

Q200 = 622 A0.621 (ST+1)-0.236 59

Q500 = 764 A0.620 (ST+1)-0.244 63

QT = Peak runoff rate for return period of T-years, in cfs


A = Drainage area, in miles2
ST = Basin storage, the percentage of the drainage basin occupied by lakes,
reservoirs, swamps, and wetland. In-channel storage of a temporary nature,
resulting from detention ponds or roadway embankments, is not included in the
computation of ST

Basin Characteristic Range of Applicability

Drainage Area (A) 0.06 miles2 (38.4 acres) to 2,647 miles2


Storage Area (ST) 0% to 74.33%

_____________
Reference: Verdi (2006)

See Figure B-4 for zone delineation.

Appendix B: Hydrology Design Aids B-11


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-12: USGS Regression Equations – Natural Flow Conditions - Region 3

Standard
Error of
Prediction
Peak Runoff Equation (%)

Q2 = 72.7 A0.741 (ST+1)-0.589 87

Q5 = 164 A0.704 (ST+1)-0.587 62

Q10 = 250 A0.686 (ST+1)-0.592 56

Q25 = 390 A0.668 (ST+1)-0.601 53

Q50 = 517 A0.656 (ST+1)-0.608 53

Q100 = 664 A0.646 (ST+1)-0.616 54

Q200 = 833 A0.638 (ST+1)-0.625 56

Q500 = 1094 A0.629 (ST+1)-0.638 59

QT = Peak runoff rate for return period of T-years, in cfs


A = Drainage area, in miles2
ST = Basin storage, the percentage of the drainage basin occupied by lakes,
reservoirs, swamps, and wetland. In-channel storage of a temporary nature,
resulting from detention ponds or roadway embankments, is not included in the
computation of ST

Basin Characteristic Range of Applicability

Drainage Area (A) 0.41 miles2 (262.4 acres) to 3,244 miles2


Storage Area (ST) 0.18% to 48.04%
_____________
Reference: Verdi (2006)

See Figure B-4 for zone delineation.

Appendix B: Hydrology Design Aids B-12


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-13: USGS Regression Equations – Natural Flow Conditions - Region 4

Standard
Error of
Prediction
Peak Runoff Equation (%)

Q2 = 171 A0.628 (ST+1)-0.401 36

Q5 = 321 A0.618 (ST+1)-0.395 39

Q10 = 447 A0.614 (ST+1)-0.396 43

Q25 = 636 A0.610 (ST+1)-0.401 48

Q50 = 797 A0.609 (ST+1)-0.406 53

Q100 = 975 A0.608 (ST+1)-0.411 57

Q200 = 1171 A0.608 (ST+1)-0.416 62

Q500 = 1461 A0.609 (ST+1)-0.424 69

QT = Peak runoff rate for return period of T-years, in cfs


A = Drainage area, in miles2
ST = Basin storage, the percentage of the drainage basin occupied by lakes,
reservoirs, swamps, and wetland. In-channel storage of a temporary nature,
resulting from detention ponds or roadway embankments, is not included in the
computation of ST

Basin Characteristic Range of Applicability

Drainage Area (A) 0.20 miles2 (120 acres) to 2,833 miles2


Storage Area (ST) 0% to 34.12%
_____________
Reference: Verdi (2006)

See Figure B-4 for zone delineation.

Appendix B: Hydrology Design Aids B-13


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-14: USGS Nationwide Regression Equations for Urban Conditions

Standard
Error
Peak Runoff Equation R2 (%)

UQ2 = 2.35A0.41 SL0.17 (i2 + 3)2.04 (ST + 8)-0.65 (13 - BDF)-0.32 IA0.15 RQ20.47 0.93 38

UQ5 = 2.70A0.35 SL0.16 (i2 + 3)1.86 (ST + 8)-0.59 (13 - BDF)-0.31 IA0.11 RQ50.54 0.93 37

UQ10 = 2.99A0.32 SL0.15 (i2 + 3)1.75 (ST + 8)-0.57 (13 - BDF)-0.30 IA0.09 RQ100.58 0.93 38

UQ25 = 2.78A0.31 SL0.15 (i2 + 3)1.76 (ST + 8)-0.55 (13 - BDF)-0.29 IA0.07 RQ250.60 0.93 40

UQ50 = 2.67A0.29 SL0.15 (i2 + 3)1.74 (ST + 8)-0.53 (13 - BDF)-0.28 IA0.06 RQ500.62 0.92 42

UQ100 = 2.50A0.29 SL0.15 (i2 + 3)1.76 (ST + 8)-0.52 (13 - BDF)-0.28 IA0.06 RQ1000.63 0.92 44

UQ500 = 2.27A0.29 SL0.16 (i2 + 3)1.86 (ST + 8)-0.54 (13 - BDF)-0.27 IA0.05 RQ5000.63 0.90 49

UQT = Peak discharge, in cfs, for the urban watershed for recurrence interval T.
SL = Main channel slope, in ft/mile, measured between points which are 10 and 85 percent of
the main channel length upstream from the study site. For sites where SL is greater than
70 ft/mile, 70 ft/mile is used in the equations.
A = Contributing drainage area, in miles2.
i2 = Rainfall intensity, in inches, for the 2-hour 2-year occurrence.
ST = Basin storage, the percentage of the drainage basin occupied by lakes, reservoirs,
swamps, and wetland. In-channel storage of a temporary nature, resulting from detention
ponds or roadway embankments, is not included in the computation of ST.
BDF = Basin development factor, an index of the prevalence of the drainage aspects of (a)
storm sewers, (b) channel improvements, (c) impervious channel linings, and (d) curb
and gutter streets. The range of BDF is 0-12. A value of zero for BDF indicates the
above drainage aspects are not prevalent, but does not necessarily mean the basin is
non-urban. A value of 12 indicates full development of the drainage aspects throughout
aspects throughout the basin. See Chapter 2, Section 2.2.3 & Example 2.2-2 of this
document for details of computing BDF.
IA = Percentage of the drainage basin occupied by impervious surfaces, such as houses,
buildings, streets, and parking lots.
RQT = Peak discharge, in cfs, for an equivalent rural drainage basin in the same hydrologic area
as the urban basin, and for recurrence interval T.
_____________
Reference: Sauer et al. (1983).

Appendix B: Hydrology Design Aids B-14


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-15: Urban Watershed Regression Equations for Tampa Bay Area

Standar
d
Error
Peak Runoff Equation R2 in %

Q2 = 3.72 A1.07 BDF1.05 SL0.77(DTENA + 0.01)-0.11 0.92 33

Q5 = 7.94 A1.03 BDF0.87 SL0.81 (DTENA + 0.01)-0.10 0.90 32

Q10 = 12.9 A1.04 BDF0.75 SL0.83 (DTENA + 0.01)-0.10 0.88 35

Q25 = 214 A1.13 (13 - BDF)-0.59 SL0.73 0.85 37

Q50 = 245 A1.14 (13 - BDF)-055 SL0.74 0.83 39

Q100 = 282 A0.918 (13- BDF)-0.51 SL0.76 0.83 42

QT = Peak runoff rate for return period of T-years, in cfs


A= Drainage area, in miles2
BDF = Basin development factor, dimensionless; see Example 2.2-2 and the
discussion on Nationwide Regression Equations in Chapter 2, Section 2.2.3
of this document.
SL = Channel slope, in ft/mile, measured between points at 10 and 85 percent of
the distance from the design point to the watershed boundary.
DTENA = Surface area of lakes, ponds, and detention and retention basins, expressed
as a percentage of drainage area.

Watershed Characteristic Range of Applicability

Drainage Area 0.34 miles2 (220 acres) to 3.45


miles2
Noncontributing internal drainage 0 to 0.3 percent of watershed area
Soil-infiltration index 2.05 to 3.89 inches
Total impervious area 19 to 61 percent of watershed area
Hydraulically connected impervious area 5.5 to 53 percent of watershed area
Effective impervious area 5.5 to 40 percent of watershed area
Channel slope 4.6 to 23.6 ft/mile
Lake and detention basin area 0 to 3.5 percent of watershed area
Basin development factor 3 to 12 (dimensionless)
_____________
Reference: Lopez and Woodham (1983).

Appendix B: Hydrology Design Aids B-15


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-16: Urban Watershed Regression Equations for Leon County, Florida

Standard
Error
Peak Runoff Equation R2 in %

Outside Lake Lafayette Basin Inside Lake Lafayette Basin

Q2 = 10.7 A0.766 IA1.07 Q2 (LL) = 1.71 A0.766 IA1.07 0.99 18

Q5 = 24.5 A0.770 IA0.943 Q5 (LL) = 4.51 A0.770 IA0.943 0.98 18

Q10 = 39.1 A0.776 IA0.867 Q10 (LL) = 7.98 A0.776 IA0.867 0.98 20

Q25 = 63.2 A0.787 IA0.791 Q25 (LL) = 14.6 A0.787 IA0.791 0.98 22

Q50 = 88.0 A0.797 IA0.736 Q50 (LL) = 22.1 A0.797 IA0.736 0.97 24

Q100 = 118 A0.808 IA0.687 Q100 (LL) = 32.4 A0.808 IA0.687 0.97 25

Q500 = 218 A0.834 IA0.589 Q500 (LL) = 71.7 A0.834 IA0.589 0.97 30

QT = Peak runoff rate outside Lake Lafayette Basin for return period T, in cfs.
A= Drainage area, in miles2
IA = Impervious area, in percentage of drainage area.
QT (LL) = Peak runoff rate inside Lake Lafayette Basin for return period T, in cfs.

Watershed Characteristic Range of Applicability

Drainage Area 0.26 miles2 (166 acres) to 15.9 miles2


Impervious area 5.8 to 54 %
Channel slope 11.9 to 128 ft/mile
Basin development factor 0 to 8 (dimensionless)
Main Channel Length 0.58 to 6.50 miles
Storage (area of ponds, lakes, swamps) 0 to 4.26 percent

_____________
Reference: Franklin and Losey (1984).

Appendix B: Hydrology Design Aids B-16


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-17: USGS Watershed Regression Equations for West Central Florida

_______________
Reference: Hammett and DelCharco (2001).

Appendix B: Hydrology Design Aids B-17


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-18: USGS Watershed Regression Equations’ Range of Applicability for


West Central Florida

_____________
Reference: Hammett and DelCharco (2001).

Appendix B: Hydrology Design Aids B-18


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-19: Department Intensity Duration Frequency (IDF) Regression Equation


Constants and Coefficients
(Page 1 of 3)

Polynomial Coefficients
for a Third Degree Polynomial
Rainfall Storm Frequency
Zone in Years A B C D
1 2 11.0983 -2.47240 0.00711 0.01886
1 3 11.97845 -2.67930 0.02444 0.01812
1 5 11.82413 -2.28931 -0.07735 0.02535
1 10 12.01819 -1.91394 -0.20146 0.03519
1 25 13.48736 -1.84775 -0.32753 0.04818
1 50 13.12334 -1.04283 -0.52846 0.06176

2 2 10.57745 -2.10106 -0.08181 0.02557


2 3 10.89437 -1.83103 -0.19244 0.03537
2 5 10.85901 -1.50267 -0.27902 0.04121
2 10 12.30743 -1.94991 -0.22855 0.03903
2 25 12.81040 -1.40033 -0.43207 0.05602
2 50 14.17099 -1.56750 -0.47317 0.06168

3 2 11.87566 -2.78202 0.02345 0.02058


3 3 11.40436 -2.01001 -0.18000 0.03550
3 5 11.42451 -1.65788 -0.29070 0.04438
3 10 11.51866 -1.25713 -0.41757 0.05430
3 25 11.30909 -0.30052 -0.70475 0.07704
3 50 12.16856 -0.12834 -0.82217 0.08822

4 2 12.75884 -3.55763 0.21171 0.00678


4 3 12.36825 -2.82718 0.00820 0.02248
4 5 11.81456 -2.18321 -0.14397 0.03283
4 10 12.54028 -2.13586 -0.20440 0.03866
4 25 12.76532 -1.45996 -0.42819 0.05666
4 50 14.56743 -2.19263 -0.30685 0.04897

5 2 12.89666 -3.55805 0.21227 0.00619


5 3 12.49905 -2.90429 0.04609 0.01794
5 5 12.28117 -2.34803 -0.11099 0.02995
5 10 13.68290 -2.93192 -0.00385 0.02241
5 25 12.69696 -1.22300 -0.49561 0.06173
5 50 13.36862 -0.83912 -0.66880 0.07724

Appendix B: Hydrology Design Aids B-19


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

(Page 2 of 3)

Polynomial Coefficients
for a Third Degree Polynomial
Rainfall Storm Frequency
Zone in Years A B C D
6 2 14.09519 -4.17207 0.31773 0.00029
6 3 14.98331 -4.44963 0.35683 -0.00224
6 5 14.54762 -3.89935 0.22564 0.00674
6 10 14.35386 -3.10140 -0.01003 0.02525
6 25 16.15961 -3.48135 -0.00160 0.02677
6 50 15.67671 -2.52635 -0.26055 0.04609

7 2 12.10821 -2.79255 0.02002 0.02053


7 3 12.43560 -2.56458 -0.06903 0.02787
7 5 12.51872 -2.17764 -0.19805 0.03849
7 10 12.49556 -1.67116 -0.34901 0.05017
7 25 12.92209 -1.11084 -0.55019 0.06666
7 50 13.29550 -0.70432 -0.70152 0.07933

8 2 11.51282 -2.10568 -0.16578 0.03515


8 3 11.13440 -1.44999 -0.34027 0.04808
8 5 11.41155 -1.34465 -0.38409 0.05149
8 10 11.54908 -0.89694 -0.53000 0.06319
8 25 10.92111 0.51710 -0.93480 0.09473
8 50 11.58787 0.73605 -1.04111 0.10384

9 2 11.08062 -1.66022 -0.28464 0.04453


9 3 11.54667 -1.49353 -0.35960 0.05071
9 5 11.76664 -1.38391 -0.39880 0.05352
9 10 12.08400 -1.00328 -0.53661 0.06491
9 25 12.38592 -0.27352 -0.77352 0.08370
9 50 14.16172 -0.73486 -0.75377 0.08518

10 2 11.33384 -1.86569 -0.22813 0.04005


10 3 11.32916 -1.38557 -0.36672 0.05012
10 5 11.19083 -0.93165 -0.48526 0.05836
10 10 10.84265 -0.18976 -0.69575 0.07495
10 25 11.83969 0.09353 -0.84451 0.08783
10 50 11.59208 1.00204 -1.10384 0.10762

Appendix B: Hydrology Design Aids B-20


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

(Page 3 of 3)

Polynomial Coefficients
for a Third Degree Polynomial
Rainfall Storm Frequency
Zone in Years A B C D
11 2 10.09256 -2.25031 0.01661 0.01544
11 3 9.30810 -1.21537 -0.25504 0.03590
11 5 9.02699 -0.47796 -0.46784 0.05263
11 10 10.23814 -1.23242 -0.27724 0.03685
11 25 11.68811 -1.61200 -0.25239 0.03706
11 50 9.94772 0.31312 -0.73271 0.07222

_____________
I = A + BX + CX2 + DX3 X = loge (time in minutes)

These equations were derived from the rainfall curves and are not exact representations
thereof. Appropriate values for X are 8 to 180 minutes.

Appendix B: Hydrology Design Aids B-21


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Table B-20: Example Application of Department IDF Regression Equations

EXAMPLE

Zone 6 - 50 years
_____________
I = A + BX + CX2 + DX3 X = loge (time in minutes)
I = 15.67671 - 2.52635X - 0.26055X2 + 0.04609X3

Time I (curve) I (calculated)


8 min 9.4 9.7
10 min 8.9 9.0
20 min 7.2 7.0
30 min 5.9 5.9
40 min 5.1 5.1
50 min 4.5 4.6
60 min 4.1 4.1
2 hr 2.67 2.7
3 hr 2.02 2.0
4 hr 1.65 1.59*
5 hr 1.40 1.34*
10 hr 0.87 0.92*
15 hr 0.65 0.94*
20 hr 0.54 1.09*
24 hr 0.47 1.25*

_____________
* These values are provided for comparison purposes only, since the regression
equations are not valid beyond a 3-hour period.

Appendix B: Hydrology Design Aids B-22


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Figure B-1: Kinematic Wave Formula for Determining Overland Flow Travel Time

Appendix B: Hydrology Design Aids B-23


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Figure B-2: Overland Flow Velocities for Various Land Use Types

Appendix B: Hydrology Design Aids B-24


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Equations and assumptions from Figure B-3

Ref: Chapter 15, Part 630, National Engineering Handbook, May 2010

Figure B-3: Velocity versus slope for Shallow Concentrated Flow

Appendix B: Hydrology Design Aids B-25


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

Figure B-4: Regions for USGS Regression Equations – Natural Flow Conditions

Appendix B: Hydrology Design Aids B-26


January 2019

Drainage Design Guide


Appendix B: Hydrology Design Aids

_____________
Reference: Hammett and DelCharco (2001).

Figure B-5: Regions for USGS Regression Equations for Natural Flow Conditions in
West Central Florida

Appendix B: Hydrology Design Aids B-27


January 2019

Drainage Design Guide


Appendix C: Open Channel Flow Design Aids

APPENDIX

C. OPEN CHANNEL FLOW DESIGN AIDS

Appendix C: Open Channel Flow Design Aids


January 2019

Drainage Design Guide


Appendix C: Open Channel Flow Design Aids

TABLE OF CONTENTS

C. Open Channel Flow Design Aids ......................................................................... C-1


Example C.1 – Geometric Elements......................................................................... C-1
Example C.2 – Geometric Elements......................................................................... C-2
Figure C-1: Trapezoidal Channel Geometry .......................................................... C-4
Figure C-2: Nomographs for the Solution of Manning’s Equation .......................... C-5
Figure C-3: Trapezoidal Channel Capacity Chart .................................................. C-6
Figure C-4: Open Channel Geometric Relationships for Various Cross Sections .. C-7

Appendix C: Open Channel Flow Design Aids C-i


January 2019

Drainage Design Guide


Appendix C: Open Channel Flow Design Aids

C. OPEN CHANNEL FLOW DESIGN AIDS


The nomographs in Figures C-1 through C-3 can be used as desktop aides for open
channel flow calculations. The purpose of each nomograph is:

Figure C-1 Area, Hydraulic Radius, and Top Width of Trapezoidal Channels
Figure C-2 Normal Depth Velocity for a General Cross Section
Normal Depth Velocity in a Circular Pipe
Figure C-3 Normal Depth in a Trapezoidal Channel

Figure C-1 can be used to solve Example C.1 below and the Geometry of Examples
3.1-1 through 3.1-4 in Chapter 3 of this document.

EXAMPLE C.1 – GEOMETRIC ELEMENTS

Given: Depth = 1.0 ft


Trapezoidal Cross Section shown below

1 1 ft 1
4 4
2.9 ft

Calculate: Area, Wetted Perimeter, Hydraulic Radius, Top Width, and Hydraulic
Depth

Water Area
A  a  bd  zd 2
a  (2.9 1)  4(1) 2  6.9 ft 2

Wetted Perimeter
P  b  2d z 2  1
P  2.9  (2  1) 4 2  1  11.146  11.1 ft

Hydraulic Radius
bd  zd 2
Rr
b  2d z 2  1

Appendix C: Open Channel Flow Design Aids C-1


January 2019

Drainage Design Guide


Appendix C: Open Channel Flow Design Aids

(2.9  1)  4(1) 2
r  0.619  0.62 ft
2.9  (2  1) 4 2  1

Top Width
T  b  2 zd
T  2.9  (2  4 1)  10.9 ft

Hydraulic Depth
A 6.9
D   0.63 ft
T 10.9

This problem can also be solved using nomographs

This example is solved in the lower right hand corner of Figure C-1

EXAMPLE C.2 – GEOMETRIC ELEMENTS

Determine Normal Depth for Standard Ditch and Narrow Ditch given in Chapter 3,
Example 3.1-4 using Figure C-3.

Standard Ditch:
Qn (25)(0.04)
Solve for 8 1
 8 1
 0.193
b 3S 2
5 3 (0.005) 2

The average value of z is (6 + 4) / 2 = 5

d
From Figure C-3,  0.22
b

d  0.43b  0.22(5)  1.1 ft.

Using a trial and error procedure to solve Manning’s Equation, normal depth = 1.12’

Narrow Ditch:

Qn (25)(0.04)
Solve for 8 1
 8 1
 0.501
b 3S 2
3.5 3 (0.005) 2

The average value of z is (6 + 4) / 2 = 5

Appendix C: Open Channel Flow Design Aids C-2


January 2019

Drainage Design Guide


Appendix C: Open Channel Flow Design Aids

d
From Figure C-3,  0.34
b

d  0.34b  0.34(3.5)  1.2 ft.

Using a trial and error procedure to solve Manning’s Equation, normal depth = 1.25’

Appendix C: Open Channel Flow Design Aids C-3


January 2019

Drainage Design Guide


Appendix C: Open Channel Flow Design Aids

Figure C-1: Trapezoidal Channel Geometry

Appendix C: Open Channel Flow Design Aids C-4


January 2019

Drainage Design Guide


Appendix C: Open Channel Flow Design Aids

Figure C-2: Nomographs for the Solution of Manning’s Equation

Appendix C: Open Channel Flow Design Aids C-5


January 2019

Drainage Design Guide


Appendix C: Open Channel Flow Design Aids

Figure C-3: Trapezoidal Channel Capacity Chart

Appendix C: Open Channel Flow Design Aids C-6


January 2019

Drainage Design Guide


Appendix C: Open Channel Flow Design Aids

Figure C-4: Open Channel Geometric Relationships for Various Cross Sections

Appendix C: Open Channel Flow Design Aids C-7


January 2019

Drainage Design Guide


Appendix D: Gutter Flow Using HEC-RAS

APPENDIX

D. GUTTER FLOW USING HEC-RAS

Appendix D: Gutter Flow Using HEC-RAS


January 2019

Drainage Design Guide


Appendix D: Gutter Flow Using HEC-RAS

TABLE OF CONTENTS

D. Gutter Flow Using HEC-RAS .......................................................................... D-1


D.1 Introduction .................................................................................................... D-1
D.2 Example of Gutter Flow Using HEC-RAS ...................................................... D-2

Appendix D: Gutter Flow Using HEC-RAS D-i


January 2019

Drainage Design Guide


Appendix D: Gutter Flow Using HEC-RAS

D. GUTTER FLOW USING HEC-RAS

D.1 INTRODUCTION
Gutter flow is a form of open channel flow. Most gutter flow is associated with pavement
drainage and storm drain design, and is, therefore, discussed in Storm Drain Design,
Chapter 6 of this document. Some situations may warrant a more detailed approach to
gutter flow than presented in the Storm Drain Chapter. The gutter flow equation is:

0.56 5 3 8 3 12
Q SX T S
n
where:

Q = Discharge, in ft3/sec

n = Manning's roughness coefficient

SX = Cross Slope, in ft/ft

T = Spread, in ft

S = Slope of the energy gradient, in ft/ft

The gutter flow equation is a normal depth equation that can be used in a manner
similar to Manning’s Equation. The slope of the energy gradient is the same as the
longitudinal slope of the gutter for normal depth of flow in the gutter. The equation
cannot be solved if the slope is zero or negative. While zero and negative slope
conditions should be avoided when designing a project, you will sometimes encounter
these conditions when analyzing existing or retrofit situations.

You can use the HEC-RAS model to analyze open channels with flat or reverse slopes,
so HEC-RAS is applicable to analyzing gutter flow with zero or negative slopes. In HEC-
RAS, the friction losses between cross sections are estimated using Manning’s
Equation. The Manning’s roughness coefficient can be adjusted, in effect, to make
HEC-RAS use the gutter flow equation to determine the friction losses.

If the gutter has a typical triangular cross section, such as a gutter against a curb or
barrier wall, the area and the hydraulic radius can be solved using the cross slope, S X,
and the spread, T:

Appendix D: Gutter Flow Using HEC-RAS D-1


January 2019

Drainage Design Guide


Appendix D: Gutter Flow Using HEC-RAS

SXT 2
A
2

P T

where:

P = Wetted perimeter, in ft

Note that T is an approximation of P when the cross slope is relatively small.

A SXT 2 SXT
R  
P 2T 2

Substituting into Manning’s Equation:

2
1.486  S X T 2  S X T  3 12 1.486 5 8 1 0.47 5 3 8 3 12
Q    S  S X 3T 3 S 2  SX T S
n  2  2 
2
(2)2 n3 n

Therefore, Manning’s Equation can be manipulated to solve the gutter flow equation if
the Manning’s roughness coefficient is reduced by a ratio of 0.47/0.56 = 0.84. The
roughness value normally used in gutter analysis is 0.016 (see Appendix B, Table B-2).
The reduced value that should be used in HEC-RAS is 0.0134 or 0.013.

D.2 EXAMPLE OF GUTTER FLOW USING HEC-RAS


An existing four-lane divided rural highway with zero percent grade will be widened to
six lanes by adding lanes in the median. The new inside lanes will slope toward the
median. A barrier wall will be erected in the median to prevent cross-over accidents.
The inside shoulder will be 12 feet wide with a 0.06 ft/ft cross slope.

The shoulder will not be warped to provide a grade along the barrier wall. Instead, the
water collecting against the barrier will be allowed to seek out the nearest inlet despite
the flat grade. Pipe will be installed parallel to the barrier wall to connect the inlets.
Occasionally, a pipe will be jacked and bored under the existing lanes to outfall the flow
from the median storm drain systems. The maximum distance between inlets will be
500 feet. Analyze the maximum spread next to the barrier wall.

Appendix D: Gutter Flow Using HEC-RAS D-2


January 2019

Drainage Design Guide


Appendix D: Gutter Flow Using HEC-RAS

D.2.1 Solution
The flow is assumed to divide halfway between the two inlets and flow in both
directions. So the flow from one side of the inlet comes from 250 feet away. Table D-1
shows the flow rate at each cross section that will be used in the HEC-RAS analysis.
Calculate the flow rates using the rational equation. Calculate the drainage area by
multiplying the width of 36 feet by the distance from the midway point between the
inlets. The rainfall intensity used is four inches per hour. The runoff coefficient is 0.95.

Table D-1: Discharges


Location Area Q
(distance from inlet) (acres) (cfs)
0 0.2066 0.785
1 0.2058 0.782
4 0.2033 0.773
10 0.1983 0.754
25 0.1860 0.707
50 0.1653 0.628
100 0.1240 0.471

The total flow into the inlet is 2 x 0.785 = 1.67 cfs. The capacity chart for a Type D DBI
from Appendix I (Inlet Efficiencies) shows that the depth above the inlet is less than 0.1
feet (which is a conservative estimate of the capacity of a barrier wall inlet). This depth
will be lower than critical depth, so the profile in HEC-RAS will start at critical depth.
Critical depth is not affected by the adjustment to Manning’s “n” because critical depth is
independent of the channel roughness.

The geometry of the shoulder next to the barrier wall is entered into HEC-RAS at
Station 0, which will be next to the inlet. See Figure D-1, below:

Appendix D: Gutter Flow Using HEC-RAS D-3


January 2019

Drainage Design Guide


Appendix D: Gutter Flow Using HEC-RAS

Figure D-1: HEC-RAS Input


The geometry is copied to the other desired cross section locations. Since the profile
begins at critical depth, the first few cross sections should be located close to each
other. The first cross section had to be located only one foot away to avoid a
conveyance ratio warning.

The flow data is entered. A flow of 0.01 cfs is entered at Station 250 because HEC-RAS
cannot use a value of zero when analyzing Steady State conditions. The downstream
boundary condition is set at critical depth. Figure D-2, below, shows the computed
profile:

Figure D-2: Computed Profile

Appendix D: Gutter Flow Using HEC-RAS D-4


January 2019

Drainage Design Guide


Appendix D: Gutter Flow Using HEC-RAS

The top width, which is equivalent to the spread, does not exceed six feet. Therefore,
the inlets prevent spread onto the travel lanes with a considerable safety factor.

Although spread will not be a problem, nuisance ponding will probably develop since the
elevation along the barrier will not be perfectly level. Although this will not be a hazard,
silt will collect next to the barrier and may require more maintenance.

Appendix D: Gutter Flow Using HEC-RAS D-5


January 2019

Drainage Design Guide


Appendix E: Partial Depth Pipe Flow Graphs

APPENDIX

E. PARTIAL DEPTH PIPE FLOW GRAPHS

Appendix E: Partial Depth Pipe Flow Graphs


January 2019

Drainage Design Guide


Appendix E: Partial Depth Pipe Flow Graphs

TABLE OF CONTENTS

Figure E-1: Circular Pipe Partial Flow Capacity Chart .................................................. E-1
Figure E-2: Circular Pipe Relative Flow, Area, Hydraulic Radius, and Velocity ............ E-2
Figure E-3: Horizontal Elliptical Pipe Relative Flow, Area, and Velocity ....................... E-2
Figure E-4: Vertical Elliptical Pipe Relative Flow, Area and Velocity ............................ E-3
Figure E-5: Pipe-Arch Relative Flow, Area, and Velocity ............................................. E-3

Appendix E: Partial Depth Pipe Flow Graphs E-i


January 2019

Drainage Design Guide


Appendix E: Partial Depth Pipe Flow Graphs

Qn
Q = Flow Rate (cfs)
D8 3 S 1 2
D = Pipe Diameter (ft)
S = Pipe Slope (ft/ft)
d = Normal Depth (ft)
Ref 1987 FDOT Drainage Manual
Figure E-1: Circular Pipe Partial Flow Capacity Chart

Appendix E: Partial Depth Pipe Flow Graphs E-1


January 2019

Drainage Design Guide


Appendix E: Partial Depth Pipe Flow Graphs

Figure E-2: Circular Pipe Relative Flow, Area, Hydraulic Radius, and Velocity

Figure E-3: Horizontal Elliptical Pipe Relative Flow, Area, and Velocity

Appendix E: Partial Depth Pipe Flow Graphs E-2


January 2019

Drainage Design Guide


Appendix E: Partial Depth Pipe Flow Graphs

Figure E-4: Vertical Elliptical Pipe Relative Flow, Area and Velocity

Figure E-5: Pipe-Arch Relative Flow, Area, and Velocity

Appendix E: Partial Depth Pipe Flow Graphs E-3


January 2019

Drainage Design Guide


Appendix F: Application Guidelines for Pipe End Treatments

APPENDIX

F. APPLICATION GUIDELINES FOR PIPE END TREATMENTS

Appendix F: Application Guidelines for Pipe End Treatments


January 2019

Drainage Design Guide


Appendix F: Application Guidelines for Pipe End Treatments

TABLE OF CONTENTS

Table F-1: Application Guidelines for Pipe End Treatments - Part A ............................F-1
Table F-2: Application Guidelines for Pipe End Treatments - Part B ............................F-2

Appendix F: Application Guidelines for Pipe End Treatments F-i


January 2019

Drainage Design Guide


Appendix F: Application Guidelines for Pipe End Treatments

Table F-1: Application Guidelines for Pipe End Treatments - Part A


(see page F-3 for notes)

Standard Description Application Inlet End


Plan
Index Cross Side Hydraulic
Type Pipe Size Median Application Ke
Drain Drain Performance

U Type Concrete
430-010 Single 15" thru 30" Limited Limited Yes Yes Fair 0.7
With Grate

430-011 U Type Concrete Single 15" thru 30" Limited No Yes Limited Good 0.5 to 0.7

Concrete Energy
430-012 Single 30" thru 72" Limited No No No NA NA
Dissipater

Flared End Single


430-020 Yes No Yes Yes Good 0.5
Section Concrete 12" thru 72"

Cross Drain
Single & Multiple
430-021 Mitered End Yes No Yes Yes Fair 0.7
15” thru 72"
Section

0.7 w/o,
Side Drain Mitered Single & Multiple
430-022 No Yes No Yes Fair 1.0 w/
End Section 15” thru 60"
grate

Single &Multiple (a)


430-030 Straight Concrete No Limited Yes Excellent 0.2
15" thru 54" Yes

430-031 Straight Concrete Single & Double 60" Yes No Limited Yes Excellent 0.2

430-032 Straight Concrete Single & Double 66" Yes No Limited Yes Excellent 0.2

430-033 Straight Concrete Single & Double 72" Yes No Limited Yes Excellent 0.2

430-034 Straight Concrete Single 84" Yes No Limited Yes Excellent 0.2

430-040 Winged Concrete Single 12" thru 48" Yes No Yes Yes Very Good 0.3

Appendix F: Application Guidelines for Pipe End Treatments F-1


January 2019

Drainage Design Guide


Appendix F: Application Guidelines for Pipe End Treatments

Table F-2: Application Guidelines for Pipe End Treatments - Part B


(see page F-3 for notes)

Standard Description Outlet End Safety


Economic
Plans Erosion Permitted Traffic-Safe
Index Type Pipe Size Applicable Rating
Tolerance Location Grate Available
U Type
Single
430-010 Concrete With Yes Very Good Inside CZ Required Good
15" thru 30"
Grate
U Type Single Grate Required
430-011 Yes Good Yes Fair
Concrete 15" thru 30" Inside CZ
Concrete
Single
430-012 Energy Yes Excellent Outside CZ No NA
30" thru 72"
Dissipater
Flared End
Single (c) (c)
430-020 Section Yes No Very Good
12" thru 72" Very Good Outside CZ
Concrete
Cross Drain
Single & Multiple (d)
430-021 Mitered End Yes Good No Very Good
15” thru 72" Outside CZ
Section
Side Drain
Single & Multiple (e)
430-022 Mitered End Yes Good Yes Good
15” thru 60" Inside CZ
Section

Straight Single & Multiple


430-030 Limited Good Outside CZ No Fair
Concrete 15" thru 54"
Straight
430-031 Single & Double 60" Limited Good Outside CZ No Fair
Concrete
Straight
430-032 Single & Double 66" Limited Good Outside CZ No Fair
Concrete
Straight
430-033 Single & Double 72" Limited Good Outside CZ No Fair
Concrete
Straight
430-034 Single 84" Limited Good Outside CZ No Fair
Concrete
Winged Single
430-040 Yes Good Outside CZ No Good
Concrete 12" thru 48"

Appendix F: Application Guidelines for Pipe End Treatments F-2


January 2019

Drainage Design Guide


Appendix F: Application Guidelines for Pipe End Treatments

LEGEND:
(a) For back of sidewalk location, see Standard Plans, Index 425-060.
(b) For temporary construction or use on a minor facility.
(c) Construction of optional toewall and concrete jacket may be necessary. Flared end
section sizes 12 inch and 15 inch may be located as close as 8 feet beyond the
outside edge of the shoulder.
(d) Mitered end section sizes 15 inch, 18 inch, and 24 inch may be located as close
as 8 feet beyond the outside edge of the shoulder.
(e) Mitered end section size 30 inch and larger does not require a grate if pipe is
located outside CZ and is offset from approach ditch alignment.

NOTES:
1. All end treatments must be selected to satisfy hydraulic suitability with proper
consideration given to safety and economics.
2. CZ denotes clear zone; it was formerly CRA, denoting clear recovery area.
3. Grates should not be placed on outlet ends unless positive debris protection is
provided at inlet end.
4. Additional notes concerning application restrictions may be shown on individual
indexes.
5. Economic ratings are based on statewide average costs.
6. End treatments with a Ke of 0.5 or greater should be used only in areas of low
design velocities and negligible debris.
7. Pipe sizes are circular, Class III B wall, concrete pipe. Elliptical pipe and corrugated
pipe are to be checked for fit. Metal pipe sizes should be reviewed using 2 ⅔-inch
x ½-inch corrugation up to 30 inches and 3-inch x 1-inch corrugation for larger
sizes.

Appendix F: Application Guidelines for Pipe End Treatments F-3


January 2019

Drainage Design Guide


Appendix G: Risk Evaluations

APPENDIX

G. RISK EVALUATIONS

Appendix G: Risk Evaluations


January 2019

Drainage Design Guide


Appendix G: Risk Evaluations

TABLE OF CONTENTS

G. Risk Evaluations ................................................................................................ G-1


G.1 Risk Evaluation ............................................................................................... G-1
G.1.1 Risk Assessment .......................................................................................... G-1
G.1.2 Economic Analysis ........................................................................................ G-2
G.2 Sample Risk Analysis .................................................................................... G-3
G.2.1 Alternatives Considered ................................................................................ G-3
G.2.1.1 Calculations for Alternative 1 ................................................................. G-4
G.2.1.2 Calculations for Alternative 2 ................................................................. G-7
G.2.1.3 Calculations for Alternative 3 ................................................................. G-8

Appendix G: Risk Evaluations G-i


January 2019

Drainage Design Guide


Appendix G: Risk Evaluations

G. RISK EVALUATIONS
G.1 Risk Evaluation
All designs with floodplain encroachments should include an evaluation of the inherent
flood-related risks to the highway facility and to the surrounding property. In the traditional
design process, the level of risk is seldom quantified, but is instead implied through the
application of predetermined design standards. For example, the design frequency,
backwater limitations, and limiting velocity are parameters for which design standards can
be set.

Two other approaches, however, are available that quantify risk on projects involving
highway facilities designed to encroach within the limits of a floodplain. These are risk
assessment and economic analysis.

Consideration of capital costs and risks should include, as appropriate, a risk analysis or
risk assessment that includes:

 The overtopping flood or the base flood, whichever is greater

 The greatest flood that must pass through the highway drainage structure(s),
where overtopping is not practicable

G.1.1 Risk Assessment


A risk assessment is a subjective analysis of the risks engendered by various design
alternatives, without detailed quantification of flood risks and losses. It may consist of
developing the construction costs for each alternative and subjectively comparing the
risks associated with each alternative. A risk assessment usually is more appropriate for
small structures or for structures whose size is highly influenced by non-hydraulic
constraints. There are no well-defined procedures or criteria for performing risk
assessments. However, an attempt should be made to screen projects and determine the
level of analysis required. Some items to consider are:

 Backwater
a. Is the overtopping flood greater than the design flood (100-year)?
b. Is the overtopping flood greater than the check flood (500-year)?
c. Is there potential for major flood damage from the overtopping flood?
d. Could flood damage occur even if the roadway crossing wasn't there?
e. Could flood damage be significantly increased by the backwater caused

Appendix G: Risk Evaluations G-1


January 2019

Drainage Design Guide


Appendix G: Risk Evaluations

by the proposed structure?


f. Could flood damage occur to offsite property owners?

 Traffic-Related Losses
a. If the design flood is exceeded and the roadway is overtopped, is there
a detour available?
 Roadway and/or Structure Repair Costs
a. Is the overtopping flood greater or less than the design flood (100-year)?
b. Is the embankment constructed from erosion-resistant material, such as
a clay type soil?
c. Does the embankment have good erosion-resistant vegetation cover?
d. How long will the duration of overtopping be?
e. Will the cost of protecting the roadway and/or structure from damage
exceed the cost of providing a relief structure?
f. Is there damage potential to the structure caused by scour, debris, or
other means during the lesser of the overtopping flood or the design
flood (100-year)?
If the risk assessment indicates the risks warrant additional study, a detailed analysis of
alternative designs and associated costs is necessary to determine the design with the
least total expected cost (LTEC) to the public.

G.1.2 Economic Analysis


An economic analysis (sometimes called risk analysis) encompasses a complete
evaluation of all quantifiable flood losses and the costs associated with them for each
structure alternative. This can include damage to structures, embankments, surrounding
property, traffic-related losses, and scour or stream channel change.

The level of expense and effort required for an economic analysis is considerably higher
than for a risk assessment, and selection of the process to be used should be based on
the size of the project and the potential risk involved.

Further details of the economic analysis process and procedures for using it have been
documented in HEC-17 (USDOT, FHWA, 1981). The full-scale detailed risk analysis
described in HEC-17 would not be necessary for normal stream crossings, but would
apply to unusual, complex, or high-cost encroachments involving substantial flood losses.

An example of a simple risk analysis follows.

Appendix G: Risk Evaluations G-2


January 2019

Drainage Design Guide


Appendix G: Risk Evaluations

G.2 SAMPLE RISK ANALYSIS

An existing double 10-foot x 4-foot concrete box culvert (CBC) crossing is the subject for
this analysis.

G.2.1 Alternatives Considered


Alternative 1: Extend existing double 10-foot x 4-foot CBC (60 feet total length) with no
change to road. Overtops at about a 17-year frequency; flooding at the site has not
caused any accidents.
Alternative 2: New quad 10-foot x 5-foot CBC (60 feet total length). Raise road to meet
FDOT 50-year HW (Headwater) criteria and closely match existing 100-year HW.
Overtops at frequencies greater than 50 years.
Alternative 3: Bridge

Table G-1: Alternatives Data


Alternative 1 Alternative 2 Alternative 3
Annual Capital Costs $ (i.e., Construction
Costs)

Annual Risks Costs $

Total Costs $

Appendix G: Risk Evaluations G-3


January 2019

Drainage Design Guide


Appendix G: Risk Evaluations

G.2.1.1 Calculations for Alternative 1

(A) Capital Costs

Quantities are from the Department’s Culvert Design Program.

Extend 20 feet right Concrete 43.1 CY Steel 6,622 lbs


Extend 8 feet left 23.5 CY 3,283 lbs
Total quantity Concrete 66.6 CY Steel 9,905 lbs

Unit prices $477/CY $0.53/lb

Total capital cost = $37,018 = $31,768 + $5,250


To convert to annual capital cost, use capital recovery factor (CRF) based on a discount
rate of 7 percent and a 20-year design life.

CRF  i where: n = 20 and i = 0.07


1 (1 i) n

Annual capital costs = $37,018 x 0.0944 = $3,494

(B) Additional Economic Costs

The following discussion estimates the additional losses associated with extending the
existing culvert and allowing the road to overtop. The losses usually consist of
embankment (and pavement), backwater, and traffic.

No embankment losses are expected. The existing road and culvert overtop, but there
is no history of embankment or pavement loss.

There will not be any additional backwater losses compared to Alternative 2. Both
Alternative 1 and Alternative 2 have essentially the same backwater characteristics.

There may be additional traffic losses associated with Alternative 1 when compared
with Alternative 2, which would raise the road to reduce overtopping potential. Traffic-
related costs consist of running time costs, lost time costs, and accident costs.
Running time costs were estimated, lost time costs were ignored (detour length added
only 1 mile to the travel distance), and accident costs were estimated but were found
to be insignificant.

Appendix G: Risk Evaluations G-4


January 2019

Drainage Design Guide


Appendix G: Risk Evaluations

Assume traffic would have to be detoured:

1 day for 25-year storm event (roadway tops at about a 17-year event)
2 days for 50-year storm event
3 days for 100-year storm event
4 days for 200-year storm event

The additional detour distance is 0.5 mile on a two-lane undivided roadway and 0.5
mile on a four-lane divided roadway.

Additional running costs = Cost per mile x ADT x additional detour length (miles)
Assume cost per mile = $0.35/mile

$25 yr = $0.35 x 27250 vpd x 1.0 mi x 1 day = $9,538

$50 yr = $0.35 x 27250 vpd x 1.0 mi x 2 days = $19,075

$100 yr = $0.35 x 27250 vpd x 1.0 mi x 3 days = $28,615

$200 yr = $0.35 x 27250 vpd x 1.0 mi x 4 days = $38,150

Additional accident costs: These are additional costs due to increased travel distance
due to the need to detour.

Additional detour length is 0.5 mi on a two-lane undivided roadway and 0.5 mi on a


four-lane divided roadway.

Accident cost = crash rate x vehicle miles x cost per crash

Vehicle miles = ADT x additional detour distance x number of days of detour

Get the crash rate and the cost per crash from the FDOT Safety Office.

Crash rate = 1.9 crashes/million vehicle miles for urban, two-lane, undivided
roadways
0.8 crashes/million vehicle miles for urban, four-lane, divided
roadways
Cost per crash = $28,000 for urban, two-lane, undivided roadways
$26,000 for urban, four-lane, divided roadways

$25 = ($28,000 x [27,250 x 0.5 x 1] x 1.9) + ($26,000 x [27,250 x 0.5 x 1] x 0.8)


$25 = $1,008.25

Appendix G: Risk Evaluations G-5


January 2019

Drainage Design Guide


Appendix G: Risk Evaluations

Using the same method, with 50-year detour = 2 days, 100-year detour = 3 days,
and 200-year detour = 4 days:

$50 = $2,016.50
$100 = $3,024.75
$200 = $4,033.00

Traffic losses in the following table are the sum of increased running costs and
increased accident losses.

Table G-2: Summary of Economic Losses


Losses ($)
Frequency (yr) Embankment & Total Losses ($)
Backwater Traffic
Pavement
5 0 0 0 0
10 0 0 0 0
15 0 0 0 0
9,538 + 1,008.25 =
25 0 0 10, 546.25
10,546.25
50 0 0 21,091.50 21,091.50
100 0 0 31,639.75 31,639.75
200 0 0 42,183.00 42,183.00

Appendix G: Risk Evaluations G-6


January 2019

Drainage Design Guide


Appendix G: Risk Evaluations

Table G-3: Summary of Annual Risk Costs


Exceed.
Freq. (yr) Losses ($) Average Loss ($) Delta Prob. Annual Risk Costs ($)
Prob.
5 0.2 0

10 0.1 0

15 0.07 0

5,273.13 0.03 158.19

25 0.04 10,546.25

15,818.88 0.02 316.38

50 0.02 21,091.50

26,365.63 0.01 263.66

100 0.01 31,639.75

36,911.38 0.005 184.56

200 0.005 42,183.00

42,183.00 0.005 210.92

0 42,183.00

Total Annual Risk Costs 1,133.71

G.2.1.2 Calculations for Alternative 2

Replace with quad 10’ x 5’ CBC

(A) Capital Costs

Concrete (from box culvert program) = 219.7 CY @ $477/CY = $104,797


Steel (from box culvert program) = 42,251 lbs @ $0.53/lb = $22,393

(B) Rebuild 400’ of Roadway

Structural Course (2’ x 24’) = 1,067 SY @ $3.40/SY = $3,628


Base group 9 = 1,067 SY @ $6.16/SY = $6,573
Neglect earthwork costs

Total Capital Costs = $137,391

Annual Capital Cost = Total x CRF = $12,970

Appendix G: Risk Evaluations G-7


January 2019

Drainage Design Guide


Appendix G: Risk Evaluations

This alternative would overtop at frequencies greater than the 50-year storm event and
would, therefore, have some annual risk costs. These risks were not calculated because
the annual cost alone is greater than the total cost for Alternative 1. If the capital costs for
Alternative 2 were less than the total cost for Alternative 1, it would be necessary to
calculate the other costs associated with this alternative.

G.2.1.3 Calculations for Alternative 3

57-foot-long x 44-foot-wide flat slab bridge

(A) Capital Costs

57 feet x 44 feet x $40/sf = 2,508 sf x $40/sf = $100,320

Annual cost using CRF = 0.0944 = $9,470

(B) Costs Not Estimated

Roadway fill and new base and asphalt. At a minimum, 900 feet of roadway would
have to be rebuilt to raise the grade to meet the bridge. (Bridge would be raised to
meet FDOT drift clearance requirements.)

Standard 1H:2V front slopes encroach into roadside ditches. Since the upstream
roadside ditch conveys substantial flow, it may not be possible or wise to reduce
its capacity. Vertical walls and/or additional right of way may be necessary.

Miscellaneous factors include driveway connections within the raised roadway


section, and the aesthetics of the raised road and bridge.

Table G-4: Cost Comparisons


Cost Item Alternative 1 Alternative 2 Alternative 3
Annual Capital Costs (i.e., Construction
$3,494 $12,970 $9,740
Costs)
Annual Risks Costs $1,134 0 0
Total Costs $4,628 $12,970 $9,740

Alternative 1 is the most economical alternative and the most desirable when considering
other impacts.

Appendix G: Risk Evaluations G-8


January 2019

Drainage Design Guide


Appendix H: Shoulder Gutter Transition Slope at Bridges

APPENDIX

H. SHOULDER GUTTER TRANSITION SLOPE AT BRIDGES

Appendix H: Shoulder Gutter Transition Slope at Bridges


January 2019

Drainage Design Guide


Appendix H: Shoulder Gutter Transition Slope at Bridges

H. SHOULDER GUTTER TRANSITION SLOPE AT BRIDGES

H.1 SLOPE CREATED BY THE SHOULDER/GUTTER TRANSITION

If the profile grade line (PGL) of the road is flat, there will be a slope away from the
bridge created by the shoulder/gutter transition. The degree of slope will depend on the
width of the shoulder and the cross slopes of the bridge deck and the roadway shoulder.
Figure H.1 shows a transition with a 10-foot shoulder and standard cross slopes for the
bridge deck and roadway shoulder.

The drop from the edge of the travel lane to the bottom of the gutter at the end of the
bridge barrier wall is:

0.02 (10.33) = 0.206 feet

Distance from edge of travel lane to bottom of gutter


Shoulder cross slope

Appendix H: Shoulder Gutter Transition Slope at Bridges H-1


January 2019

Drainage Design Guide


Appendix H: Shoulder Gutter Transition Slope at Bridges

10’
4”
8’

Figure H-1: Shoulder/Gutter Transition at Bridge End

The drop from the edge of the travel lane to the bottom of the gutter at the end of the
transition is:

0.06 (8) + 0.25 = 0.730 feet

3-inch drop from lip of gutter to bottom of gutter


Distance from edge of travel lane to lip of gutter
Shoulder cross slope

The drop of the gutter bottom in the transition is 0.730 – 0.206 = 0.524 feet. The length
of the transition is 25 feet. The slope of the bottom of the gutter is 0.524/25 = 0.0210, or
2.10%.

Appendix H: Shoulder Gutter Transition Slope at Bridges H-2


January 2019

Drainage Design Guide


Appendix I: Inlet Efficiencies

APPENDIX

I. INLET EFFICIENCIES

Appendix I: Inlet Efficiencies


January 2019

Drainage Design Guide


Appendix I: Inlet Efficiencies

TABLE OF CONTENTS

Figure I-1: Type 1, SX = 0.02 ......................................................................................... I-1


Figure I-2: Type 1, SX = 0.03 ......................................................................................... I-1
Figure I-3: Type 1, SX = 0.04 ......................................................................................... I-2
Figure I-4: Type 1, SX = 0.05 ......................................................................................... I-2
Figure I-5: Type 1, SX = 0.06 ......................................................................................... I-3
Figure I-6: Type 3, SX = 0.02 ......................................................................................... I-3
Figure I-7: Type 3, SX = 0.03 ......................................................................................... I-4
Figure I-8: Type 3, SX = 0.04 ......................................................................................... I-4
Figure I-9: Type 3, SX = 0.05 ......................................................................................... I-5
Figure I-10: Type 3, SX = 0.06 ....................................................................................... I-5
Figure I-11: Type 5, SX = 0.02 ....................................................................................... I-6
Figure I-12: Type 5, SX = 0.03 ....................................................................................... I-6
Figure I-13: Type 5, SX = 0.04 ....................................................................................... I-7
Figure I-14: Type 5, SX = 0.05 ....................................................................................... I-7
Figure I-15: Type 5, SX = 0.06 ....................................................................................... I-8
Figure I-16: Type S (Shoulder Gutter Inlet), SX = 0.06 .................................................. I-8
Figure I-17: Sump Conditions for Types 2, 4 & 6; SX = 0.02 ......................................... I-9
Figure I-18: Sump Conditions for Types 2, 4 & 6: SX = 0.04 ....................................... I-10
Figure I-19: Sump Conditions for Types 2, 4, 6 & S; SX = 0.06 ................................... I-11
Figure I-20: Type 9 Inlet .............................................................................................. I-12
Figure I-21: Type 10 Inlet ............................................................................................ I-12
Figure I-22: Closed Flume Inlet, SX = 0.02 .................................................................. I-13
Figure I-23: Closed Flume Inlet, SX = 0.03 ................................................................. I-13
Figure I-24: Closed Flume Inlet, SX = 0.04 .................................................................. I-14
Figure I-25: Closed Flume Inlet, SX = 0.05 .................................................................. I-14
Figure I-26: Closed Flume Inlet, SX = 0.06 .................................................................. I-15
Figure I-27: Ditch Bottom Inlets ................................................................................... I-16

Appendix I: Inlet Efficiencies I-i


January 2019

Drainage Design Guide


Appendix I: Inlet Efficiencies

Figure I-1: Type 1, SX = 0.02

Figure I-2: Type 1, SX = 0.03

Appendix I: Inlet Efficiencies I-1


January 2019

Drainage Design Guide


Appendix I: Inlet Efficiencies

Figure I-3: Type 1, SX = 0.04

Figure I-4: Type 1, SX = 0.05

Appendix I: Inlet Efficiencies I-2


January 2019

Drainage Design Guide


Appendix I: Inlet Efficiencies

Figure I-5: Type 1, SX = 0.06

Figure I-6: Type 3, SX = 0.02

Appendix I: Inlet Efficiencies I-3


January 2019

Drainage Design Guide


Appendix I: Inlet Efficiencies

Figure I-7: Type 3, SX = 0.03

Figure I-8: Type 3, SX = 0.04

Appendix I: Inlet Efficiencies I-4


January 2019

Drainage Design Guide


Appendix I: Inlet Efficiencies

Figure I-9: Type 3, SX = 0.05

Figure I-10: Type 3, SX = 0.06

Appendix I: Inlet Efficiencies I-5


January 2019

Drainage Design Guide


Appendix I: Inlet Efficiencies

Figure I-11: Type 5, SX = 0.02

Figure I-12: Type 5, SX = 0.03

Appendix I: Inlet Efficiencies I-6


January 2019

Drainage Design Guide


Appendix I: Inlet Efficiencies

Figure I-13: Type 5, SX = 0.04

Figure I-14: Type 5, SX = 0.05

Appendix I: Inlet Efficiencies I-7


January 2019

Drainage Design Guide


Appendix I: Inlet Efficiencies

Figure I-15: Type 5, SX = 0.06

Figure I-16: Type S (Shoulder Gutter Inlet), SX = 0.06

Appendix I: Inlet Efficiencies I-8


January 2019

Drainage Design Guide


Appendix I: Inlet Efficiencies

Figure I-17: Sump Conditions for Types 2, 4 & 6; SX = 0.02

Appendix I: Inlet Efficiencies I-9


January 2019

Drainage Design Guide


Appendix I: Inlet Efficiencies

Figure I-18: Sump Conditions for Types 2, 4 & 6: SX = 0.04

Appendix I: Inlet Efficiencies I-10


January 2019

Drainage Design Guide


Appendix I: Inlet Efficiencies

Figure I-19: Sump Conditions for Types 2, 4, 6 & S; SX = 0.06

Appendix I: Inlet Efficiencies I-11


January 2019

Drainage Design Guide


Appendix I: Inlet Efficiencies

Figure I-20: Type 9 Inlet

Figure I-21: Type 10 Inlet

Appendix I: Inlet Efficiencies I-12


January 2019

Drainage Design Guide


Appendix I: Inlet Efficiencies

Reference: Muste, Craig, Bayraktar

Figure I-22: Closed Flume Inlet, SX = 0.02

Reference: Muste, Craig, Bayraktar

Figure I-23: Closed Flume Inlet, SX = 0.03

Appendix I: Inlet Efficiencies I-13


January 2019

Drainage Design Guide


Appendix I: Inlet Efficiencies

Reference: Muste, Craig, Bayraktar

Figure I-24: Closed Flume Inlet, SX = 0.04

Reference: Muste, Craig, Bayraktar

Figure I-25: Closed Flume Inlet, SX = 0.05

Appendix I: Inlet Efficiencies I-14


January 2019

Drainage Design Guide


Appendix I: Inlet Efficiencies

Reference: Muste, Craig, Bayraktar

Figure I-26: Closed Flume Inlet, SX = 0.06

Appendix I: Inlet Efficiencies I-15


January 2019

Drainage Design Guide


Appendix I: Inlet Efficiencies

D I T C H B O T T O M I N L E T C A P A CI T Y
Fo r C o n d it io n s W h e re th e H G in th e Bo x is B e lo w th e G ra te

30

T yp e H re t .
20
T yp e G .
FLOW (cfs)

T yp e D & E re t.
T yp e H c .i .
T yp e E c . i.
T yp e F
T yp e C re t .
10
T yp e C c . i.

0
0 0.5 1 1.5 2
HEAD (ft) above Grate Top

1. The above graph should be used where the hydraulic gradient in the inlet box is below
the top of the grate. For other conditions, see the discussion below.
2. The above is based on 50% debris blockage and inlets without slots.
3. The symbols on the curves do not represent-measured data points. They are
calculated points from the equation and coefficients in the research report by the
University of South Florida titled “Investigation of Discharge through Grated Inlets”,
February 1993, WPI No. 0510611. Contact the FDOT Research Center at 850-414-
4615 to obtain a copy. The grate flow areas used in the equations are from U.S.
Foundry & Mfg. Corp.

Figure I-27: Ditch Bottom Inlets

Where the hydraulic gradient is above the top of the grate, the system capacity may
control the flow through the grate. The total system loss is a sum of friction losses
and various minor losses, including the loss across the grate. In this situation, the
loss across the grate is typically small but can be calculated from:

2
Vg
Head Loss [feet] = K
2g
Vg2
K
2g

Appendix I: Inlet Efficiencies I-16


January 2019

Drainage Design Guide


Appendix I: Inlet Efficiencies

Where K = 0.46 for reticuline grates; 3.2 for cast iron grates
Vg = velocity [fps] across the grate based on the grate full face area
(grate width x grate length)
g = acceleration of gravity

Example:
A DBI is needed to capture 5 cfs in a depressed area behind the sidewalk. The
hydraulic gradient due to friction loss in the system is estimated to be 0.8 ft above the
grate.

Try a Type C DBI:


Full Face Area = 2.33' x 3.0' = 7.0 ft2,
then Vg =Q / A = 5 / 7 = 0.7 fps
Assume a Cast iron grate is used. Then K = 3.2
Then: Head loss = K x Vg2/2g = 3.2 x (0.7) 2 / 64.4 = 0.02 ft

This is an insignificant amount of head loss and is typical of most design situations.
Where a DBI accepts high flow rates (usually under high head conditions) as perhaps
in a stormwater pond, the additional loss could be substantial and may dictate a larger
inlet (large grate area.)

Appendix I: Inlet Efficiencies I-17


January 2019

Drainage Design Guide


Appendix J: Exfiltration Design

APPENDIX

J. EXFILTRATION DESIGN - HYDRO-GEOTECHNICAL AIDS -


K FOR SATURATED AND COMPACTED LAB SOIL
SPECIMENS

Appendix J: Exfiltration Design


January 2019

Drainage Design Guide


Appendix J: Exfiltration Design

Table J-1: Coefficient of Permeability (k) for Saturated and Compacted Laboratory
Soil Specimens
SOIL TYPICAL SOIL CLASSIFICATION PERMEABILITY
NAME UNIFIED AASHTO (ft/day)
Well-graded gravels or gravel- GW A-3 300 to 0.3
sand mixtures with little or no Pervious
fines
Poorly graded gravels or gravel- GP A-3 3 x 104 to 30
sand mixtures with little or no Very pervious
fines
Silty gravels, gravel-sand-silt GM A-2-4 3 to 3 x 10-3
mixtures Semi-pervious to
pervious
Clayey gravels, gravel-sand-clay GC A-2-6 3 x 10-3 to 3 x 10-5
mixtures Impervious
Well-graded sands or gravelly SW A-3 30 to 0.3
sands with little or no fines Pervious
Poorly graded sands or gravelly SP A-3 300 to 3
sands with little or no fines Pervious
Silty sands, sand-silt mixtures SM A-2-4 3 to 3 x 10-3
Semi-pervious to
pervious
Clayey sands, sand-clay mixtures SC A-6 3 x 10-3 to 3 x 10-5
Impervious
Inorganic silts and very fine ML A-6 3 to 3 x 10-3
sands, rock flour, silty or clayey Semi-pervious to
fine sands or clayey silts with pervious
slightly plasticity
Inorganic clays of low to medium CL A-7 3 x 10-3 to 3 x 10-5
plasticity, gravely clays, sandy Impervious
clays, silty clays, lean clays
Organic silts and organic silty OL A-6 0.3 to 3 x 10-3
clays of low plasticity Semi-pervious to
pervious
Inorganic silts, micaceous or MH A-6 0.03 to 3 x 10-4
diatomaceous fine sandy or silty Semi-pervious to
soils, elastic silts pervious
Organic clays of high plasticity, CH A-8 3 x 10-3 to 3 x 10-6
fat clays Impervious
NOTE: Table adapted from Drainage Manual Volume 2, FDOT 1987.

Appendix J: Exfiltration Design J-1


January 2019

Drainage Design Guide


Appendix J: Exfiltration Design

Table J-2: Coefficient of Permeabilty (k) for SCS Hydrological Soils


HYDROLOGICAL SOIL CLASSIFICATION PERMEABILITY
TYPE CHARACTERISTICS (ft/day)
A Soils that have high infiltration rates even when 60
thoroughly wetted and a high rate of water
transmission
B Soils that have moderated infiltration rates when 48
thoroughly wetted and a moderated rate of water
transmission
C Soils that have slow infiltration rates when 24
thoroughly wetted and a slow rate of water
transmission
D Soils having very slow infiltration rates when 12
thoroughly wetted and a very slow rate of water
transmission
A/D Soils Type A under saturated natural conditions that 60
can be adequately drained, considering that
drainage is feasible and practical.
B/D Soils Type B under saturated natural conditions that 36
can be adequately drained, considering that
drainage is feasible and practical.
C/D Soils Type C under saturated natural conditions that 12
can be adequately drained, considering that
drainage is feasible and practical.
NOTE: Table adapted from Applicant’s Handbook: Regulation of Stormwater Management Systems.
SJRWMD, 2005

Appendix J: Exfiltration Design J-2


January 2019

Drainage Design Guide


Appendix K: Drainage Well Design – Reasonable Assurance Report

APPENDIX

K. DRAINAGE WELL DESIGN – REASONABLE ASSURANCE


REPORT

Appendix K: Drainage Well Design – Reasonable Assurance Report


January 2019

Drainage Design Guide


Appendix K: Drainage Well Design – Reasonable Assurance Report

K. DRAINAGE WELL DESIGN – REASONABLE ASSURANCE


REPORT

K.1 REASONABLE ASSURANCE REPORT

Submit a Reasonable Assurance Statement, in a report form, to the Department for


review and approval.

1) The report should provide a reasonable assurance that the buoyant stormwater
discharge into a Class G-III aquifer (total dissolved solids greater than 10,000 mg/L) via
the drainage well(s) has a minimum potential to:

i) Due to buoyancy, rise into a preferential pathway in a Class G-II aquifer


system [Underground Source of Drinking Water (USDW) total dissolved solids
less than 10,000 mg/L)]
ii) Impact any surface water bodies in the vicinity of the project via
groundwater discharge
iii) Cause mounding that may rise into USDW and/or manifest on a land
surface as a spring due to pressurization of the wells that are located nearby or
receive stormwater flow from a building roof at elevation + XXX feet.
The report should be signed and sealed by a Florida Licensed Professional Geologist/
Engineer with hydro-geological expertise to:

 Document the existence of sufficient confining strata above the base of the well
casing, to minimize the potential of injected buoyant stormwater to rise into a G-II
aquifer system.
 Show the injected stormwater has minimum potential to impact any surface water
bodies in the vicinity of the project that may be affected via groundwater
discharge.
 Show the pressurization of the well due to its location at elevation + XXX feet or
receiving stormwater flow from a building roof elevation + XXX feet will not cause
mounding that may rise into USDW and/or manifest as a spring on a land
surface.

2) The report may be submitted at this time or its submission may be deferred until the
stormwater drainage well(s) construction permit has been issued. However, it must be
submitted for Department review and approval prior to the Department’s authorization to
use the stormwater drainage well(s).

Appendix K: Drainage Well Design – Reasonable Assurance Report K-1


January 2019

Drainage Design Guide


Appendix K: Drainage Well Design – Reasonable Assurance Report

3) Indicate whether you wish to submit this report before or subsequent to Department
issuance of a stormwater drainage well construction permit. The Department strongly
advises the former option to most efficiently achieve the operational authorization of the
well(s).

4) Note that no fluid should be discharged into the stormwater drainage well(s) without
written authorization from the Department to use the well(s).

5) Issuance of a well construction permit does not obligate the Department to authorize
the use of well(s), unless the well(s), information required by the permit, and this report
qualifies for an authorization.

Appendix K: Drainage Well Design – Reasonable Assurance Report K-2


January 2019

Drainage Design Guide


Appendix L: Exfiltration Design - Sample Plans

APPENDIX

L. EXFILTRATION DESIGN - SAMPLE PLANS

Appendix L: Exfiltration Design - Sample Plans


January 2019

Drainage Design Guide


Appendix L: Exfiltration Design - Sample Plans

Appendix L: Exfiltration Design - Sample Plans L-1


January 2019

Drainage Design Guide


Appendix L: Exfiltration Design - Sample Plans

Appendix L: Exfiltration Design - Sample Plans L-2


January 2019

Drainage Design Guide


Appendix L: Exfiltration Design - Sample Plans

Appendix L: Exfiltration Design - Sample Plans L-3


January 2019

Drainage Design Guide


Appendix M: Pipe Corrosion Tables and Charts

APPENDIX

M. PIPE CORROSION TABLES AND CHARTS

Appendix M: Pipe Corrosion Tables and Charts


January 2019

Drainage Design Guide


Appendix M: Pipe Corrosion Tables and Charts

TABLE OF CONTENTS

Figure M-1: Estimated Service Life – 16 ga. Galvanized Steel Pipe ........................... M-1
Table M-1: Estimated Service Life – 16 ga. Galvanized Steel Pipe ............................ M-2
Figure M-2: Estimated Service Life – 16 ga. Aluminized Steel Pipe ............................ M-3
Table M-2: Estimated Service Life – 16 ga. Aluminized Steel Pipe ............................. M-4
Figure M-3: Estimated Service Life – 16 ga. Aluminum Pipe ...................................... M-5
Table M-3: Estimated Service Life – 16 ga. Aluminum Pipe........................................ M-6
Figure M-4: Estimated Service Life – 60” Dia. Reinforced Concrete Pipe, S = 1500... M-7
Table M-4: Estimated Service Life – 60" Dia. Reinforced Concrete Pipe, S = 1500 .... M-8

Appendix M: Pipe Corrosion Tables and Charts M-i


January 2019

Drainage Design Guide


Appendix M: Pipe Corrosion Tables and Charts

Figure M-1: Estimated Service Life – 16 ga. Galvanized Steel Pipe

Appendix M: Pipe Corrosion Tables and Charts M-1


January 2019

Drainage Design Guide


Appendix M: Pipe Corrosion Tables and Charts

Table M-1: Estimated Service Life – 16 ga. Galvanized Steel Pipe


Resistivity
pH 1000 1500 2000 3000 4000 5000 7000 10000 15000 20000 30000 40000 ≤50000
5.0 7 10 12 15 17 19 21 24 27 29 32 34 36
5.1 7 10 12 15 17 19 21 24 27 29 32 34 36
5.2 8 10 13 16 18 19 22 25 28 30 33 35 37
5.3 8 11 13 16 18 20 22 25 28 30 33 35 37
5.4 8 11 13 16 19 20 23 25 28 31 34 36 37
5.5 9 12 14 17 19 21 23 26 29 31 34 36 38
5.6 9 12 14 17 19 21 24 26 29 32 35 37 38
5.7 10 13 15 18 20 22 24 27 30 32 35 37 39
5.8 10 13 15 18 21 22 25 27 30 32 36 38 39
5.9 11 14 16 19 21 23 25 28 31 33 36 38 40
6.0 11 14 16 20 22 23 26 28 32 34 37 39 41
6.1 12 15 17 20 22 24 26 29 32 34 37 40 41
6.2 13 16 18 21 23 25 27 30 33 35 38 40 42
6.3 13 16 19 22 24 25 28 31 34 36 39 41 43
6.4 14 17 19 22 24 26 29 31 34 36 40 42 43
6.5 15 18 20 23 25 27 30 32 35 37 40 43 44
6.6 16 19 21 24 26 28 31 33 36 38 41 44 45
6.7 17 20 22 25 27 29 32 34 37 39 42 45 46
6.8 18 21 23 26 29 30 33 36 39 41 44 46 48
6.9 20 23 25 28 30 32 34 37 40 42 45 47 49
7.0 22 25 27 30 32 34 36 39 42 44 47 49 51
7.1 24 27 29 32 34 36 39 41 44 46 50 52 53
7.2 28 31 33 36 38 40 42 45 48 50 53 55 57
7.3 34 37 39 42 45 46 49 52 54 57 60 61 64
7.4 - 9.0 34 37 42 49 55 60 69 80 95 107 126 142 155
Estimated Service Life: (SL) = 17.24{Log10R - Log10[2160-2490(Log10pH)]} for 5<pH7.3
(SL) = 1.84 R0.41 for 7.3 <pH <9

Appendix M: Pipe Corrosion Tables and Charts M-2


January 2019

Drainage Design Guide


Appendix M: Pipe Corrosion Tables and Charts

Figure M-2: Estimated Service Life – 16 ga. Aluminized Steel Pipe

Appendix M: Pipe Corrosion Tables and Charts M-3


January 2019

Drainage Design Guide


Appendix M: Pipe Corrosion Tables and Charts

Table M-2: Estimated Service Life – 16 ga. Aluminized Steel Pipe


Resistivity
pH 1000 1500 2000 3000 4000 5000 7000 10000 15000 20000 30000 40000 ≤50000
5.0 19 28 34 43 49 54 61 69 78 84 93 99 104
5.1 20 29 35 44 50 55 62 70 79 85 94 100 105
5.2 21 30 36 45 51 56 63 71 80 86 95 101 106
5.3 22 31 37 46 52 57 65 72 81 87 96 102 107
5.4 24 32 39 48 54 59 66 74 82 89 98 104 109
5.5 25 34 40 49 55 60 67 75 84 90 99 105 110
5.6 26 35 41 50 56 61 69 76 85 91 100 106 111
5.7 28 37 43 52 58 63 70 78 87 93 102 108 113
5.8 29 38 44 53 59 64 72 79 88 94 103 109 114
5.9 31 40 46 55 61 66 73 81 90 96 105 111 116
6.0 33 41 48 56 63 68 75 83 91 98 106 113 118
6.1 34 43 50 58 65 69 77 84 93 100 108 115 119
6.2 36 45 51 60 67 71 79 86 95 101 110 116 121
6.3 38 47 54 62 69 73 81 88 97 104 112 119 123
6.4 41 50 56 65 71 76 83 91 100 106 115 121 126
6.5 43 52 58 67 73 78 86 93 102 108 117 123 128
6.6 46 55 61 70 76 81 88 96 105 111 120 126 131
6.7 49 58 64 73 79 84 92 99 108 114 123 129 134
6.8 53 62 68 77 83 88 95 103 112 118 127 133 138
6.9 57 66 72 81 87 92 100 107 116 122 131 137 142
7.0 to 8.5 63 72 78 87 93 98 105 113 122 128 137 143 148
8.6 46 55 61 70 76 81 88 96 105 111 120 126 131
8.7 36 45 51 60 67 71 79 86 95 101 110 116 121
8.8 29 38 44 53 59 64 72 79 88 94 103 109 114
8.9 24 32 39 48 54 59 66 74 82 89 98 104 109
9.0 19 28 34 43 49 54 61 69 78 84 93 99 104
Estimated Service Life (SL) = 50{Log10R - Log10[2160 - 2490(Log10pH)]} for 5.0 <pH<7.0
(SL) = 50(Log10R - 1.746) for 7.0 <pH <8.5
(SL) = 50{Log10R - Log10{2160 - 2490 Log10[7 - 4(pH - 8.5)]}} for 8.5<pH <9.0

Appendix M: Pipe Corrosion Tables and Charts M-4


January 2019

Drainage Design Guide


Appendix M: Pipe Corrosion Tables and Charts

Figure M-3: Estimated Service Life – 16 ga. Aluminum Pipe

Appendix M: Pipe Corrosion Tables and Charts M-5


January 2019

Drainage Design Guide


Appendix M: Pipe Corrosion Tables and Charts

Table M-3: Estimated Service Life – 16 ga. Aluminum Pipe


Resistivity
pH ≥200 400 600 800 1000 1200 1400 1600 1800 2000 2300 2700 3200 3800 4500 ≤5000
4.5 & 9.0 36 39 40 41 41 42 42 42 43 43 43 43 44 44 44 45
4.6 & 8.9 38 41 42 43 43 44 44 45 45 45 45 46 46 47 47 48
4.7 & 8.8 40 43 44 45 46 46 47 47 47 48 48 48 49 49 50 51
4.8 & 8.7 42 45 46 48 48 49 49 50 50 50 51 51 52 52 53 54
4.9 & 8.6 44 48 49 50 51 52 52 53 53 54 54 55 55 56 56 57
5.0 & 8.5 46 50 52 53 54 55 56 56 57 57 58 58 59 59 60 61
5.1 49 53 56 57 58 59 60 60 61 61 62 62 63 64 65 66
5.2 & 8.4 52 57 59 61 62 63 64 65 65 66 67 67 68 69 70 71
5.3 55 61 64 66 67 68 69 70 71 71 72 73 74 75 76 77
5.4 & 8.3 59 66 69 71 73 74 75 76 77 78 79 80 81 82 83 84
5.5 63 71 75 78 80 81 83 84 85 86 87 88 90 91 92 93
5.6 & 8.2 68 78 82 85 88 90 91 93 94 95 97 98 100 102 104 105
5.7 74 85 91 95 98 100 102 104 106 107 109 111 113 116 118 119
5.8 & 8.1 81 95 102 107 110 114 116 119 121 122 125 128 131 134 137 138
5.9 89 107 115 122 127 131 134 138 140 143 146 150 154 158 163 165
≥6.0 & ≤8.0 100 122 133 142 149 154 159 164 168 171 176 182 188 194 200 204
Where:
SL = Years to first perforation
Service Life (SL) = Tp / (RpH + Rr ) Tp = Thickness of pipe (inches)
RpH = Corrosion rate for pH (inches/year)
Rr = Corrosion rate for resistivity (inches/year)

Appendix M: Pipe Corrosion Tables and Charts M-6


January 2019

Drainage Design Guide


Appendix M: Pipe Corrosion Tables and Charts

Figure M-4: Estimated Service Life – 60” Dia. Reinforced Concrete Pipe, S = 1500

Appendix M: Pipe Corrosion Tables and Charts M-7


January 2019

Drainage Design Guide


Appendix M: Pipe Corrosion Tables and Charts

Table M-4: Estimated Service Life – 60" Dia. Reinforced Concrete Pipe, S = 1500
Chlorides
pH 15000 13000 11000 9000 7000 5000 3000 2000 1000 750 500 250
5.0 88 93 99 107 118 135 164 192 250 278 324 360
5.1 89 94 101 109 119 136 165 193 251 279 325 360
5.2 90 95 102 110 121 137 167 194 252 281 327 360
5.3 91 96 102 111 122 138 167 195 253 282 327 360
5.4 92 97 103 111 122 139 168 196 253 282 328 360
5.5 92 97 103 112 123 139 168 196 254 282 328 360
5.6 93 98 104 112 123 140 169 196 254 283 329 360
5.7 93 98 104 112 123 140 169 197 254 283 329 360
5.8 93 98 104 113 124 140 169 197 255 283 329 360
5.9 93 98 105 113 124 140 170 197 255 284 330 360
≥6.0 94 99 105 113 124 141 170 197 255 284 330 360

Conversion Factors for Different Size Culverts SL Reduction Factors for Sulfates
Pipe Dia. Mult. By Pipe Dia. Mult. By Sulfate Content Subtract from SL
12” 0.36 48” 0.76 1500 0
18” 0.36 60” 1.00 3200 5
24” 0.41 72” 1.25 4900 10
30” 0.48 84” 1.51 6600 15
36” 0.54 96” 1.77 8300 20
42” 0.65 108” 2.04 10000 25
Note: Sulfate derating not applicable
when Type V cement is used.

Service Life (SL) = 1000(1.107CC0.717D1.22K-0.37W-0.631) - 4.22x1010(pH-14.1) - 2.94x10-3(S) + 4.41


Where: C = Sacks of cement per cubic yard D = Steel depth in concrete K = Environmental chloride concentration in ppm
W = Total percentage of water in the mix S = Environmental sulfate content in ppm

Appendix M: Pipe Corrosion Tables and Charts M-8


January 2019

Drainage Design Guide


Appendix N: A Rationale for Stormwater Rule Standards

APPENDIX

N. A Rationale for Stormwater Rule Standards

Appendix N: A Rationale for Stormwater Rule Standards


January 2019

Drainage Design Guide


Appendix N: A Rationale for Stormwater Rule Standards

N. A RATIONAL FOR STORMWATER RULE STANDARDS


The following is an excerpt from a paper titled “The Evolution of Florida’s
Stormwater / Watershed Management Program” by Eric H. Livingston, FDEP.

The overriding standards of the Stormwater Rule are the state’s water quality standards
and appropriate regulations established in other FDEP rules. Therefore, an application
for a stormwater discharge permit must provide reasonable assurance that stormwater
discharges will not violate state water quality standards. Because of the potential
number of discharge facilities and the difficulties of determining the impact of any facility
on a waterbody or the latter’s assimilative capacity, the Department decided that the
Stormwater Rule should be based on design and performance standards.

The performance standards established a technology-based effluent limitation against


which an applicant can measure the proposed treatment system. Compliance with the
rule’s design criteria created a presumption that the desired performance standards
would be met which, in turn, provided a rebuttable presumption that water quality
standards would be met. If an applicant wanted to use Best Management Practices
(BMPs) other than those described in the rule, then a demonstration must be made that
the BMP provides treatment that achieves the desired pollutant removal performance
standard. The actual design and performance standards are based on a number of
factors which will subsequently be discussed.

1. Stormwater Management Goals - Stormwater management has multiple


objectives including water quality protection, flood protection (volume, peak
discharge rate), erosion and sediment control, water conservation and reuse,
aesthetics and recreation. The basic goal for new development is to assure that
the post-development peak discharge rate, volume, timing and pollutant load
does not exceed pre- development levels. However, BMPs are not 100%
effective in removing stormwater pollutants while site variations can also make
this goal unachievable at times. Therefore, for the purposes of stormwater
regulatory programs, the Department (water quality) and the state’s regional
Water Management Districts (flood control) have established performance
standards based on risk analysis and implementation feasibility.

2. Rainfall Characteristics - An analysis of long term rainfall records was undertaken


to determine statistical distribution of various rainfall characteristics such as
storm intensity and duration, precipitation volume, time between storms, etc. It
was found that nearly 90% of a year’s storm events occurring anywhere in
Florida produce a total of 1 inch of rainfall or less. Also, 75% of the total annual
volume of rain falls in storms of 1-inch or less. Finally, the average inter-event
time between storms is approximately 80 hours (5).

Appendix N: A Rationale for Stormwater Rule Standards N-1


January 2019

Drainage Design Guide


Appendix N: A Rationale for Stormwater Rule Standards

3. Runoff Pollutant Loads - The first flush of pollutants refers to the higher
concentrations of storm water pollutants that characteristically occur during the
early part of the storm with concentrations decaying as the runoff continues.
Concentration peaks and decay functions vary from site to site depending on
land use, the pollutants of interest, and the characteristics of the drainage basin.
Florida studies (6, 7) indicated that for a variety of land uses the first .5 inch of
runoff contained 80-95 percent of the total annual loading of most stormwater
pollutants. However, first flush effects generally diminish as the size of the
drainage basin increases and the percent impervious area decreases because of
the unequal distribution of rainfall over the watershed and the additive phasing of
inflows from numerous small drainages in the larger watershed. In fact, as the
drainage area increases in size above 100 ac the annual pollutant load carried in
the first flush drops below 80% because of the diminishing first flush effect.

4. BMP Efficiency and Cost Data - Numerous studies conducted in Florida during
the Section 208 program generated information about the pollutant removal
effectiveness of various BMPs and the costs of BMP construction and operation.
Analysis of this information revealed that the cost of treatment increased
exponentially after “secondary treatment” (removal of 80% of the annual load)
(8).

Selection of Minimum Treatment Levels - After review and analysis of the above
information, and after extensive public participation, the Department set a
stormwater treatment objective of removing at least 80% of the average annual
pollutant load for stormwater discharges to Class III (fishable/swimmable) waters.
A 95% removable level was set for storm water discharges to sensitive waters
such as potable supply waters (Class I), shellfish harvesting waters (Class II) and
Outstanding Florida Waters. The Department believed that these treatment levels
would protect beneficial users and thereby establish a relationship between the
rule’s BMP performance standards and water quality standards.

References:

5. Wanielista, M.P., et. al. Precipitation, Inter-event Dry Periods, and Reuse Design
Curves for Selected Area of Florida. Final report submitted to Florida
Department of Environmental Regulation, 1991.

6. Wanielista, M.P., et. al. Stormwater Management Practices Evaluations. Reports


submitted to East Central Florida Regional Planning Council, 1977.

7. Miller, R.A. Percentage Entrainment of Constituent Loads in Urban Runoff, South


Florida, USGS WRI Report 84-4329, 1985.

8. Wanielista, M.P., et. al. Stormwater Management Manual. Prepared for Florida
Department of Environmental Regulation, 1982.

Appendix N: A Rationale for Stormwater Rule Standards N-2

You might also like