Finite Element Analysis of Embankments On Soft Ground PDF
Finite Element Analysis of Embankments On Soft Ground PDF
by
Darien Russell
MSc, BSc
June, 1992
Summary
The objectives of the research were threefold. Firstly, to improve .the
numerical modelling capability for reinforced embankments constructed
over soft compressible soils containing vertical drains. Secondly, to
demonstrate the ability to model accurately such embankments. Finally, to
develop simplified procedures to be used in the design of embankments
over soft soils.
The modifications to the finite element program, CRISP, included the
incorporation of three additional elements: modelling the reinforcement, the
soil/reinforcement interface and the vertical drains. The facility to vary
permeability with stress level was also implemented. A technique for
modelling the consolidation of soil containing vertical drains in plane strain
finite element analyses was developed and validated.
The modified program was validated in three ways. Firstly, each element
was used to analyse simple problems so that the correct formulation was
ensured. Secondly, a series of analyses was carried out of problems for
which analytical solutions were available; these problems involved collapse
of undrained subsoils and consolidation around a single vertical drain.
Thirdly, an analysis of a case history of an embankment constructed over a
normally consolidated clay, improved with vertical drains, was performed.
Based on the results of the previous finite element analyses, and an
additional analysis of an idealized two-stage constructed embankment,
simple design procedures have been proposed. Firstly, a method for the
design of single stage embankments and, secondly, a method for the
calculation of subsoil strength increases in multi-stage construction, which
can be used in conjuction with limit equilibrium analyses.
It is concluded that the finite element method is a useful technique for the
analysis of reinforced embankments over soft soils containing vertical
drains.
II
Acknowledgement
The research presented in this Thesis was carried out under the
supervision of Dr. C.C. Hird and Dr. l.C. Pyrah, as part of a SERC funded
project.
The Author would very much like to thank his supervisors for their
invaluable help and encouragement with all aspects of the research project.
Their numerous comments during the preparation of this Thesis are
gratefully acknowledged.
Dr. R.A. Jewell was involved with the SERC contract and participated in
several discussions during the project, his comments were both timely and
useful.
The Author would like to express his gratitude to all members of the
geotechnics group at the University of Sheffield who always found time to
discuss the work. The staff at the University's Computer Services
department provided advice and facilities whenever necessary.
III
Table of Contents
1. Introduction
I
2
4
5
6
7
8
9
9
II
11
12
14
17
18
21
Iv
23
25
27
29
29
23
3.1. Crisp
29
30
31
3.2. Reinforcement Element
3.2.1. Constitutive Relationship
3.2.2. Reinforcement Element Stiffness Matrix
Formulation
3.2.3. Equivalent Nodal Forces
3.2.4. Analyses Using the Reinforcement Element
3.3. Interlace Element
3.3.1. Interface Element Stiffness Matrix Formulation
3.3.2. Transformation to Gobat Coordinate System
3.3.3. Interface Element Constitutive Relationship and
Modes of Behaviour
3.3.4. Equivalent Nodal Forces
3.3.5. Analyses Using the Interface Element
3.4. Drainage Element
3.4.1. Element Formulation
3.4.2. Transformation to Two-Dimensions
3.4.3. Incorporation of the Drainage Element in a
Two-Dimensional Mesh
3.4.4. Analyses Using the Drainage Element
3.5. Variation of Permeability with Stress Level
3.6. Summary
4.1. Introduction
4.1.1. Plasticity Theory
4.1.2. reinforced embankments
4.1.3. Application of Plasticity Theory to Reinforced
Embankment Problems
4.2. Finite Element Analysis of Plasticity Problems
4.2.1. Meshes for Equivalent Loading Problems
4.2.2. Material Properties
4.2.3. Results
4.2.4. Effect of the interface element
4.2.5. predicted displacements for a rough footing on a
uniform strength/limited depth subsoil
4.2.6. further ana'ysis using modified cam-clay
33
34
35
38
39
39
39
41
42
44
45
46
46
47
48
49
49
51
52
52
52
54
55
58
58
60
61
64
69
73
4.3. Embankments with Constant Side Slope
4.3.1. Finite Element Analyses of Constant Side Slope
Loading
4.3.2. Results
4.3.3. A comparison of limit equilibrium and finite
element analysis for embankments with constant
side slope
4.4. Summary
vi
77
80
81
83
86
90
90
93
104
107
114
117
120
121
121
122
122
122
124
126
126
126
129
129
129
131
135
138
138
139
140
145
147
7.1. Introduction
147
148
149
154
172
172
179
183
184
189
7.5. Summary
198
200
200
8.2. Conclusions
201
201
202
203
204
205
205
206
206
VII
207
218
224
226
References
VIII
List of Symbols
List of Symbols
a
a0
b
b*
B
B
B
Cohesion intercept
C,
CI'
cc
ch
Coefficient of consolidation
CII
Compression index
D
e
e0
E
E
F0
FS
G
Factor of safety
Elastic shear modulus
Initial shear modulus
Hd
Embankment height
Embankment design height
Hydraulic gradient
k,1
Permeability
Horizontal permeability
x
List of Symbols
Interface element normal stiffness
Ic
lc
Vertical permeability
Drain permeability
k0 Initial permeability
'e
Link matrix
LL
Liquid limit
RIr
Nr
P1
Plasticity index
'i
List of Symbols
r
rJr
Time
Transformation matrix
Pore pressure
depth
Overall average degree of consolidation
U
V
Specific volume
Volume
wr
wn
we
xl
z
a
13
o
L.
List of Symbols
Strain
qlp'
.11
K
(?.-ic)I?
Poissons ratio
Stress
a1,
Normal stress
Vertical stress
Superscripts
Effective stress
*
T
Virtual displacement
Transformed matrix
XIII
List of Symbols
Subscripts
ax
Axisymmetric conditions
p1
xiv
List of Figures
List of Figures
Chapter 3- The Finite Element Program CRISP and Modifications
3.1
3.4 Interface element shear stress sign convention and stress regimes.
3.5 Drainage element.
Chapter 4- Collapse of Undrained Subsoils
4.1
4.4
xv
List of Figures
4.11 Load displacement curves for load controlled analyses: (a) subsoil of
uniform strength/limited depth; (b) subsoil with strength increasing
linearly with depth.
4.12 Surface settlement profiles for load controlled analyses: (a) subsoil
of uniform strength/limited depth ; (b) subsoil with strength increasing
linearly with depth.
4.13 Vertical stress distribution at failure predicted using three different
finite element meshes compared with plasticity theory.
4.14 Plasticity analysis of a rough rigid footing on a subsoil of uniform
strength/limited depth: (a) slip line mesh; (b) hodograph.
4.15 Comparison of the lateral displacements predicted by finite element
and plasticity analyses: (a) 2D; (b) 40; (c) 6D; (d) 80 from toe.
4.16 Stress and strength profiles used for the modified Cam-clay analysis.
4.17 Load settlement curves for a displacement controlled analysis using
modified Cam-clay to model a subsoil with strength increasing
linearly with depth.
4.18 Comparison of finite element and plasticity predicted vertical stress
distributions at failure beneath a rough footing on a subsoil with
strength increasing linearly with depth modelled using modified
Cam-clay.
4.19 Error in the undrained shear strength profile used for the modified
Cam-clay analysis (after Potts and Ganendra, 1991).
4.20 Proposed practical embankment shape to approximate ideal loading
predicted by plasticity theory (after Jewell, 1988) and a proposed
simplified constant side slope design profile.
4.21 Typical load displacement curves for a constant side slope analysis.
Curves shown for a uniform strength/limited depth analysis with
xID=5.
4.22 Comparison of the finite element predicted collapse load of a
constant side slope embankment with plasticity theory for a subsoil
with uniform strength/limited depth.
4.23 Comparison of the finite element predicted collapse load of a
constant side slope embankment with plasticity theory for a subsoil
with strength increasing linearly with depth.
4.24 Displacement vectors at failure for embankments with constant side
slopes on a subsoil of uniform strength/limited depth: (a) x/D=1; (b)
XdID3; (c) Xd/D5.
xvi
List of FIGures
4.27 Maximum shear strain contours at failure for a constant side slope
embankment on a subsoil with strength increasing linearly with depth
(pxIs4.8).
Chapter 5- Consolidation Around a Single Vertical Drain
5.1
Unit cells and direction of flow: (a) axisym metric conditions; (b) plane
strain conditions.
5.2
5.3 Finite element mesh used for unit cell analysis comparison with
Hansbo (1981).
5.4 Comparison of finite element and analytical results for consolidation
of a unit cell: (a) L=0.0; (b) L=0.5; (c) L=3.0: (d) L5.0.
5.5 Finite element mesh used for unit cell analyses: comparison with
Jamiolkowski et al (1983).
5.6
5.7
xvii
List of Figures
5.13 Errors in geometry matched plane strain analyses: (a) based on
settlements; (b) based on pore pressure.
5.14 Development of surface settlement profile in geometry matched
plane strain analysis: (a) without smear or well resistance; (b)
without smear but with well resistance.
5.15 Comparison of pore pressure distribution from axisymmetric and
geometry matched plane strain analyses without smear or well
resistance (degree of consolidation=65%).
5.16 Excess pore pressure distribution after the first consolidation
increment from the axisymmetric finite element analysis without
smear or well resistance.
5.17 Comparative results for axisymmetric and matched plane strain
analyses with vertical flow (k.,=k,j, degree of consolidation based on
excess pore pressure: (a) without smear or well resistance; (b)
without smear but with well resistance.
5.18 Errors in geometry matched plane strain analyses for unit cells with
vertical permeability (k,=k,), based on excess pore pressure.
Chapter 6 - Case History: Porto Tolle
6.1
List of Figures
6.12 Mesh used for plane strain analysis.
6.13 Surface settlement at centreline.
6.14 Maximum horizontal displacement.
6.15 Lateral movement profiles at the inclinometer position: (a) at the end
of the construction stage; (b) at the end of the consolidation stage.
6.16 Excess pore pressure on centreline of the embankment (full plane
strain): (a) 19.7m below ground level; (b) 12.6m below ground level.
Chapter 7- Multi-Stage Embankment Construction
7.1 Idealized two-stage embankment geometry.
7.2
List of Figures
7.19 Undrained shear strength anisotropy of normally consolidated clays.
7.20 Estimation of vertical stress increase in a soil due to embankment
loading: (a) one-dimentional; (b) Gray (1936).
7.21 Predicted strength increase near centreline.
7.22 Predicted absolute strength near centreline.
7.23 Contours at the end of the consolidation stage predicted by the finite
element analysis: (a) strength increase; (b) absolute strength.
7.24 Contours of strength increase at the end of the consolidation stage:
(a) predicted by Method G; (b) predicted by Method H.
7.25 Contours of absolute strength at the end of the consolidation stage:
(a) predicted by Method G; (b) predicted by Method H.
7.26 Contours of strength difference at the end of the consolidation stage:
(a) between Method G and finite element; (b) between Method H
and finite element
7.27 Porto Tolle predicted: (a) degree of consolidation; (b) total vertical
stress.
7.28 Porto Tolle predicted: (a) undrained shear strength increase; (b)
absolute undrained shear strength..
7.29 Contours of strength increase at the end of the consolidation stage
at Porto Tolle predicted by: (a) the finite element analysis; (b) the
simplified method.
7.30 Contours of the strength difference predicted by finite element
analysis and the simplified method at the end of the consolidation
stage at Porto Tolle.
Appendix A - Undrained Shear Strength of Modified Cam-clay
Al
A2
A3
xx
List of Figures
Appendix C - Development of Pore Pressures Due to Ramp Loading
Cl Ramp loading scheme.
xxi
List of Tables
List of Tables
Chapter 6 - Case History: Porto Tolle
6.1 Summary of Porto Tolle soft clay material properties.
6.2 Summary of parameters used to model Porto Tolle clay.
Chapter 7- Multi-Stage Embankment Construction
7.1 Summary of assumptions in the simple methods used to predict
strength increases.
xxi
Introduction
1. Introduction
1.1. Embankments Over Soft Soils
Recent social and economic development around the world has brought
about an increase in the construction of embankments used in highway and
railway systems, flood and irrigation projects and harbour and airport
installations. This coupled with increasing urbanisation has required the
geotechnical engineer to design and construct embankments over
increasingly weak and compressible soils.
Often the soil is sufficiently weak that the embankment cannot be
constructed in a single lift, in the area available; in such situations it is
necessary to improve the soil on which the embankment is to be placed.
This improvement can take several forms depending on the soil conditions
encountered.
Where settlement is the major consideration the rate at which settlement
occurs can be increased with the installation of vertical drains (Barron,
1948; Hansbo, 1981; Holtz et al, 1987; Holtz et al, 1991) and the
post-construction settlements therefore reduced. Preloading the
compressible soils with a surcharge is a technique used with both
embankments and settlement sensitive structures (Johnson, 1970a) and is
often used in conjunction with vertical drains (Johnson, 197Db).
Other techniques may have to be considered to ensure stability of the
embankment during construction. Stage construction can be used to
construct embankments with relatively steep side slopes (Jardine and
Hight, 1987; Ladd, 1991; Leroueil et al, 1991). This method relies on the
increase in the undrained shear strength of the subsoil during consolidation
and is therefore most beneficial when used with vertical drains. An
Introduction
increasingly popular technique is the use of tensile reinforcing material,
such as geogrids, geotextiles or steel reinforcement, placed at the base of
the embankment fill (Bonaparte and Christopher, 1987). The reinforcement
provides lateral restraint and effectively increases the bearing capacity of
the subsoil.
Other less common expedients include the use of light-weight fills, drainage
trenches, stone columns, lime columns, piles, replacement of the subsoil,
electra-osmosis, electro-injection, dynamic compaction or a combination of
the above (brief descriptions and further references for these techniques
can be obtained from: Pilot et al, 1987; Delmas et at, 1987, Lerouiel et al,
1991).
In this research three of the more popular embankment construction
techniques have been examined: 1) reinforcement at the base of the fill
material, 2) vertical drains in the subsoil and 3) multi-stage construction with
vertical drains.
Introduction
Introduction
(1991) examined three case histories and demonstrated that the factor of
safety calculated using an effective stress analysis was approximately twice
as high as that calculated using an undrained strength analysis.
In this Thesis, where appropriate, the philosophy of an undrained strength
analysis has been used for multi-stage embankment construction (Chapter
7). Undrained analysis based on the initial undrained shear strength has
been used as a conservative method of analysis for quickly constructed
single stage em bankments (Chapter 4).
Introduction
Limit equilibrium analysis has been extended to the analysis of
embankments with reinforcement at the base of the fill (Fowler, 1982;
Jewell, 1982). The reinforcement provides an additional resisting force, the
magnitude of which is dependent on both the fill material and the surface
undrained shear strength of the subsoil.
Plasticity Theory
Often an accurate estimation of the collapse load of a geotechnical structure
can be made using plasticity theory. Jewell (1988) has suggested that the
plasticity solutions for two idealized soil profiles (Mandel and Saleon, 1972;
Davis and Booker, 1973) have direct relevance to the stability analysis of
embankments constructed on soft soils. These plasticity solutions predict
the collapse loads for rigid strip footings on idealized subsoils. The solutions
can also model the shear stress distribution on the underside of the footing.
This shear stress can be interpreted as the action of an infinitely stiff
reinforcement; the varying roughness factors define the bond between the
soil and the reinforcement (Houlsby and Jewell, 1988).
The plasticity solutions have been shown to match well limit equilibrium
analysis for the idealized conditions (Jewell, 1988) but further research is
required before the method can be directly applied to embankment design.
Introduction
Introduction
Introduction
solutions to the same problem if incorrect discretization or solution
techniques are used.
Introduction
In drained conditions the Mohr-Colomb yield criterion is often used and for
mathematical convenience the flow rule is usually associated (Smith, 1982).
Such models produce excessive dilatancy and may predict an incorrect
response.
Critical State Models
Critical state theory (Schofield and Wroth, 1968; Roscoe and Burland, 1968;
Atkinson and Bransby, 1978) provides a realistic framework for the
prediction of the behaviour of normally and lightly over-consolidated clays.
These models can predict strain hardening for such soils and have been
used widely in numerical analyses to provide accurate predictions of
displacements and stresses. The increased complexity of these models
produces increasing difficulty in the analysis in which the equilibrium
condition is more difficult to enforce. This requires a larger number of
loading iterations or a more complex solution procedure.
Introduction
finite element method is therefore an excellent tool with which to study the
behaviour of embankments constructed over soft ground.
However, accurate numerical predictions require careful analysis and it is
often necessary to conduct a study in which material parameters, boundary
conditions and solution techniques are varied. Therefore, It may be
unrealistic to perform a finite element analysis at an initial stage of the
design process, instead simple design procedures may be used for
preliminary analysis.
The aims of the present research were:
1. To modify an existing finite element program, CRISP (Gunn and
Britto, 1984), to better model reinforced embankments over soft soils
containing vertical drains (Chapter 3).
2. To benchmark the modified program against analytical solutions for
both the undrained collapse of idealized soils and the consolidation
behaviour of soil surrounding a single vertical drain (Chapters 4 and
5).
3. To develop procedures so that soil systems containing vertical drains
can be analysed accurately using the finite element method (Chapter
5).
4. To compare finite element predictions with observed behaviour for an
embankment constructed over soft soil incorporating vertical drains
(Chapter 6).
5. To develop simple procedures which aid routine design of
embankments on soft soils containing vertical drains, reinforcement
and stage constructed embankments (Chapters 4 and 7).
10
Literature Review
11
Literature Review
Literature Review
in good agreement with the observed values. Embankment stresses did not
compare favourably with those measured but the authors expressed
concern over the measured values. The finite element predicted tension
zones agreed closely with visible tension cracks.
Smith and Hobbs
Smith and Hobbs (1974) compared finite element analyses with centrifuge
model tests. The subsoil was represented with an undrained
elastic-perfectly plastic soil model. The effect of the elastic parameters and
the distance of the mesh boundaries were studied. The authors showed
that, for their analyses, the finite element method predicted collapse loads
agreed well with limit equilibrium calculations.
Wroth and Simpson
Wroth and Simpson (1972) analysed a test embankment constructed at
King's Lynn, UK. The subsoil was modelled using Cam-clay (Schofield and
Wroth, 1968) and the embankment loading by equivalent vertical loads
applied over a number of increments. Undrained analyses were carried out
representing the short term conditions. The predicted vertical and horizontal
displacements were in good agreement with those observed. However, the
predicted pore pressures were not in such good agreement. The authors
suggested that the disagreement may have been caused by consolidation
near the drained boundaries during construction.
The authors also carried out drained analyses to simulate long term
conditions. The predicted displacements were again in good agreement with
those observed.
13
Literature Review
14
Literature Review
Literature Review
Literature Review
Hird and Kwok (1990a) presented results from two embankment analyses
using the modified version of CRISP. Interface elements were used to
model the soil/reinforcement interfaces and it was shown that useful
information regarding the transmission of shear stresses from the soil to the
reinforcement could be extracted.
Hird and Kwok (1990b) carried out a parametric study of an idealized
embankment considering the effect of the reinforcement stiffness and the
subsoil depth. The authors concluded that provided the reinforcement was
sufficiently stiff and strong: 1) subsoil deformations would be significantly
reduced, 2) the development of tension in the reinforcement would be in
accordance with the principles suggested by Jewell (1988) and 3) very stiff
reinforcement may cause arching in the embankment fill. Study of the effect
of subsoil depth showed that for a subsoil of constant strength the effect of
reinforcement reduced with increasing depth.
Hird and Pyrah (1990) performed a parametric study of a second trial
embankment constructed at Stanstead Abbotts, UK The embankment was
modelled with both elements and equivalent vertical loads. The finite
element predicted vertical displacements compared well with those
observed. The horizontal displacements agreed less well and according to
the authors may have been attributed to the relatively coarse mesh. The
analyses modelling the embankment using equivalent vertical loads
produced unrealistic reinforcement stress and strain profiles, this may have
been due to the large deformations which occurred in a highly compressible
peat layer.
Literature Review
Literature Review
before the final loading stage which was then modelled as completely
undrained. The finite element analysis could not correctly predict failure as
the clay subsoil exhibited strain softening behaviour. However, the predicted
displacements and excess pore pressures due to the additional fill agreed
favourably with the observed values.
Imperial College
Several finite element studies have been carried out at Imperial College.
Bond (1984) compared different drainage conditions and illustrated the
strengthening effects of a surface crust for a 'typical' soft clay using an
extended modified Cam-clay constitutive model to represent the subsoil.
The same set of soil parameters were use by Smith (1984) to investigate
some effects of stage constructed embankments. The embankment was
constructed in Im lifts with partial dissipation of excess pore pressure
specified after each increment. Smith observed large principal stress
rotations during the first undrained loading stage and that these rotations
were greatest below the toe of the embankment. Interestingly no rotations
greater that 5 were observed during subsequent loading or consolidation
stages. Smith also plotted the variation of the undrained shear strength ratio
with the maximum principal effective stress and the vertical effective stress
(s)a 1' and sJa ') against embankment height and observed that neither of
these ratios fell below the normally consolidated value.
Further details of these analyses can be found in Jardine and Hight (1987).
Schafer
Schafer (1987) carried out fully coupled consolidation analyses of reinforced
embankments using a finite element program which included an extended
version of modified Cam-clay, used to model the soil, and bar elements,
used to model the reinforcement.
19
Literature Review
The program was used to analyse the Mohicanville Dike (Schafer, 1987;
Duncan et al, 1988). The subsoil consisted of peat and soft clay on which a
previous embankment had failed. Steel reinforcement was used to allow the
construction of the dike to the design height. The finite element analysis of
these highly complex subsoil conditions showed good agreement for
reinforcement stresses and subsoil displacements. However, the predicted
excess pore pressures were less good. The discrepancies were explained
as being due to creep occurring in the subsoil and ageing of the fill
material.
Schafer (1987) also analysed the test embankments at St Alban, Canada,
constructed on Champlain Clay. The author found that the initial results
were not in good agreement with the observed behaviour. A reinterpretation
of preconsolidation pressures improved the quality of the analysis with
respect to predicted failure, displacements and excess pore pressures. The
author suggested that the lack of agreement, when using the measured soil
properties, occurred because of the inability of the constitutive model to
reproduce the brittle behaviour of Champlain Clay.
Almeida
Almeida (1984) used a version of CRISP to compare finite element
analyses with centrifuge tests (Almeida et al 1985; Almeida et al, 1986). The
finite element analyses did not accurately predict failure, but this may have
been due to the embankment being modelled as elastic. Subsoil stress
paths predicted by Almeida were similar to those predicted by Smith (1984).
Almeida (1984) implemented a variation of permeability with stress level
relationship and presented results indicating improved pore pressure
prediction when using this algorithm. The predictions of excess pore
pressure using fully coupled consolidation were, generally, encouraging.
20
Literature Review
Literature Review
22
Literature Review
method the derived plane strain vertical drain spacing is only applicable for
one degree of consolidation.
A consolidation analysis of the embankment was compared with a partially
drained analysis in which the bulk compressibility of the pore water was
manipulated. The agreement of the finite element analyses with the
observed behaviour was not good but the consolidation and partially
drained analyses produced similar results.
2.3. Discussion
2.3.1. Summary of Embankment Finite Element Analyses
The finite element analyses reviewed above have been divided into two
categories: firstly, undrained analyses (which have been used to assess the
short term behaviour of a single stage constructed embankment) and,
secondly, analyses which have considered the dissipation of the excess
pore pressures due to the consolidation of the subsoil. Particular attention
has been placed on analyses which have modelled reinforcement andlor
vertical drains.
The development of the constitutive models used to represent the subsoil
has been an important aspect of the analysis of embankments constructed
over soft soils. The earliest analyses used an elastic subsoil model (Brown
and King, 1966) but quickly developed with a non-linear elastic model
(dough and Woodward, 1967), an elastic-perfectly plastic model (Smith
and Hobbs, 1974) and a critical state model (Wroth and Simpson, 1972).
Improvements of the constitutive modelling for the embankment material
have also been made and several authors have indicated that accurate
modelling of the embankment is essential to ensure realistic predictions.
Kwok (1987) performed a parametric study of embankment models. Hird
and Pyrah (1990) compared the use of finite elements, using an
23
Literature Review
Literature Review
25
Literature Review
26
Literature Review
Literature Review
28
3.1. Crisp
3.1.1. The History of Crisp
The computer programs known as CRISP were developed at Cambridge
University. Work started in 1975 when the program was originally called
MZOL but in 1976, after additional work, it was renamed CRISTINA. The
program was developed further and in 1982 given the name CRISP
(itical state Erogram). The 1982 version of the program (Gunn and
Britto, 1982), referred to as CRISP82, has been used in previous research
of reinforced embankments at the University of Sheffield (Kwok, 1987).
In 1984 a new version of the program was produced (CRISP84). The
double precision version of the 1984 program has been modified, by the
Author, and used in this research. In the Thesis this version of the program
will be referred to as CRISP.
Other versions of the program are available. In 1987 a reduced program
was made available to coincide with the publication of Britto and Gunn
29
CnSD
and Modifications
(1987). In 1990 the program was extensively rewritten for use on personal
computers (CRISP9O).
31
(Cook, 1981). If excessively large strains occur the solution may no longer
be reliable.
Embankment Construction
CRISP allows for the simulation of the construction of an embankment by
specifying an initial mesh containing elements which represent the fill
material. CRISP then allows these elements to be added as the analysis
proceeds. To avoid using a large number of thin elements the self weight of
the added elements is applied over a number of increments. Excavation can
be simulated by removing elements.
Restarted Analyses
The original CRISP program outputted information at the end of each
increment so that an analysis could be restarted from some earlier point.
The file produced became extremely large for any reasonable number of
increments. In order to reduce its size the program has been modified so
that restart information is only saved for specified increments. The analysis
can then be restarted from any one of these increments.
33
1
y
I=-i
S node
(ci)
strain
(b)
(ar)
and the
34
Crisp and Modifications
For the element shown in Figure 3.la the shape functions, N, in the local
coordinate system, (), are
N=[ O.5(_1) o.5(^1)
(1_2) ]
........................(3.2)
.................................................................(3.3)
where
ON
(34
In which x' is the coordinate system defining the element length, i.e. x'=O at
node I and x'=L at node 2, where L is the length of the element.
However, the shape functions are in terms of the element local coordinate
system, (p,). Using the chain rule for differentiation it can be shown that
35
(35)
N_ Nd,
ox'
, dx'
'36
---
dx' - L
Differentiating the shape functions, Equation 3.2 and using the chain rule
B
=[
(0.5 - ) (0.5
-F
2 ]
.................................(
3.7)
....
...............................................(3.8)
To transform Equation 3.9 into the x' coordinate system it must be multiplied
by , which from Equation 3.6 is L12. Thus
(Ke)xirrJ'l B TDBIt
. .............................................( 3.10)
36
matrix, (Kj 1,, to the global element stiffness matrix K in global coordinate
system (x,y).
ICT T(K)_r
................ (3.1 1)
where
000
0 0
0 0 ............................................ (3.12)
0000 Cs
cs0
r =
(3.13)
...................................................... (3.14)
37
cx
Q.5(a1+c)
LL1+L2
modified cj eo me try
deformed geometry
L1
.................................................... (3.15)
tJ'1 B Tad()
................................ ................(3.1 6)
39
CIISD and Modifications
U)
U)
a)
I-
In
T=6+qtan o
.1-I
U)
L
a
a)
-c
14
U)
nodal point
relative displacement
(b)
(a)
top
bottom
8q a
The shape functions for the interface element are defined in Equation 3.2
and can be rewritten
N=[
Ni (112 tS3 ]
We = Nrae
where
-N, 0 -N2 0 N2 0 N1 0 -N3 0
0 -N1 0 -N2 0 N2 0 N., 0 -N3
L
.. ( 3.21)
Nr-iF
N ]
ci=Dw 9 ..............................................................(3.22)
The vector of stresses, a, has components of normal stress,
a,,
and shear
stress, t. The constitutive matrix has components of shear stiffness, k5, and
normal stiffness, k,.
40
D-1 k
.............................. (3.23)
0 k
Equation 3.23 implies that there is no coupling of the shear and normal
stiffnessses of the interface.
The principal of virtual work can now be used to define a set of nodal forces
which are in equilibrium with the state of internal stress. It is assumed that a
set of virtual nodal displacements, a0*, cause a set of virtual nodal relative
displacements,
where
w=Nra............................................................. (3.211.)
and from the principal of virtual work
,jlJ*Tad()
aTFe .._. J1
(3.25)
Substituting for a and w, in Equation 3.25 and dividing both sides by a.*T
Fe= 1
N r3N rd(
) a e
.............................................. (3.26)
............................................. (3.27)
(3.28)
41
t
0
0
0
0
0
T=
0
t
0
0
0
0
0
0
t
0
0
0
0
0
0
t
0
0
0
0
0
0
t
0
0
0
0
0
0
t
...... . (3.29)
in which
coscc sina
-Sifla
................. (3.30)
43
POSITION a
UPPER___________
VIIRFAcE\+
REIN
FOREI(NT
LOVER
IUERFAcE
,0
POSITION b
E7_
/0
ck ci
\+\
Figure 3.4 - Interface element shear stress sign convention and shear
stress regimes.
shear stress regimes which may develop, by differential movement of the fill
and subsoil, in a reinforced embankment.
3.3.4. Equivalent Nodal Forces
The equivalent nodal forces, which are in equilibrium with the internal
stresses, are calculated for the interface element at the end of each
increment. Stresses (c, an), which act on a plane through the centre of the
interface element, are used to calculate the nodal forces F.
Consider an infinitesimally small section of the element. The forces which
are in equilibrium with the stresses are
dFx = (tcoscx-.. sin a)
cm
ciiy=(tsina-i-cmncosa)c
(3.31a)
(3.31b)
44
......................................................... (3.32)
The work done by the nodal forces due to the virtual displacements is
equivalent to the work done by the internal stresses
(dFe)=f1 N(rcosaansina)d,
(dFe) = J 11 N(tsina+ancosa)d
(3. 33a)
(3.33b)
45
................... (3.31)
where
KS=J' BTDBd(vol)
stiffness matrix
L=f" BmNPd(vot)
link matrix
,tf'
permeability matrix
ETkEd(VOt)
Crlso and Modifications
In
displacement node
+ pore pressure node
pore pressure shape functions, E are the first derivative of the pore
pressure shape functions, k is the permeability matrix, D is the constitutive
matrix, m=(
...................... (3.35)
where
cosa sina
1= 0
o
o
o
0
0
o
o 00
0
0 00
cosa sincx
0 cosa Sina 0 0
o
0
o -0 1 0
o
0
o
a 01
o
(3. 36)
The element length, L, and angle of orientation, a, are defined as for the
reinforcement element, Figure 3.2.
47
KSAd+LAu=AF1
continuity
LTAd-AtcDAu=AF2
where Ad, Au, AF1 and AF2 are increments of nodal displacement, pore
pressure, nodal force and nodal flow respectively.
Considering each term separately:
1 KSAd - is the change of nodal displacement, Ad, caused by an
increment of nodal force AF1.
2 LAu - is the change of nodal pore pressure, Au, caused by an
increment of nodal force AF1.
3 LTAd - is the change of nodal displacement, Ad, caused by an
increment of flow, AF2.
4 AtcDAu - is the change of nodal pore pressure, Au, caused by.an
ncrement of flow, AF2.
Terms 1 and 4 are relatively easy to understand, in that they have the usual
finite element meaning. KS is the force stiffness ascribed to the element
multiplied by the defined area; similarly, Q can be considered as the flow
48
LT
would be non zero) but a bar element with force and flow stiffness but no
real thickness. The element stiffness matrix, Equation 3.34, is reduced to
IKS 0
Ke L 0 Atb
...................(3.37')
49
Where k0 and e0 are the initial values of the permeability and void ratio
respectively and Ck is the permeability change index. Tavenas et al (1983)
showed that the permeability change index can be reasonably
approximated as Ck=O.5e0.
The permeability stress level relationship, Equation 3.38 (with Ck=O.5e0),
has been implemented in CRISP. The option is available when using
modified Cam-clay and if the option is selected the program calculates a
new horizontal and vertical permeability at the end of each increment.
Analyses, of an idealized triaxial test, were carried out in which the
permeability and void ratio were monitored and found to agree with
calculated values using Taylor's relationship. The variation of permeability
with stress level was used in an analysis of the Porto Tolle case history
(Chapter 6). A comparison of analyses of this case history showed
negligible difference with and without varying the permeability with stress
level.
Two aspects of the selection of permeability parameters which have been
found to be more significant are the initial values of both vertical and
horizontal permeabilities and ensuring that the coefficient of consolidation
throughout the clay depth has been modelled correctly (Section 7.3.2).
50
3.6. Summary
The finite element program CRISP is a versatile program which can be used
to model geotechnical problems. However, in order to model accurately
reinforced embankments over soft soils incorporating vertical drains several
modifications were necessary:
I The implementation of a bi-linear elastic one-dimensional element to
model the reinforcing material in plane strain analyses.
2 The implementation of an elastic-perfectly plastic relative
displacement interface element to model the soillreinforcement
interface in plane strain analyses.
3 The implementation of a one-dimensional bar element which can be
used to model vertical drains in axisymmetric or plane strain
analyses.
4 The implementation of an algorithm which allows the permeability to
be varied with stress level when using the modified Cam-clay
constitutive model.
The modified version of the program has been used in all analyses
presented in the remaining Chapters.
51
52
fI
cr1'
if3
a
x
Coincident upper and lower bound solutions can be found using methods
which depend on the solution of hyperbolic partial differential equations.
One such solution technique is the method of characteristics (Eg. Atkinson,
1981). This method involves the extension of two sets of slip lines from
known boundaries, the directions of the slip lines being defined by the
boundary conditions. At the intersection of two slip lines the state of stress
can be calculated. The slip lines are referred to as a and 13, or stress,
characteristics and are associated with positive and negative shear stresses
respectively, Figure 4.1. For undrained soil the a and
13
characteristics
intersect at 900.
From the slip line field it is possible to consti uct a vector displacement
diagram which defines the relative displacement of areas bounded by slip
lines. The displacement diagram is referred to as a hodograph.
53
Reinforcement
Subsoil
Plasticity theory assumes that the soil is perfectly plastic satisfying the
normality condition and that the flow rule is associated. It is also assumed
that the coaxiality condition, which states that the directions of principal
stress and principal strain increment coincide, is valid.
A set of strain increment characteristics, also referred to as velocity
characteristics or zero extension lines, can be defined in a similar way to the
stress characteristics shown in Figure 4.1. For an undrained soil which is
undergoing purely plastic deformations and has an associated flow rule,
characteristics of stress and strain increment coincide.
54
has been termed P and the force in the reinforcement due to the spreading
of the subsoil as ftidn (Jewell, 1988).
The stability of an embankment reinforced at its base has been discussed
by Hird and Jewell (1990). Three classes of failure can be identified: internal
stability of the embankment, instability of the subsoil and overall instability
involving the embankment and the subsoil. The second of these has been
selected for investigation; comparisons of finite element analyses and
plasticity theory are made for relevant subsoil conditions.
1972)
and secondly, a
55
pX
su
subsoil with strength increasing
linearly and indefinitely with depth
p
z
(b)
Figure 4.3 - Idealized soil profiles for plasticity analyses: (a) uniform
strength/limited depth; (b) strength increasing linearly with depth.
subsoil with strength increasing linearly and indefinitely with depth, Figure
4.3b (Davis and Booker, 1973).
Plasticity theory gives the distribution of vertical stress on the underside of
the rigid footing, shown in a non-dimensional form in Figure 4.4. These are
ideal distributions which make the best possible use of the available subsoil
strength and provide a meaningful comparison for finite element analyses.
56
CollaDse of Undrained Subsoils
25
20
15
Jo
b
10
5
x/D
10
5
01
0
5
1ox/Su.
10
D=4.Om
39 interface elements
1488 degrees of freedom
3D
1OD
DI
restrained
horizontally
restrained
horizontally
restrained vertically
and horizontally
Figure 4.5 - Finite element mesh used for uniform strength/limited depth
analyses (mesh shown for smooth case).
B=1 2.Om
1.5B
S
C
S.
C
S
I)
BI
horizontally and
vertically restrained
Figure 4.6 - Finite element mesh used for strength increasing linearly
with depth analyses (mesh shown for smooth case).
59
For the rough loading analysis a row of interface elements was included
along the surface and no horizontal movement was allowed.
60
4
E
1:
-10
-12
10
15
20
Shear strength S (lcPa)
25
30
4.2.3. Results
Displacement Controlled Analyses
In order to model the rigid strip footing assumed by the plasticity solutions,
plane strain finite element analyses were carried out by applying increments
of uniform vertical displacement along the loaded surface. Failure was
deemed to have occurred when a further increase in vertical displacement
caused no significant increase in vertical stress at any point on the
displaced boundary. Failure was reached in approximately 100 increments.
The normal stresses in the upper interface elements (equivalent to the
vertical stress distributions at the subsoil surface) were compared with the
vertical stress distribution predicted by plasticity theory. The load-settlement
curves for three points on the subsoil surface beneath the footing, for each
of the four cases, are shown in Figure 4.8.
61
700
x0
Ax=5D
600
500
xlOD
400
I1-
300
U
1-
200
100
C
10
15
settlement (mm)
20
25
x0
Ax5D
600
'500
a.
x=1 OD
400
.1-
m
300
U
.1-
200
100
10
15
settlement (mm)
20
25
Collause of Undrained Subsoils
70
-x0
-Ax5D
60
,50
x1 OD
B-
SI..
40
1S
.
0
30
.4I-
S
)
20
10
0*
0
10
15
settlement (mm)
20
25
-.x0
-Ax5D
-xlOD
60
50
ag 40
S
I.4ii
30
U
.4I-
> 20
10
10
15
settlement (mm)
20
25
For soil of uniform strength and limited depth the distribution of total vertical
stress for both the rough and the smooth footings are shown in
non-dimensional form, together with the distribution predicted by plasticity
theory, Figure 4.9. The corresponding distributions for the case of strength
increasing linearly with depth is shown in Figure 4.10.
Load Controlled Analyses
In addition to the displacement controlled analyses, load controlled analyses
were performed in which the ideal plasticity distributions were applied in
increments at the ground surface. The load was applied in 100 equal
increments, i.e. 1 % of the failure load predicted by plasticity theory was
applied in each increment. Only rough cases were examined.
As the applied load reached the plasticity load, the incremental
displacements increased rapidly, Figure 4.11. An attempt was made to
increase the load above that predicted by the plasticity solution by 5%
applied over 50 increments, but the analyses failed due to numerical
problems. The settlements profiles under full plasticity load are shown in
Figure 4.12 for soils of uniform strength and limited depth and soil of
increasing strength with depth.
CollaDse of Undrained Subsoils
25
20
15
In
10
5
01
0
5
x/D
I
10
10
5
0
10
/x/SLb
100
90
,- 80
,70
260
>.
4C)
4m
a 40
a.
0
U
220
10
0
10
10
15
20
settlement (mm)
100
90
- 80
.a70
U
260
o 40
a.
U
220
10
0
20
10
10
20
settlemeni (mm)
30
40
40
30
?20
E
.10
Ew
.11-
110
10
20
20
10
10
20
30
40
20
30
I
- 6
I
0
10
12.
30
25
20
:10
U)
15
10
5
C
10
x/D
the number of rows of linear strain triangles was increased, from three to
four, the vertical stress distribution at failure still over-predicts the plasticity
solution by a large amount, 24% at xID=5, even though the number of
degrees of freedom is similar to the original mesh including interface
elements.
This result shows that the interface element can be used to analyse,
accurately and economically, the behaviour at rough soil boundaries. The
interface element also allows interface shear and normal stress information
to be extracted easily from the analysis.
Towards the centreline the finite element displacements are larger than the
plasticity displacements (Figures 4.15b, c and d). This may have been
caused by the rigid area near the centreline in the hodograph, Figure 4.14b,
being larger than that predicted by plasticity theory. This over-prediction of
the extent of the rigid zone is a result of the relatively coarse slip line field
which was used for the graphical solution, Figure 4.14a. If a finer slip field
had been drawn, then the vertex of block 1, Figure 4.14a, would have
coincided with the centreline of the footing, as predicted by plasticity theory.
This would have reduced the extent of the rigid block and increased the
lateral displacements predicted from Figure 4.14b.
At a distance 20 from the toe, Figure 4.14a, the trend reverses and the
plasticity displacements are larger than the finite element displacements.
Two factors may have contributed to this. Firstly, the undrained behaviour is
modelled by using a high value for the bulk modulus of water, but any
non-infinite bulk modulus will result in some volumetric strain which will tend
to produce smaller lateral displacements. This error is cumulative, becoming
more significant with distance from the centreline. Secondly, if some of the
soil was still elastic at increment 90 then a reduced lateral displacement
would result.
73
0
0
2
I\
4]
S
40
80
120
stress (kPa)
150
ID
20
30
40
effective
vertical
stress
'\
-- pore water
\ \
pressure
D 81
101
121
\'
Figure 4.16 - Stress and strength profiles used for the modified
Cam-clay analysis.
The analysis was conducted for a rough footing resting on a soil with
strength increasing linearly with depth, the material parameters, using
standard notation, were X=0.25, ic=0.07, f=3.0, M=1.0 and v'0.3. The
predicted stress and strength profiles, Figure 4.16, were defined according
to the equations developed in Appendix A. An error in the strength profile of
Figure 4.16, pointed out by Potts and Ganendra (1991), is discussed below.
The analysis was performed using the mesh shown in Figure 4.6. Additional
interface elements were included at the surface to model the rough
boundary and the interface strength was s 0 =4.25kPa. The undrained shear
strength was calculated to increase at a rate p=2.7OkPaIm. The load
settlement curves for three points at the subsoil surface are shown in Figure
4.17. Failure was defined to have occurred after 600 increments with a
displacement of 0.06m.
At failure the vertical stress distribution beneath the footing can be
compared with the plasticity solution, Figure 4.18. The prediction made
using the modified Cam-clay model is in close agreement with the plasticity
74
90
80
,70
U
0
.60
j50
m 40
U
C,
30
> 20
10
0
0.1
0.2
0.3
0.4
settlement (mm)
0.5
0.6
25
20
,15
10
5
0
0
4
ox/Su
.....................................(4.1)
The two predictions of the undrained shear strength profile, Figure 4.16 and
Equation 4.1, are shown in Figure 4.19. Also shown is the percentage error
between the two predictions. The maximum error is 22% at the surface and
rapidly decreases to less than 5% beyond 2m below ground level.
A best fit line through the profile predicted by Potts and Ganendra produces
values of s=4.13kPa and p2.6lkPa/m (compared with the values
s=4.25kPa and p=2.7OkPaIm used in Figure 4.18). If these values are
used to replot the data in Figure 4.18 a negligible difference occurs in the
position of the data points.
The effect of the over-large interface strength (4.25kPa rather than
3.49kPa) appears to have had little effect on the final result The analysis
76
4
E
1a.
0
8
10
12
0
10
15
20
25
30
35
40
Figure 4.19 - Error in the undrained shear strength profile used for the
modified Cam-clay analysis (after Potts and Ganendra, 1991).
showed that slip occurred on the interface beneath the loaded area even
though the shear strength was lower at some integration points in
neighbouring soil elements. This implies that yielding was constrained in
those elements to some degree due to the discretization error at the
surface.
for
FSyH
suo
FS-yH
SUD
design curve
px/Suo
x/D
design curve
(a)
H fill
(b)
Hfill
Hd
Hd
7-
x
practical crosssection
(c)
Xd I
x
practical crosssection
(d)
constant side, slope
practical crosssection
78
Figure 4.20a and b. The plasticity curve can then be used to define the
minimum safe single stage side slope length, Xd, Figures 4.20a and b, and
cross section, Figures 4.20c and d.
The embankment shape which corresponds to plasticity theory involves a
uniform slope within the plasticity region (x>O) and a stable, steeply sloping
wedge beyond (x<O). Jewell (1988) compared limit equilibrium analyses of
the plasticity cross sections (Figures 4.20c and d) with the p'asticity
solutions. Limit equilibrium analysis of subsoil with strength increasing
linearly with depth were in good agreement with the plasticity curve but limit
equilibrium analysis of the uniform strength and limited depth case
exceeded the plasticity distribution by a large amount, this difference is
discussed in Section 4.3.2. The embankment shape suggested by the
plasticity cross section would be difficult to construct and would take
additional land due to the triangle of soil required to avoid a vertical soil face
at the toe.
Also shown on Figures 4.20c and cj is a proposed embankment profile with
a constant side slope which starts at the toe (x=O) of the plasticity solution
and at a distance Xd intersects the plasticity profile at the required design
height, Hd. This profile would be easier to construct and would require less
land than the plasticity cross section suggested by Jewell (1988). In order to
investigate the proposed constant side slope profile a series of finite
element analyses was carried out
Finite element analyses were performed for fully reinforced embankments
with constant side slopes on both types of idealized subsoil. The results of
these analyses were compared with the plasticity solution to assess the
validity of a design method using the plasticity distribution to predict the
slope length of a constant side slope embankment.
79
80
450
400
350
u 300
0
0
U
250
200
150
100
50
0
350
50
250 150 50
settlement (mm)
150
250
350
Figure 4.21 - Typical load displacement curve for a constant side slope
analysis. Curve shown for a uniform strength limited depth subsoil
analysis with XdID=5.
4.3.2. Results
The failure of the analysis was assessed from load-displacement curves for
three points on the subsoil surface; these were located beneath the toe, the
crest and the centreline. Typical curves are presented in Figure 4.21.
Clearly defined failure loads could be assessed for all analyses as the load
acting at the increment after which large displacements occurred at one or
more of the observed points.
For each of these analyses the vertical stress applied beneath the crest was
normalised with s and plotted against the plasticity solution, Figure 4.22
and Figure 4.23 for the uniform strength and limited depth case and the
strength increasing linearly with depth case respectively.
81
CoIIaDse of Undrained Subsoils
25
20
.15
(I)
10
5
0
6
5
4
x/D or xd/D
10
30
25
20
15
10
5
0
- Plasticity theory
0
10
px/S or pxJS
-....
- -
84
85
4.4. Summary
A modified version of the finite element program CRISP has successfully
predicted the undrained collapse of a strip footing on two idealized soil
profiles using an elastic-perfectly plastic constitutive model. Both rough and
86
87
89
3D view
Plan View
arrows indicate
direction of flow
(c>\
(a
-elk-'II--II-(b
Figure 5.1 - Unit cell and direction of flow: (a) axisymmetric conditions
(b) plane strain conditions.
consolidation analysis for subsequent use in a stability calculation. Despite
its merits, however, the finite element analysis of vertical drains has only
occasionally been reported (Zeng et al, 1987; Kumamoto et al, 1988;
Sanchez and Sagaseta, 1990).
In this Chapter a methodology for the finite element analysis of ground
incorporating vertical drains is developed. An obvious difficulty is that most
numerical analyses of embankment problems are conducted for plane strain
conditions, whereas the soil around an individual vertical drain is more
appropriately modelled as axisymmetric, Figure 5.1. It is necessary,
91
Applied stress
NL
Drain
Smear zone
Figure 5.2 - Unit cell adopted for analytical solution (After: Hansbo,
1981).
therefore, to find a logical basis for the plane strain analysis. As previous
approaches to this problem have left scope for improvement, Section 2.2.2,
the equivalence of plane strain and axisymmetric consolidation has been
re-examined and a new matching procedure developed.
92
The numerical analyses have been performed using the modified version of
CRISP, Chapter 3, using the drainage element, Section 3.4, to model the
vertical drains. Validation of the analysis was undertaken in two stages.
Firstly, the results of axisymmetric analyses were compared with Hansbo's
(1981) closed form solutions for consolidation around a single vertical drain.
Secondly, comparative plane strain and axisymmetric analyses were
conducted to test the matching procedure under the simplified conditions
assumed in its derivation. In Chapter 6 the matching procedure is tested
under more realistic ground conditions.
93
The available theories vary in their degrees of rigour and complexity. The
most rigourous, but mathematically most complex, solutions are based on
Blot's (1941) consolidation theory and were derived for equal strain
boundary conditions by Yoshikuni and Nakando (1974) and Onoue (1988).
Both solutions allow for combined vertical and radial drainage in the soil and
finite drain permeability (well resistance), but only Onoue's solution
considers smear. Although the upper boundary is displaced uniformly,
variation of vertical strain is permitted with both radius and depth. Less
rigourous solutions, allowing for both well resistance and smear but
neglecting vertical flow in the soil, have been obtained by Barron (1948),
Hansbo (1981) and Zeng and Xie (1989). These are genuine equal strain
analyses where the vertical strain is assumed to be uniform with radius and
depth; the latter is clearly an approximation if well resistance is significant.
The solutions of Hansbo and Zeng and Xie are relatively easy to compute
compared with either Barron's solution or the more rigourous equal strain
solution. Solutions for free strain boundary conditions have been obtained in
the absence of well resistance (Barron, 1948) and even then tend to be
complex.
It has been shown (Hansbo, 1981; Zeng and Xie, 1989; Onoue, 1988) that
Hansbo's relatively simple theory compares well, not only with the
philosophically similar solutions of Barron and Zeng and Xie, but also with
more rigourous solutions. Being also simple to apply, the theory has gained
wide acceptance and has been used in parametric studies (Jamiolkowski et
al, 1983). It has therefore been used to validate the finite element analysis.
Under an instantaneous step loading the average degree of consolidation
based on pore pressure, tih , on a horizontal plane at depth z and at time t is
predicted by Hansbo to be
Vh=1_eTI41(5.1)
94
orizontally
estrained,
rn pervious
hon
resi
dral
eler
dral
bou
i.Om
I .JFE1
tica Ily
restrained,
imoervious
Figure 5.3 - Finite element mesh used for unit cell analyses comparison
with Hansbo (1981).
100
90
c8
c 70
0
1U
-a
0
VI
c 50
0
(3
40
30
20
10
0
0.01
0.1
10
time factor
a
-o
0
11
c 50
0
0
40
30
20
10
0
0.01
0.1
10
time factor
100
finite element
--Honsbo
-*Yoshilcuni
90
c 50
0
U
.40
30
20
10
0
0.01
0.1
'U
time factor
100
-R-
B0
finite element
-Hansbo
70
Yoshikuril
90
60
C
0
U
1-
50
40
L
0)
.
30
20
10+-
0.01
0.1
'U
time factor
(a)
fixed in x
fixed
drainc
eleme
or dn
bound
=20.Om
___] in y
R=1O.Om
(b)
R=1O.Om
Figure 5.5 - Finite element meshes used for unit cell analyses:
comparisons with Jamiolkowski et al (1983).
99
for a fixed drain spacing. In this case the analysis could be conducted more
economically by using a single row of elements, Figure 5.5b, whose upper
and lower boundaries were subject to the same conditions as those applied
to the corresponding boundaries of the unit cell. To model the effect of
smear, soil elements near the drain were reduced in size and given a
permeability 10 times lower than the main body of soil (i.e. k/k5=10). An
increase in the drain radius was also used to provide a value of s=rJr in the
desired range.
Results for these three sets of predictions are presented alongside the
theoretical predictions in Figures 5.6, 5.7 and 5.8. In each figure the time
factor for 90% consolidation at mid-depth of the drain, Th , is expressed as
a ratio of the corresponding theoretical factor for an ideal case in which
there is no smear and no well resistance. Once again the agreement
between the finite element analyses and the theoretical results is generally
excellent. The trends shown in Figures 5.6, 5.7 and 5.8 are in close
agreement with those shown by Jamiolkowski et al (1983). This is true
despite a subtle difference of approach in the latter case, namely that in
order to establish ThgQ the average degree of consolidation was not just
evaluated at the mid-depth but over its entire length.
As noted above, the finite element results were obtained under free strain
conditions yet have been compared with an equal strain theory. It is of
interest to plot the displacements of the upper boundary of the mesh to
examine how far the boundary conditions for the finite element analyses
deviate from the theoretical ones. The development of the surface
displacement with time for three representative analyses is given in Figure
5.9; the settlements are expressed as proportions of the final settlement,
which itself is uniform across the unit cell. The only significant
non-uniformity occurs temporarily above the unusually thick smear zone in
Figure 5.9c. The results explain why very similar solutions were obtained for
free and equal strain conditions by Barron (1948).
100
4
3.5
3
U
0
01
I-
25
0
0i
0.5
B
7
6
a
-05
0
014
03
01
I-
2
1
0
10
15
20
25 30 35
drain length (m)
40 45
50
101
B
7
2
1
1.2
1.4
1.6
1.8
2
2.2 2.4
smear zone radius/drain radius s
2.6
2.8
The spatial variation of the vertical strain within the unit cell, neglected by
Hansbo, is also readily obtainable from the finite element analyses. Figure
5.10 shows the variation obtained on the periphery of the unit cell at various
stages of an analysis with high well resistance, when the largest variation of
strain might be anticipated. The strain is expressed as a proportion of the
final strain, which was uniform with depth. It is demonstrated that the
variation of strain can be significant, although its effect on the overall rate of
settlement, as given by Hansbo's solution, appears small.
102
-10
Th=0.0025
-Th=0.025
g-2o
C
1-
m-
Th=0.1 25
-9-
Th=0.5
-Th=2.5
-50
C
-60
1:;:
-90
-100
-10
g-20
Th=0.025
4-
C
S-
Th=0.1 25
-9-
-40
Th=0.5
-50
Th=2.5
-ATh=1 2.5
-70
-80
-90
-100
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
r/R
0
Th=0.0025
-iTh0.025
-10
TIi=0.1 25
-aTh=0.5
-40
I,
.
C
-50
-N-
Th=2.5
-60
Th=12.5
-90
-100
U U.1 U.Z U..
- ThaxR2
ax
where 2B is the drain spacing in plain strain and the subscripts (or subscript
extensions) p1 and ax identify the plane strain or axisymmetric conditions
respectively. Either or both the drain spacing and the soil permeability may
be manipulated to satisfy Equation 5.3. Unfortunately, the resulting values
do not then apply for other than the chosen degree of consolidation, since
104
a
Th=O.0025
-Th=0.025
-0.1
-0.2
Th=O.1 25
-aTh=0.5
-0.3
-0.4
Th=2.5
-ATh12.5
-0.6
-0.7
-0.8
-0.9
10 20 30 40 50 60 70 80 90 100
vertical strain/final vertical strain (%)
Figure 5.10 - Variation of vertical strain with depth at unit cell periphery
in finite element analysis with well resistance (q/k=150m2).
plane strain and axisymmetric unit cells, equality of the average degree of
consolidation at every time and at every level in the cell is required. Hence
105
(5.4)
from Equation 5. I
IELI
II hpI
(55
J.Lax
or
= C h t
(5 6
B2 .Lpj R2.Lax
Two types of matching are defined, firstly, geometry matching in which the
permeability of the soil is the same in the plane strain and axisymmetric
analyses and the width of the plane strain unit cell is adjusted. If the soil
parameters are identical in the axisymmetric and plane strain unit cell, then
Equation 5.6 becomes
B2 gg=R2 .t
........................................................................(5.7)
and
^*ln(s)_]
= (R2_-_2!(2!z_z2)
q., Q)
(5.8)
........................................(5.9)
or
1In()+*ln(s)_
............................................... (5:10)
106
or
Qww2j .......................................................................(5.12)
A second procedure, referred to as permeability matching, is to maintain the
same drain spacing in the plane strain and axisymmetric analyses and to
adjust the plane strain horizontal soil permeability to achieve a matched
analysis. From Equation 5.6 with B=R
.................................................(5.13)
substituting for the factors
and i.La(
2k
3[ln()+.ln(s)_..]
and
Qw = ..........................................................................( 5.15)
that soil elements near the drain remained sufficiently small. The previously
adopted values of the elastic parameters (E' and u') and horizontal
permeability of the soil were retained, except when permeability was
changed in accordance with the matching rules. As before the drain was
assumed to possess no stiffness. In analyses with smear, the size of the
smear zone and its permeability were fixed so that s=5 and kII=2. In the
analyses with well resistance, the discharge capacity was chosen to give
L=2.
The equivalence of these plane strain and axisymmetric analyses can be
assessed on the basis of both settlement and pore pressure. Results for
the overall average degree of consolidation, Ti, calculated from the
settlements of the surface nodes are given in Figure 5.11. The style of
calculation was the same as that used to compute Ti,, at a given depth from
nodal pore pressures as described previously. The corresponding results for
tih at the mid-depth of the drain are shown in Figure 5.12. For both Ti and
Uh it
0
10
Axisymmetric
Permeability matched
60
S
!
Geometry matched
-70
o 80
90
100
0.01
0.1
IU
Time factor Th
i :
Permeability matched
50
-60
S
-70
Geometry matched
0 80
90
100
o.01
10
0.1
Time factor Th
0
-10
Axisymmetric
Permeability matched
-60
0
f -70
Geometry matched
-80
-90
-100
0.01
0.1
I',
Time factor Th
Axisymmetric
Permeability matched
-50
-60
Geometry matched
1:;:
-90
-100
0.01
0.1
'U
Time factor Th
Axisymmetric
70
60
matched
50
40
U
Geometry matched
30
0)
U
20
10
a
b.0 1
0.1
'U
Time factor Th
70
-a 60
Permeability matched
50
40
30
Geometry matched
DI
20
10
C
O.01
0.1
IU
Time factor Th
100
90
80
Axisymmefric
70
matched
a 60
50
40
Geometry matched
! 30
Dl
o 20
10
0
0.01
0.1
'U
lime factor Tb
Axisymrnetric
matched
50
40
Geometry matched
30
o 20
10
0
0.01
0.1
10
Time factor Th
10
F-'
'-F
x
a
..
10
15
0.01
0.1
10
TImB factor Th
10
F'
'-F
x
a
10
15
0.01
0.1
10
Time factor Th
axisymmetric analysis (Figure 5.9a) and clearly violates the equal strain
assumption on which the matching theory is based. The improvement of the
matching procedure obtained with well resistance, most marked at small
time factors, corresponds to the attainment of slightly more uniform
settlement profiles, Figure 5.14b. Fortunately the differential settlements
seen in these unit cell analyses are less likely to occur in full embankment
analyses, because of the stiffness contribution of the embankment
elements, and superior matching can be anticipated.
21n(n)-1
.....................(5.15)
- 3(ln(n)_)
From Equation 5.15 the ratio of the pore pressures at the periphery of the
test unit cell. Figure 5.5, is 0.72. This compares well with the finite element
predicted value of 0.68, from Figure 5.15. Equation 5.15 can therefore be
used to predict the pore pressure which would develop in an axisymmetric
unit cell from plane strain analysis results.
A notable difference in the theoretical solution (Barron, 1948; Hansbo,
1981) and the finite element solution to the consolidation of the unit cell is
the excess pore pressure distribution at low time factors. As pointed out by
114
-10
gzo
4-
C
S-
Th=0.75
-aTh=1.75
-40
-50
-N-
Th=3.75
C
11-
-60
1:;:
-90
-100
-10
g-20
Th=O.25
4-
C
0
Th=0.75
-B-
Th=1 .75
. -50
-N-
Th=3.75
C
1-
1:;:
-90
-100
115
60
o 50
0.
40
0.30
m
0
-20
1
S
a
x
Lii
0.1
0.2 0.3
0.7
0.8
0.9
100
90
a
-
80
70
m
I.
III
m
L
a.
L40
30
20
10
0
0.1
the soil. Specimen results for tih at the mid-depth of the drain are shown in
Figure 5.17 and indicate that the introduction of vertical flow does not
invalidate the matching. In fact the matching errors in 1h in Figure 5.18, are
significantly reduced. This can be explained by the action of the vertical flow
117
100
go
Axisymmetrl
c 50
0
U
40
30
20
10
0-I-
0.01
0.1
'U
Time factor Th
100
90
c8O
70
0
60
50
40
20
10
0.01
0.1
10
Time factor Th
10
a
6
4
9 -2
a
-4
-6
-a
-10
-12
0.01
0.1
10
Time factor Th
Figure 5.18 - Errors in geometry matched plane strain analyses for unit
cells with vertical permeability (k,=kh) based on excess pore pressure.
119
5.4. Summary
A methodology has been developed for representing vertical drains in plane
strain finite element analyses of embankments on soft ground, and has
been validated by analysing the consolidation of soil around a single drain.
The methodology includes the use of a drainage element to represent the
vertical drains, Section 3.4. The performance of CRISP has been checked
under axisymmetric conditions against the theoretical solutions of Hansbo
(1981) and found to be satisfactory. The trends obtained by Jamiolkowski et
al (1983) in a parametric study of welJ resistance and smear have been
reproduced successfully.
A procedure has been derived which permits equivalent plane strain and
axisymmetric analyses to be conducted. Throughout the analysis, the
average degrees of consolidation are matched at every level in the soil,
although the pore pressures at corresponding points are not the same. The
equivalent plane strain analysis can be achieved by manipulating the drain
spacing andlor the horizontal permeability of the soil. The validity of the
proposed matching procedure has been examined for drains with and
without well resistance and smear, installed in a uniform soil with linear
compressibility characteristics. The procedure is generally successful; at no
time does the error in the average degree of consolidation, either at the
mid-depth of the drain or over its whole length exceed 11%.
In practice the soil behaviour is likely to be non-linear and to vary with
depth. Also in a full embankment analysis significant lateral strains may
occur in the soil. Chapter 6 describes analyses of a case history in which a
preloading embankment was constructed over a soil into which vertical
drains had been installed. The matching procedure is used to model the
vertical drains in these more realistic soil conditions and comparisons of
observed and finite element predicted behaviour are made.
120
121
122
yellow to brown
medium dense
sand and silty
sand
110
o 20
a,
light bown
soft silty day
.0
0.
U,
-D
30
brown dense
silty sand
0
-4
-8
-12
E
16
0
20
-24
-28
-32
100
200
300
Stress (kPa)
400
500
600
VALUE
Compression Index
C, = 0.30 - 0.45
Triaxial compression
4'
= 29
4',
= 32
4)'
4)'
PARAMETER
Horizontal permeability
Vertical permeability
Total unitweight
y= 18.6kN/m3
w, = 35-40%
Liquid limit
LL=50 - 60%
Plasticity index
P1=25 - 40%
The cross section and plan view of the embankment are shown in Figure
6.3a and b and the rate of construction in Figure 6.3c. Four types of vertical
drain were installed in the clay stratum, Figure 6.3b. The finite element
analyses concentrate on the Geodrain area, although little variation in
performance of the four zones was observed (Hansbo et al, 1981). The
embankment was constructed at a constant rate to a height of 5.5m in a
period of 3.5 months; this implies a load due to the fill of 99kN1m 2 beneath
the crest (unit weight of fill l8kN/m3 , Jamiolkowski and Lancellotta, 1984).
330m
Jetted
Soildrain
Geodrain
Sandwick
(a)
30m
5.5m
66m
(b)
___ 120
E
80
-D
D 40
0
a,
>
00
b
12
1978
1977
6.2.4. Instrumentation
The trial embankment was heavily instrumented with settlement plates.
several types of piezometer and vertical and horizontal inclinometers
(Garassino et al, 1979). Field measurements were taken during the
construction and consolidation stages.
drained
ho rizoni
restrain
d ra i nag
element
or drair
boundar
rntally
med
drain rad
O.031m
tically restrained
drained
R=2.Om
127
Value
A.
0.16
0.032
2.58
1.16
M1
0.92
v'
0.3
4.lxlO9mIs
Ic
3.5xi 010m/s
128
0
Theory
A
Finite element
b0
LI,
E -15
0
.o-20
.r
.1-
0.
ci-25
30
35
0
10
20
30
40
50
60
70 80
Theory
A
Finite element
b0
E 15
0
-C
-I-
a
0-25
30
35
0
10 20 30
40 50 60
70 80
at a constant rate over 50 equal time increments totalling 3.5 months. The
consolidation was modelled using 50 equal time increments totalling 10
months.
The first analysis used drainage elements, Section 3.4, to itode1 the vertica
drain, with a discharge capacity of l4Om3Iyear corresponding to the
minimum likely value (Holtz et at, 1991). The second modelled the dra n as
infinitely permeable by setting the excess pore pressure to zero at the left
hand mesh boundary, Figure 6.4. The average surface settlement for each
analysis is plotted against time in Figure 6.6; as cai be seen even The
lowest likely discharge capacity has a negligible effect on the rate of
consolidation. The maximum hydraulic gradient in the drain, used for the
first analysis, occurred transiently near the surface and was equal to 0.11.
As the actual discharge capacity is likely to be significantly higher than that
assumed, the effect of well resistance can be ignored and the vertical drain
modelled as a boundary with zero excess pore pressure.
In the remainder of this Chapter the analysis without vertical drainage
elements will be referred to as the axisymmetric analysis.
131
?ioa
E
'-200
E 300
41-
.0
U -
a
1L
a
800
900
6
8
time (months)
10
12
14
The settlement of the top right hand corner of the unit cell is compared with
observed values in Figure 6.8.
During construction the agreement between predicted and observed pore
pressures and settlements is very good, therefore validating the choice of
material parameters. However, during the ten month consolidation period,
although the predicted settlement is still very close to that observed Figure
6.8, the predicted pore pressures show a marked deviation from the
monitored values, Figure 6.7. The finite element predictions of both pore
pressures and settlements are consistent, with the settlement reaching a
final constant value as the excess pore pressures approach zero. This trend
is not seen in the observed values. Although the observed settlement
appears to be levelling off, implying the end of primary compression, there
is little dissipation of the excess pore pressure after the end of construction.
Jamiolkowski and Lancellotta (1984) suggested that the observed pore
132
70
60
U
0
50
I-
[40
30
0
aI,
ci
x
10
0
0
- observed
6
8
time (months)
10
12
14
A finite element
Figure 6..7a - Excess pore pressure 19.7m below ground level on the
centreline of the embankment.
80
70
U
a.60
50
a-40
-30
U
20
x
10
0
0
observed
6
8
time (months)
10
12
14
A finite element
Figure 6.7b - Excess pore pressure 12.6m below ground level on the
centreline of the embankment.
133
0
- 100
observed
20D
o --500
60O
" 700
0
8O0
F-900
1000
0
6
B
time (months)
10
12
14
pressures during the consolidation stage may have been unreliable, and
stated that the unsatisfactory long term performance of the piezometers
may have been due to the presence of organic gas. However, the
malfunction of the piezometers may have masked the effect of phenomena,
such as destructuring of the soil (Burland, 1990), which would also have
resulted in higher observed excess pore pressures than predicted by the
finite element analysis. As the monitored excess pore pressures during the
consolidation stage may be in error, the material parameters used for the
axisymmetric analysis were retained in further analyses.
From Figure 6.7 it is interesting to note that the pore pressure at 19.7m
below ground level is lower than at 12.6m below ground level. This
behaviour appears incorrect as with an infinitely permeable drain the rate of
consolidation of an isotropic material should be the same at every depth,
and the effect of the small vertical permeability would imply lower pore
134
60
a-
o 30
a1
0
U
fn
x
0
0
110
00
0
6
8
time (months)
10
12
14
Figure 6.9 - Unit cell rate of consolidation based on the average excess
pore pressure at mid-depth.
0
_i00
..- 200
C
0
E 300
0
-I-I-
400
500
1-
L
11
600
o-700
U
I-
-BOO
900
6
a
time (months)
io
12
14
1
0
I0
'--2
IDI
c 3
CI
4U
E 4
5
6
0
6
8
time (months)
10
12
14
matched analysis has been plotted as the difference between the two plane
strain analyses was too small to be noticeable.
For both settlement and pore pressure the agreement between
axisymmetric and plane strain analyses is very good. It is possible to define
matching errors as
AVERAGE EXCESS PORE PRESSURE
(ERROR) pwp - DIFFERENCE IN PREDICTED
100 ........(6.1)
APPLIED LOAD
....................(6.2)
These errors are plotted against time in Figure 6.11 from which it can be
seen that the matching error is never greater than 6%.
137
drair
sand
I loads
clay
drair
sanc
restrainea
horizontally and
verticall v restrained
restrainea
138
modelled using vertical loads applied in the same number of increments and
time as for the unit cell analyses.
6.4.2. Material Parameters
Clay Layer and Drains
The material parameters for the clay layer were as defined in Table 6.2. The
drains were modelled as lines of nodes at which no excess pore pressure
could develop, thus representing an infinitely permeable drain (Section
6.3.4).
Sand Layers
Only descriptive information was available for the upper and lower sand
layers and it was therefore necessary to adopt typical parameter values to
model these materials. The layers were modelled as drained
elastic-perfectly plastic materials for which five parameters are required: the
cohesion (c'), the angle of shearing resistance (4'), the Young's modulus
(E), the Poissons ratio (v') and the unit weight (y). The cohesion was taken
to be zero, c'=O. The angle of shearing resistance was defined as 4'=44.4
following Jewell (1990). A typical value for the Poissons ratio under drained
conditions was used, v'=0.2, (Lade, 1977). The unit weight of the sand was
assumed to be the same as for the clay, i.e. y=18.6kNIm3. Again following
Jewell (1990), the Youngs modulus for the soil was derived from an
empirical relationship proposed by Hardin and Black (1966) for the initial
elastic shear modulus of soil, G1,
Gj=7OO(27() ...(6.3)
where s r=lOOkN/m2 is included for dimensional consistency,s' is assumed
to be equivalent to the mean effective stress (p') for plane strain conditions
and e is the voids ratio, assumed equal to 0.5. Equation 6.3 was used to
evaluate an initial shear modulus, G 1. In order to use a shear modulus more
139
0
E
E
-I-
a,
Ea,
-I-I-
a,
a,
100
-300
-400
-500
'II-
-600
"3
-700
a,
C
a,
-200
a,
U
U
observed
-800
I-
4C
a,
U
-900
-10004
0
6
8
time (months)
10
12
14
appropriate to the strain levels that may develop in the embankment before
yielding, a value for the shear modulus of G=0.5G1 was used. The Young's
modulus was then related to the shear modulus as E=2(1+v')G.
Three sand layers were defined, namely: z=0-7.5m, 29-42m and 42-62m
below ground level and the Young's modulus values in these layers were
5.9x104, 14.5x104 and 17.4xlO4kNIm2 respectively.
0
observed
60
80
100
12a
E
x
a
E
16C
0
6
8
time (months)
10
12
14
has not reached a constant value at the end of the period considered; this
may again be explained by the continued creep of the soil which is not
modelled in the finite element analysis.
The lateral movement profiles at the end of construction and at the end of
consolidation are plotted in Figure 6.15. The maximum values of lateral
movement are in good agreement, although the observed small inward
movement of the surface has been exaggerated in the finite element
analysis. This may be a result of modelling the embankment simply using
vertical loads which does not allow the development of outward shear
stresses at the ground surface, as may occur in practice.
The inward movement cannot be explained as being caused simply by
settlement of the embankment. If the embankment was assumed to remain
the same length (33m) and the settlement increased linearly from the toe to
the centreline value of 900mm, then the toe would move inwards by only
2mm. Most of the predicted inward movement must be caused by the initial
elastic behaviour of the sand. It may be noted that for a footing moving
rigidly into an elastic layer an inward movement at either end of the footing
is predicted (Poulos and Davis, 1974).
Pore Pressure
Piezometers were placed on the centreline of the embankment at depths of
19.7 and 12.6m below the surface. The piezometers were positioned at the
centre of a triangle of drains so as to record the highest pore pressures. In
Figure 6.16 comparisons of the observed and finite element predicted pore
pressures are shown.
The excess pore pressure calculated from the finite element program is for
a plane strain analysis and as discussed in Chapter 5 the matching
procedure ensures that the average pore pressure across the axisymmetric
and plane strain unit cell are equal at every depth and at every time.
However the distribution of pore pressure across each unit cell is not the
142
0
observed
5
finite element
10
15
25
30
35
40
45
50
200 150 100 50
0
lateral displacement (mm)
100
150
A
finite element
10
15
-a
30
35
40
45
50
0
200 150 100 50
lateral displacement (mm)
100
150-
80
70
a
a-
. 60
20
U
10
0
6
8
time (months)
observed
10
12
14
A finite element
.30
observed
6
8
time (months)
10
12
14
A finite element
same and in the present case, at the periphery of the unit cell, the
axisymmetric pore pressure is only 72% of that predicted in plane strain, as
predicted using Equation B33, Appendix B. It is therefore necessary to
correct the plane strain pore pressure at the periphery of the unit cell by
multiplying by a factor of 0.72 before comparison with the observed data
can be made.
From Figure 6.16 it can be seen that the predicted excess pore pressures
during construction are in good agreement with those observed. However,
as seen in the axisymmetric unit cell analyses, Section 6.3.5, the agreement
after the end of construction is not good. As previously discussed the likely
cause of this discrepancy is the unsatisfactory long term performance of the
piezometers due to the presence of organic gas, resulting in incorrect
recorded values. However, other factors may result in an increased
observed excess pore pressure (Eg. destructuring of the subsoil) and
should ideally be considered in future analyses.
6.5. Summary
The matching procedure developed in Chapter 5 has been applied to a
normally consolidated clay deposit modelled using modified Cam-clay. The
comparisons of the rate of consolidation based on both average surface
settlement and average pore pressure at the mid depth of axisymmetric and
plane strain unit cells are good. The matching procedure has therefore been
shown to be capable of providing an accurate representation of soil
improved using vertical drains analysed in plane strain.
A full plane strain analysis of the Porto Tolle trial embankment has been
carried out in which the matching procedure was used to calculate the
spacing of the vertical drains in plane strain. The finite element predicted
vertical and lateral displacements and pore pressure on the centreline are,
with the exception of excess pore pressures after construction, in good
agreement with the observed values. The study shows that with accurate
145
soil parameters obtained from high quality field and laboratory tests and
with careful finite element modelling accurate predictions of the behaviour of
soils improved with vertical drains can be made.
146
7. Multi-Stage Embankment
Construction
7.1. Introduction
The maximum single stage height of an embankment, with a given side
slope length, built on soft clay can be estimated using plasticity theory (as
described in Chapter 4). If the required embankment height is greater than
this then a multi-stage construction approach could be used in which the
subsoil is allowed to consolidate and gain strength between each, of
several, loading stages. Vertical drains are often installed in the subsoil to
increase the rate of consolidation and therefore shorten the required
consolidation stages.
The design of a multi-stage embankment requires the accurate prediction of
the strength increase of the subsoil and the application of the correct
analysis method to each loading stage. Three approaches have been
summarized by Ladd (1991).
I Total stress analysis - Construction and failure are assumed to occur
in a sufficiently short time so that the strength increase of the subsoil
is negligible. The shear strength of the subsoil is based on
unconsolidated undrained laboratory tests or field vane shear tests.
This method cannot take account of the increase of shear strength
due to consolidation and is only applicable to a single lift
embankment analysis.
2 Effective stress analysis - Effective strength parameters are
measured in consolidated drained laboratory tests and insitu
measured pore pressures are used to calculate the available shear
strength along a potential slip surface. The implicit assumption is
147
Private communication.
149
6.5m
1O.Om
1.2m
2.6m
1O.Om
150
30
60
90
0.8
1.2
1.6
2.0
16
24
0
2
93kPa
l9kPa/m
4-.,
0
a)
-o
8
10
(a) stress
(b) OCR
(c) strength
Figure 7.2 - Idealized soil (a) stress (b) Over-consolidation ratio (C)
Initial undrained shear strength.
excess pore water pressure beneath the centre of the embankment. This
calculation assumes that the drains have no well resistance and cause no
smearing of the soil during installation and that the subsoil undergos
one-dimensional loading.
In the finite element analysis the drains were modelled as infinitely
permeable by defining boundaries at which the excess pore pressures were
set to zero. The matching procedure, Chapter 5, was used to provide a
finite element drain spacing of 4m by modifying the horizontal subsoil
permeability to k,1=1.77x1OmIs. The positions of the drains in the finite
element analysis are shown in Figure 7.1.
Soil/Reinforcement Interfaces
In order to represent accurately the behaviour of the reinforcement it was
necessary to model the (lower) subsoil/reinforcement and (upper)
reinforcement/fill interfaces. The stiffness parameters of both interfaces
were: normal stiffness,
k=1O4kNIm2
Stiffness values in this range have been shown by Kwok (1987) to produce
satisfactory results.
The upper interface yield criterion was modelled with 4=300 and =5kNIm2
so that the full shear strength of the fill material was assumed to be
mobilised along the upper interface at failure.
The lower interface yield criterion was modelled as purely cohesive with a
value of =3.93kN/m2, equal to the initial undrained shear strength of the
subsoil surface, during the first lift and the consolidation stage. The interface
strength was increased for the final lift to represent the enhanced subsoil
surface strength. The increased interface shear strength was the average of
the undrained shear strengths in the surface elements, of each unit cell, as
shown in Figure 7.3.
152
20
16
a
.c12
.4-
a'
C
I-
-I-
I-
0
UI
6
8
10
12
Distance from foe Cm)
14
16
iB
reinforcement and
interface level
a)5
C .i-.
C
LO
4- N
C.
C
i0
..-. N
-c
position of
vertical drains
the subsoil occurred. Failure was defined as the increment during which the
increase in settlement was greater than the increase in the height of fill,
represented by the surcharge loading. This definition of failure has been
previously used by Rowe and Soderman (1987) and is discussed further in
Section 7.2.2.
Finite Element Mesh and Boundary Conditions
The mesh and boundary conditions used for the finite element analysis are
shown in Figure 7.4. Four vertical drains have been modelled at the
positions shown.
6.00
TOE
undrained
loading
5.00
CREST
E
CENTRELINE
'4.00
.
0)
w
second
lift
3.00
.
*NON
E
2.00
consolidation stage
Id
first lifI-'1
1 .00
0.00
'I
1.2-1.0 0.8 0.6 0.4 0.2 0.0 0.2 0.4 0.6 0.8
Settlement (m)
the settlement due to the loss of water from the system can no longer occur.
Further research into the change from consolidated to undrained analyses
may be required to ensure that the undrained failure predicted is accurate.
The maximum centreline surface settlement at the end of the consolidation
was 691mm, this value is comparable to an estimate of the one dimensional
consolidation settlement beneath the embankment crest, performed before
the finite element analysis, of 0.71m. This calculation was based on an
average coefficient of volume change, m=2.2x10 3m2/kN, and assuming
that 62% of the excess pore pressures dissipated during the consolidation
stage, Section 7.2.1.
The centreline settlement increased to 799mm by the end of the second lift
as a result of mainly distortional deformations, although some small volume
change did occur. At failure the final centreline settlement was 1147mm,
indicating large distortional deformations with the development of a
rotational mechanism, Figure 7.10.
The surface settlement profiles at the end of each of the three stages and at
failure are shown in Figure 7.6. Heave occurred near the toe and the
greatest settlement occurred beneath the crest. The ratio of the maximum
heave to the maximum settlement increased with the application of load.
The ratio was 0.48 after the first lift, 0.31 after consolidation, 0.38 after the
second lift and 0.68 at failure.
The lateral displacements of a section through the toe are shown in Figure
7.7. The maximum lateral movement occurs at a depth of 1-2.5m below
ground level. The lateral movement profile is typical of that observed in finite
element analyses but differs from those often observed which are
significantly more concave below the peak movement.
The maximum surface displacement is plotted against the maximum lateral
movement in Figure 7.8. The form of this curve is similar to those produced
by Jardine and Hight (1987) and Leroueil et al (1991) with the ratio of lateral
156
1000
500
E
0
1C
0
500
1000
1500
15
10
10
0
5
Distance from toe (m)
-- 2nd
ili -9-
15
20
1dhra
2
3
E
4-
a0
a
g
1200
1000
400
600
800
Lateral movement (mm)
200
1400
1200
E
E 1000
end of consolidation
4-
800
end of 2nd lift
' 60C
E
20(
200
400
600
800
1000
Maximum lateral movement (mm)
1200
158
1.60
1.40
1.20
1.00
0.80
0.60
0.40
0.20
0.00
0.00
.......................................................................(7.1)
I The finite element analysis allowed 20 days for the placing of the
first and second lifts in which time the soil adjacent to the drains
would have partially consolidated and gained strength. This would
result in a higher factor of safety as the limit equilibrium assumed
undrained loading.
2 The strength increases during the consolidation stage adopted in the
limit equilibrium calculations were significantly different to those
observed in the finite element analysis. This aspect is discussed
further in Section 7.3.
3 The limit equilibrium calculation used a factor of safety based on the
available shear strength and the applied loading. The calculated
factor of safety of 1.1, therefore, implied that the embankment height
could be increased by a factor of 1.1 across the whole section. The
finite element calculation applied a surcharge loading along the crest
and not on the side slope. The mechanism which develops in the
finite element analysis would have been different to that produced if
the embankment loading was proportionally increased at all points.
4 When predicting undrained collapse using modified Cam-clay a
large number of increments need to be used to ensure that the
collapse load is not over-predicted. The only certain method of
prediction is to perform a sensitivity study with increasing numbers
of increments until the collapse is unaffected by an increase in the
number of increments. Such a sensitivity study was not carried out
for in this case but a large number of increments were used. The
first lift failed in the 660th increment the first 200 increments were
consolidated, the second lift failed in the 690th increment with the
first 450 increments consolidated.
To compare the limit equilibrium and finite element predicted failure directly
it would be necessary to carry out a limit equilibrium analysis using the finite
160
'-
JI
I I
II
______
-
..
-------
iimii ii I
--_'-/..'MY_'//I 111111111
ISIHUSI P
45
40
35
30
25
0
20
15
IC
20
10
40
30
60
50
60
70
45
40
35
30
25
0
20
15
'C
10
20
30
40
p.
50
60
70
60
None of these elements reach the critical state but it can be seen that the
response during loading is increasingly undrained as the distance from the
drain increases. Position C, furthest from the drain, shows an almost
undrained response whilst position A, closest to the drain, shows a virtually
drained response, and is almost parallel to the K0 line during the
consolidation stage, until the undrained loading is applied. Figure 7.11b
shows the stress paths for positions close to the centreline and the general
trend of behaviour is the same as beneath the slope. Some of the soil is
approaching the critical state at failure of the embankment. These stress
paths follow a similar pattern to those observed by Almeida et al (1985,
1986) and Smith(1984).
The rotation of the principal stress direction within the subsoil is of
significance as the subsoil undrained shear strength is known to be
dependent on the orientation of principal stresses (Eg: Ladd, 1991; Hight et
al, 1987), this aspect will be discussed in more detail in Section 7.3. The
rotations beneath the crest are shown in Figure 7.12b. At position 0
principal stress directions are exchanged during the first increment with the
horizontal stress becoming the maximum principal stress. However, apart
from this, the principal stress rotations are less than 100 and conditions
beneath the crest approximate to those in a one dimensional compression
test The rotations beneath the slope, Figure 7.12a, are much larger
throughout the analysis although the rotations reduce sharply as failure is
approached. This variation of rotation across the subsoil is complicated by
the inclusion of the vertical drains but compares well with that observed by
Smith (1984).
The parameter introduced by Bishop (1966)
b= 05 ............................................................................(7.2)
1
principal stresses and a2' is the out of plane stress, ar'. Now b is redefined
163
90.0
60.0
(3
o 300
(3
.4-
a
'- 00
UI
d3
L
-U
C
-I-
U
-I-
-90.0
0.00
1.00
2.00
3.00
4.00
5.00
6.00
90.0
60.0
U
o 300
(3
-I-
a
-' 00
In
0
L
0)
0
-U
-30.0
o_
a
-4-
-4-
0
E
-90.0
0.00
1.00
2.00
3.00
4.00
500
6.Oa
164
1.0
0.5
0.0
.0
1.0
1.5
2.0
0.00
1.00
2.00
3.00
4.00
Embankment height (m)
5.00
6.00
1.0
0.5
0.0
.0
1.0
1.5
2.0
0.00
1.00
3.00
4.00
2.00
Embankment height (m)
5.00
6.00
165
as
b.=a.............(7.3)
Oi
'<c'
btb.
The variation of b* with embankment height has been plotted in Figure 7.13.
Beneath both the slope and crest during the first lift the element nearest the
drain has an out of plane principal stress that is less than the minimum in
plane principal stress. This is consistent with the soil nearest the drain
consolidating quickly, causing little out of plane stress, and the soil further
from the drain becoming increasingly undrained and undergoing stress
increases in all directions due to the increased vertical stress. During the
consolidation stage for all positions b* reaches a value between 0.0 and 0.5.
In the second lift the elements away from the drain are approaching the
failure value of b*=0.5 with the soil adjacent to the drain having a relatively
constant value. Finally the undrained loading causes all points to approach
a value of b*=O.5. This is consistent with Figure 7.11 in which the stress
paths are approaching the critical state.
The strength at the end of each increment can be calculated from the void
ratio, Appendix A, Equation A30. The ratio of the strength to the effective
vertical stress (sJc') is plotted in Figure 7.14 and the ratio of the strength to
the maximum in plane principal effective stress (sJci 1 ) is plotted in Figure
7.15. Two points are noted from these graphs; firstly, at no stage of the
analysis does either ratio fall below the normally consolidated ratio
sJcy' 1=0.218 and, secondly, the ratio sJ; ' is always equal to or greater than
the ratio s/a1' as the effective vertical stress can never be greater the
maximum in plane principal stress. These points are significant when
developing strategies for calculating strength increases.
166
0.30
0.28
- 0.26
>
b
' 0.24
0.22
0.20
0.00
1.00
2.00
3.00
4.00
Embankment heIght Cm)
5.00
6.00
0.30
0.28
- 0.26
>
b
0.24
0.22
0.20
0.00
1.00
3.00
4.00
2.00
Embankment height (m)
5.00
6.00
167
0.30
0.28
- 0.26
b
' 0.24
0.22
o.2a
0.00
1.00
3.00
4.00
2.00
Embankment h&ght (m)
5.00
6.00
0.28
- 0.26
b
(1)
0.24
0.22
0.20
0.00
1.00
2.00
3.00
4.00
Embankment height (m)
5.00
6.00
168
During the first lift the strength ratios for all points reduce until yielding
occurs. Once the soil has yie'ded the value of sJa1' is dependent on the
ratio qip'. By comparing Figure 7.11 with 7.15 it can be seen that an
increasing qlp' ratio causes an increase in s,/a' and vice versa. Points 0, E
and F are at the critical state, Figure 7.11, with a constant sJc 1' of
approximately 0.28. During the loading stages the undrained shear strength
remains virtually constant. The increase in the ratio is, therefore, a result of
a reduction in the effective stress due to increasing pore pressures.
Reinforcement and Interface Behaviour
The reinforcement strain is plotted at the end of each of the four stages in
Figure 7.16. From the profiles it can be seen that the reinforcement strain is
slightly reduced near vertical drains (at 2, 6, 10, and 14m from toe). This is
due to the soil adjacent to the drains consolidating and gaining strength
more quickly. The reinforcement strain increases during the consolidation
stage. This increase is not caused purely by the differential settlement of the
embankment but also by to an increasing lateral movement, Figure 7.7. The
reinforcement strain profile at failure shows a final reinforcement strain of
over 8%.
The interface elements can be used to plot the shear stress distributions at
the interfaces above and below the reinforcement, Figure 7.17. The sign
convention used to define the direction of the interface shear stresses is
shown in Figure 3.4.
The (upper) reinforcement/fill shear stress profile, Figure 7.17a, shows
significant oscillation of stress, particularly at failure. However, trends can
be discerned from the profiles and in all cases the shear stress on the
reinforcement is outwards for the first 8-lOm beyond the toe,indicating that
the embankment is spreading laterally and inducing a tensile strain in the
reinforcement. Beyond this (nearer the centreline) the embankment
169
0.00
1.00
2.00
3.00
4.00
! :::
7.00
8.00
9.00 +0.00
20.0
15.0
a
10.0
0
-
I.1UI
I-
a
10.1
15.
20.
0.0
2.0
4.0
6.0
10.0
8.0
Distance from toe (m)
12.0
14.0 16.0
4.0
2.0
a
UI
UI
UI
4UI
L
U
-
4.1
U)
6.
8.
0.0
2.0
6.0
8.0
10.0
Distance from toe (m)
4.0
-a I st lift
- -
12.0
14.0
16.0
III
passive
compression
- active
compression
simple
shear
.2 0.40
0
C
a)
0.30
______
U,
I-
0
a)
triaxia] compression
direct smpte shear :
0.20
U)
a)
C
0.10
0
I-
0
C
0.00 L
0.0
20.0
40.0
60.0
80.0
100.0
a1'
is known. The variation of the strength ratio with embankment height has
174
been plotted for the idealised study in Figures 7.14 and 7.15. From these
plots it can been seen that the initial strength ratio provides a lower bound
to the strength ratio during construction. However, It is difficult to simply yet
accurately estimate the value of ' and an easier quantity to evaluate is the
effective vertical stress ;'. Using this value with the initial strength ratio
must provide a conservative strength as ;'<a1 . The strength at any time
can therefore be calculated from
S()
c,
.......................(7.4.)
NC
(a) method 1
LqO
O<&Tv<ri
I
I
I
I
I
-4
-4
Xct
]
176
toe and that this angle reduces linearly towards the centreline. If the load
spreading angle was zero, this method would be equivalent to method I
(above).
Generation of Excess Pore Water Pressure due to Embankment Loading
At the end of a loading stage excess pore pressures wilt have been
generated in the subsoil. For the purpose of this investigation the simplest
possible estimation of the value of these excess pore pressure increase
(Au) has been made, i.e. that the excess pore pressure increase is equal to
the increase in total vertical stress calculated using Grays (1936) elastic
theory
AU=AY V ...............................................................................(7.5)
..........................................................................(7.8)
177
The excess pore water pressure generated during loading will reduce
during the consolidation stages. In order to calculate the effective stress
conditions, and thus the undrained shear strength, it is necessary to
estimate the amount of dissipation which occurs. When analysing a subsoil
containing vertical drains it is likely that the flow will be predominantly
horizontal and it may be appropriate to use one of the solutions for a unit
cell reviewed in Chapter 5. For this study Hansbo's (1981) solution has
been used. For the idealised two-stage analysis the vertical permeability
was negligible so that it was possible to apply Hansbo's equation without
consideration of vertical flow. In general the vertical permeability will not be
negligible and soil close to horizontal drained boundaries *ill co'.solidate
more quickly than would be predicted using purely radial flow in a unit cell
calculation. The increased rate of consolidation would lead to surface
strengthening which is important, particularly for reinforced embankments,
178
......................................................................(7.9)
eT1)
Th
(7.10)
( \
_lSjl
1+2Ko) OCR1+2Ko,)
ocIvc(1+2KoNc)
(710
(1+2k0)
Chm'y..........................................................................(7.11)
where the coefficient of volume change for the initial one-dimensional
conditions can be defined as m=AJcy,'(1+e), k is the permeability and 're, is
the unit weight of water.
The coefficient of volume change is dependent on both the void ratio and
the vertical effective stress and will therefore vary with depth. For the
idealized case described in Section 7.2, the values of m at the top and
bottom of the soft clay were 9.4x1 0 and 1 .4x1 0 3m2/kN respectively. This
181
........................................................(7.12)
where e0 and k0 are the initial void ratio and permeability respectively.
Applying Equation 7.12, to the idealized case, results in a permeability
which is five times greater at the surface than at the base. Combining the
effects of the variation of the coefficient of volume change and permeability
gives a slightly larger coefficient of consolidation at the surface than at the
base.
In the finite element analysis of the idealized case the variation of the
coefficient of volume change was automatically accounted for when using
the modified Cam-clay constitutive model. However, the permeability was
constant throughout the depth. Therefore, the upper layers of the finite
element mesh consolidated more quickly than those lower down.
In general, when performing a finite element analysis of an embankment
constructed on a soft soil it is necessary to model accurately the coefficient
of consolidation throughout the depth of the deposit. To achieve this, using
the modified Cam-clay model implemented in CRISP, it would be necessary
to define several horizontal layers of elements and define a different value
for the permeability in each.
As well as its initial variation with depth, the permeability will also display
further changes with loading. This can be considered using Equation 7.12.
The variation of permeability with stress level has been implemented in
CRISP, Section 3.4, but has not been used in the idealized analysis.
182
..(7.13)
where e0=r-1.
The undrained shear strength increases for each row of elements in each
unit cell are then averaged to give eight values of strength with depth (per
unit cell).
The matching technique (developed in Chapter 5) only ensures that the
average value of the excess pore pressure in the plane strain unit cell is
equivalent to that which would occur in an axisymmetric unit cell. The
variation of the excess pore pressure across the unit cells will be different
(Section 5.3.2) and therefore the undrained shear strength variation across
the unit cell will be different. In order to use the results from a plane strain
finite element analysis to predict undrained shear strengths which are likely
to occur in the subsoil, it is necessary to average the void ratio at various
levels in each of the un!t cells and to calculate the undrained shear strength
from these average values.
183
184
Method
Loading
dissipation in OC range
one-dimensional
no
no
Gray
no
no
one-dimensional
yes
no
Gray
yes
no
one-dimensional
no
yes
Gray
no
yes
one-dimensional
yes
yes
Gray
yes
yes
0.0
FE
-Method A
-1.0
-2.0
-3.0
Method B
-aMethod C
-4.0
4-
Method 0
-7.0
-8.0
-9.0
-10.0
0.0
2.0
1.0
6.0
3.0
5.0
4.0
Strength Increase (kPa)
7.0
8.0
0.0
-1.0
FE
-2.0
Method E
-3.0
Method F
-aMethod C
-4.0
-N-
.c
-5.0
4a
Method H
-7.0
-8.0
-9.0
-10.0
1.0
2.0
5.0
6.0
4.0
3.0
Strength increase (kPa)
7.0
8.0
0.0
FE
-Method A
-1.0
-2.0
-3.0
Method B
-eMethod C
-4.0
E -5.0
Method 0
-7.0
-B.0
-9 .c
-b.c.
0.0
5.0
10.0
15.0
20.0
Strength (kPa)
25.0
30.0
FE
-2.0
Method E
-3.0
Method F
-aMethod G
-4.0
-C
-5.0
Method H
a-6.0
-7.0
-8 .a
-9.0
-10.0
0.0
5.0
10.0
20.0
15.0
Strength (kPo)
25.0
30.0
188
5.0
_____- 10.0
0.0
__________
25.O
190
191
5.0
______- 10.0
15.0
20.0
________
25.0
5.0
#
1 u.1u
15.0
-.-----------.--
___
-
25.0
192
maxlmum=2.96kPa
mlnlmum=-O.47kPa
madmum=2.47kPa
IP%IWI
i.fl flUfl,n
193
5.0
10.0
?_15.0
-FE
20.0
Honsbo
25.0
30.0
80.0
85.0
90.0
95.0
100.0
5.0
FE
10.0
15.0
-4Onedimensional
Gray
20.0
25.0
30.01
85.0 90.0
95.0 100.0
coefficient of volume change with depth, which would have occurred in the
finite element analysis, was correctly reproduced by the simple method.
Figure 7.27b shows a comparison of the total stress increase near the
centreilne, calculated using both a one-dimensional approach and using the
Gray (1936) elastic solution with the average finite element prediction in the
unit cell closest to the centreline. The one-dimensional approach predicts a
uniform stress increase equal to the applied load, and is increasingly in
error with depth. The elastic solution represents more accurately the profile
of total stress increase with depth, but still exceeds the finite element
prediction.
After combining the predicted degree of consolidation (Figure 7.27a) and
the increase in total vertical stress (Figure 7.21 b) to predict the increase in
effective vertical stress, the undrained strength ratio, (s/a')=O.272, can
be used to predict the strength increase. These predictions are compared
with the undrained shear strengths calculated from the finite element
analysis near the centreline in Figure 7.28a. The underestimation of the
degree of consolidation causes the simplified method to underestimate the
strength at the surface. The one-dimensional calculation of total stress
increase (Method G) results in a large error in the predicted strength.
However, the superior total stress predictions using the elastic solution
(Method H) result in a better prediction of the strength increase and produce
the observed trend at the centreline.
The Gray solution has been used to calculate the stresses throughout the
soft clay layer and predictions of strength increase from both the finite
element analysis and the simplified method are presented in Figure 7.29. A
comparison of the two predictions is made in Figure 7.30 where the strength
calculated using the simplified method has been subtracted from those
predicted in the finite element analysis. The maximum error on the safe side
is 14.1 OkPa and this is due to the simplified method neglecting the effect of
vertical permeability. The maximum unsafe error is -3.37kPa and is caused
195
5.
E
lethod G
ia
lethod H
?15
20
25
30 .0
16.0
18.0
24.0
22.0
20.0
sfrengfh increase kPa)
26.0
28.0
Method C
Method H
15.0
20.0
25 .a
30.0 -I--
40.0
50.0
60.0
70.0
80.0
Strength (kpa)
90.0
100.0
_)
by the failure of the elastic solution to accurately predict the vertical stress
increase. The unsafe error is relatively small and occurs only at a depth at
which failure mechanisms are unlikely due to the increase of strength with
depth.
7.5. Summary
The idealized study has provided information on the behaviour of a
multi-stage construction of an embankment over a soft cohesive deposit. As
expected analysis showed that multi-stage construction allows a
significantly higher embankment to be constructed than would be possible
in a single lift.
Comparison of limit equilibrium and finite element analysis has proved
inconclusive. A higher factor of safety was found from the finite element
198
analysis. This was due, in part, to the consolidation which occurred during
construction, in the finite element analysis, producing strength increases
that were not allowed for in the (undrained) limit equilibrium analysis. Also
the method by which the embankment was brought to failure in the finite
element analysis was different from that used in the limit equilibrium
analysis.
The undrained shear strength increases for the idealized multi-stage
analysis have been calculated from the void ratios predicted by the finite
element analysis. Simplified calculations of strength increases have also
been made and comparisons of these methods with the finite element
analysis have shown reasonable agreement. The simplified strength
increase calculation procedures have also been applied to the Porto Tolle
case history, Chapter 6. Agreement for this case was not as good, but the
size of the error produced still indicates that the simple procedures for the
prediction of undrained shear strength increase may be used.
The good agreement achieved despite a major assumption made in the
simple methods regarding the generation of excess pore pressure during
loading. The proposed procedure used a one-dimensional approach
ignoring the pore pressures which would result from distortion. Such a
method would underestimate the excess pore pressure generated and,
therefore, would overestimate both the effective stress and the undrained
shear strength. Further investigation of this aspect is needed.
199
8.2. Conclusions
The following conclusions have been drawn from the results presented in
the previous chapters.
203
204
205
206
ApDendix A
(Al)
where a1>a2>a3.
The undrained shear strength is defined as
Su =
1"
, a1
G3
.....................(A2)
Triaxial Conditions
In a triaxial test a2 =a3 , therefore from Equations Al and A2
SU
..............................................................................(A3)
207
ADoendix A
a'
p'=(ci 1
that
3(a _ci)
M=..................................................................... (A9)
( +2a)
rearranging Equation A9 and combining with Equation A7
MTX =
6sin'(AID)
3sin4
Plane Strain
Using the definition of the stress invariant q from Equation Al and
p'(a 1 +a21+a3 )13 and substituting for the intermediate principal stress
defined in Equation A5 it can be shown that
/(a _a)
.......(All)
MpL =
(
+)
209
ADnendix A
N
N
ln(p')
if5
210
ApDendix A
(p;' (+M
2M2
From geometry and using the critical state gradient for the normal
consolidation lines and critical state lines of ?. and the gradient of the
swelling line K
i. i\ f ,\A
I PL I - I Pic
s.
.......................(P15)
iII I
j I PJ
where A=Q.K)/
Combining Equation A14 and A15
(P16)
Point L is at the critical state, therefore q=Mp', so that
=
A/7pj(tIj2
02)
qI.
A(P17')
Equation A17 defines the deviator stress at failure in terms of the initial
stresses for any normally consolidated modified Cam-clay sample subject to
undrained loading.
Triaxial Conditions
Substituting Equation A3 into Equation A17
SurxMr4'{ 2M
(/1 8)
211
App endix A
____
S UTX M
(1 +2KONC)((
6
Ovc
)2
I KONC
1 + 2KONC 2M 2
.(A19)
SUPL=MPL-J. 2M,
(420)
L )
SUpL
M (1+2I<'0) ( 1KONC 2 g
L(1+2KONC) 2ML)(A21)
3J
()
= ()...................................................................( A22)
..........................................(/\23)
and at the critical state q.=Mp'. and q.=Mp'., Equation A22 can be
rewritten as
212
Appendix A
VB
VA
In(p')
()
............(1\.24)
= ( i)
3q9.
[ (1+2Kn)ci..a ] = [ (1+2KnNr)a11A
(1+2Kn)
c4A]
(A25)
aVB
AnDendix A
(qB.'
_(qA.'i'
Hr
..aVB)
(i ^2K0)
1OCR(i +2KONc)1
avA4(1+2KnN()J
(1+2K0)
J ...................
(A27)
, I
I(1+O)
IULt((1+2IcoNc)'1
(A28)
........
where the subscripts OC and NC represent over-consolidated and normally
consolidated states respectively
Plane Strain Conditions
Similarly for plane strain conditions substituting Equation A6 into Equation
A27
(SPL) _.(SUPL) ((
+)
NC(l+2Kiwt)J[
(1+21(n)
.......
(A29)
.................................................................... (ft30)
214
A pp endix A
(A31
qcs=Mexp[1jV]( A32)
Substituting for triaxial conditions, from Equation A3, Equation A32
becomes
sul)c=fexp[l'jt]
(A33)
Substituting for plane strain conditions, from Equation A6, Equation A32
becomes
suPL =,fJexP[ r'v]
........................................................
(A34)
ADgendix A
From critical state theory it is also possible to derive an equation which
ensures zero strains in the horizontal directions for an increment of vertical
load, as
KoNC
=:011c................................................................ (A36)
where
11ONC(1+v')(lIt) ^ 3iA
3(1_2v')
=1 .....................................(A36a)
M2-11ONC
In the current research a value of G=oo has been used so that Equation
A36a becomes
_3A+JA2+4M2
....................................................(
Equations A35 and A36 can be applied for the Porto Tolle case history
(Chapter 6) where 4,'32, M=0.92 and A0.8. The coefficients of earth
pressure at rest for the normally consolidated soil are 0.47 and 0.74 for the
empirical and theoretical relationships respectively.
In all analyses presented in this Thesis the theoretical relationship, Equation
A36 has been used to calculate Analyses of unit cells with insitu
stresses defined using this relationship and loaded one-dimensionally
produced correct stress paths. However, comparative analyses using the
empirical relationship produced unrealistic stress paths.
Coefficient Of Earth Pressure at Rest for an Over-Consolidated Soil (Is)
Several empirical relationships have been suggested for the coefficient of
earth pressure at rest for over-consolidated soils, for exam Ii Schmitt (1966)
Ko =KQNC(OCR) ................................................................ (A38)
Acrnendix A
(139)
217
ADDendix B
vx= k!
.................................................................................(BI)
V- k u
where is the unit weight of water and u the excess pore pressure.
Consider a horizontal slice of soil of thickness dz. The flow in the slice at
distance x from the drain is equal to the change in volume within a block of
soil of width (B-x), so that
V=
Aooendlx B
drain
2dz
Consider next the corresponding slice of drain. As only vertical flow occurs
in the drain, the change of flow from the entrance to the exit of the slice,
dQ11is
.......................................................... (B5)
dQ2 = 7w
ax
dzdt
.....(B6)
For continuity
219
ADDendix B
dQ1 +2dQ2
=0
.....................(B7)
1w az2
1w
xo
dzdt= 0
...................................... (B8)
Rearranging
Qw(2u -o
(u
X)o2kZ2)
B9
+)
=0
(BlO)
x=O
Rearranging
( 2 u
- 2By
Qw t
Bli
(B12)
(B13)
Let be the average excess pore pressure across the section at depth z, so
that
tIB =Judx
......................................................................... (B14)
220
Aooendix B
vB2Yw[2+2k(2&z2)]
.(B15)
Now
= (TI
- -(TI v.-.j.
B27wm4^2k(2&_Z2)]1
(B17)
Integrating Equation B17 and applying the boundary condition that at t=O,
11= v.
(B18)
_____
Rearranging
t=_B2;;mvin(&)
........................................................ (B19)
where
[+ B2 (2Lz_z2)]
..................................................... (B20)
............................................................................ (B21)
221
Appendix B
v = ti o exp[_ Th ]
............ (B23)
(B24)
Ratio of Excess Pore Pressure in Axisymmetric and Plane
Strain Unit Cells
Consider a drain without well resistance or smear. It can be shown
(Hansbo, 1981) that the excess pore pressure at a distance r from the
centreline in an axisymmetric unit cell of radius R is
(L) _(r2;r)]
In
(B25)
Similarly, for a plane strain unit cell of half width B the excess pore pressure
at a distance x from the centreline is given by Equation B13 which for a
drain with no well resistance (Q - ) becomes
u=[8x_]
............................................................ (B26)
U ,k2B2
X2
-
2B2
If the axisymmetric and plane strain unit cells are matched using geometry
matching (Section 5.3) then
B= RJ(In(n)_)
(B28)
and
222
App endix B
.(B29)
k,=k
- L 'r) Upl -
,2
2R2
r.1
^ 2R2]
(B 30)
[3(ln(n)_)][j.]
................................................................ (B31)
[2ln(na)_a2+t]
(B32)
[3(ln(n)_)][2a_a2]
At the periphery of the unit cells a =1 and ignoring the small term (1/n2)
u
UpS -
21n(n)-1
........(B33)
3(ln(n)_)
223
AnDendix C
= uoe(T1)(c1)
where u0 is the initial excess pore pressure throughout the soil cylinder, Th is
the time factor for radial consolidation and
is a parameter dependent on
the radius of the drain and soil cylinder, the well resistance and the amount
of smear which has taken place on installation.
A similar equation for plane strain conditions was derived in Appendix B,
Equation B24. Both equations assume that a load has been applied
instantaneously to the top of the unit cell so that no dissipation of pore
pressure has taken place.
In practice loads cannot be applied instantaneously and a certain period for
application must be allowed. In this Appendix Hansbo's analytical solution is
modified to allow for the effect of a load applied at a constant rate, i.e. ramp
loading.
Figure Cl shows a load applied at a constant rate over of a time T so that
the final load is Q. At any time, t, the applied load, q, can be defined as
q =t
................................................................................. (C2)
....................................................................... (C3)
The excess pore pressure can then dissipate for a time t1-t.
224
ADDendix C
dq
dt
ti
= (du) e
r0
................................................................. (C4)
h = . e(10r0 dt
...................................................................(C5)
U,, =
e"')dt .................................................................(C6)
therefore
U,, = -(i - ...................................................................(C7)
Equation C7 defines the excess pore pressure developed by a ramp
loading. If a series of ramp loadings and consolidation periods are used the
Equation C7 and Equation Cl can be alternated using the calculated values
for the average excess pore pressure as input to the next stage.
225
References
References
I Abid, M.M. (1980). The consolidation behaviour of laminated clays using
the finite element method. PhD Thesis, Sheffield University.
2 Abid, M.M. and Pyrah, l.C.(1990). The consolidation behaviour of finely
laminated clays. Computers and Geotechnics, Vol. 10, PP. 307-323.
3 Almeida, M.S.S. (1984). Stage constructed embankments on soft clays.
PhD Thesis, University of Cambridge.
4 Almeida, M.S.S, Britto, A.M, and Parry, R.H.G. (1985). Centrifuged
embankments on strengthened and unstrengthened clay foundations.
Gotechnique, Vol. 35, pp. 425-441.
5 Almeida, M.S.S, Britto, A.M and Parry, R.H.G. (1986). Numerical
modelling of a centrifuged embankment on soft clay. Canadian
Geotechnical Journal, Vol. 23, pp. 103-114.
6 Atkinson, J.H. and Bransby, P.L (1981). The mechanics of soils - An
introduction to critical state soil mechanics. Pub: McGraw Hill.
7 Atkinson, J.H. (1981). Foundations and slopes. Pub: McGraw Hill.
8 Barron, R.A. (1948). Consolidation of fine-grained soils by drain wells.
Trans. ASCE, 113, Pp. 718-742.
9 Bassett, RH. (1987). Original design of the trial embankment.
Prediction Symposium on a Reinforced Embankment, King's College,
London.
10 Bassett, RH. and Guest D.R. (1990). Model and analytical
comparisons of the behaviour of reinforced embankments on soft
foundations. Proc. Int Reinforced Soil Conf., Glasgow, PP. 461-467.
11 Blot, M.A. (1941). General theory of three dimensional consolidation. J.
Appi. Phs. Vol. 12, PP. 155-1 64.
12 Bishop, A.W. (1955). The use of slip circle in the stability analysis of
slopes. Geotechnique No. 5, P p. 7-17.
13 Bishop, A.W. (1966). The strength of soils as geotechnical materials.
Gotechnique, Vol. 10, pp. 91-1 28.
226
References
14 Bishop, A.W. and Bjerrum L. (1960). The relevance of triaxial the test
to the solution of stability problems. Proc. Research Conf. on Shear
Strength of Cohesive Soils, ASCE, pp. 437-501.
15 Bonaparte, R. and Christopher, B.R. (1987). Design and Construction
of reinforced embankments over weak foundations. Transportation
Research Record 1153, pp. 26-39.
16 Bond, A. (1984). The behaviour of embankments on soft clay. MSc
Dissertation, Imperial College, London.
17 Booker, J.R. and Small, J.C. (1975). An investigation of the stability of
numerical solutions of Blot's equations of consolidation. Int. J. Solids and
Structures. Vol. 11, pp. 907-911.
18 Boultrop, E. and Holtz, R.D. (1983). Analysis of embankments on soft
ground reinforced with geotextiles. Proc. 8th ECSMFE. Vol. 1, pp. 469-472.
19 Britto, A.M and Gunn, M.J. (1987). Critical state soil mechanics via
finite elements. Pub: Ellis Horwood.
20 Brown, C.B. and King, I.P. (1966). Automatic embankment analysis:
equilibrium and instability conditions. Geotechnique, Vol. 16, No. 3, pp.
209-219.
21 Burland, J.B. (1972). A method of estimating the pore pressures and
displacements beneath embankments on soft, natural clay deposits.
Proceedings of the Roscoe Memorial lecture, pp. 505-536, Pub: G.T. Foulis
and Co.
22 Burland, J.B. (1990). On the compressibility and shear strength of
natural clays. Gotechnique Vol. 40, pp. 329-378.
23 Clough, R.W. and Woodward, R.J. (1967). Analysis of embankment
stresses and deformations Journal of Soil Mechanics and Foundations
Division ASCE , Vol. 9, SM4, pp. 529-549.
24 Cook, R.D. (1981). Concepts and applications of finite element
analysis. Pub: Wiley.
25 Croce, A., Calabresi, G. and Viglianni, C. (1973). In situ investigation
on pore pressures in soft clay. Proceedings of the 8th ICSMFE, Moscow,
Vol. 2.2, pp. 53-60.
26 Davis, E.H. and Booker, J.R. (1973). Effect of increasing strength with
depth on bearing capacity of clays. Gotechnique Vol. 23, pp. 551-563.
227
References
27 Delmas, P., Magnan, J.P. and Soyez, B. (1987). New techniques for
building embankments on soft soils. Embankments on Soft Clays, Chapter
6, Bulletin of the Public Works Research Centre, Athens, Greece.
28 Dluzewski, J.M. and Termaat, R.J. (1990). Consolidation by finite
element method in engineering problems. 2" European Speciaflity
Conference on Numerical Methods in Geotechnical Engineering, pp.
213-222.
29 Duncan, J.M. and Chang, C.Y. (1970). Non-linear analysis of stress
and strain in soils. J. Soil Mech. Fdns. Div., ASCE, Vol. 96, SM 5, pp.
1629-1 653.
30 Duncan, J.M., Schafer, V.R., Franks, L.W. and Collins, S.A. (1987).
Design and performance of a reinforced embankment for mohicanville dike
No. 2 in Ohio. Transportation Research Record 1153, pp. 15-25.
31 Fellenius, W. (1927). Erdstatische berechnungen mit reibung and
kohaesion, Ernst, Berlin.
32 Fowler, J. (1982). Theoretical design considerations for
fabric-reinforced embankments. Proc 2 mt. Con?. Geotextiles, VoJ. 3, Las
Vegas, pp. 665-670.
33 Garassino, A., Jamiolkowski, M., Lancellotta, ft and Tonghini, M.
(1979). Behaviour of pre-loading embankments on different vertical drains
with reference to soil consolidation characteristics. Proc. 7th European Conf.
Soil Mech. Brighton Vol. 3, pp. 213-21 8.
34 Goodman, R.E., Taylor, R.L. and Brekke, T.L (1968). A model for the
mechanics of jointed rock. J. Soil Mechanics and Foundation Div. ASCE,
Vol 94, pp. 637-659.
35 Gray, H. (1936). Stress distribution in elastic solids. Proc. 1 mt. Conf.
Soil Mech. and Found. Eng., Vol. 2, p. 157.
36 Griffiths, D.V. (1985). The effect of pore fluid compressibility on failure
loads in elasto plastic soils. mt. J. Numerical and Analytical Methods in
Geomechanics, Vol 9, pp. 253-259.
37 Gunn, M.J. and Britto, A.M. (1982). CRISP: users' and programmers'
guide. Engineering Department, Cambridge University.
38 Gunn, M.J. and Britto, A.M. (1984). CRISP84-users and programmers
guide. Engineering Department, Cambridge University.
228
References
39 Hight, D.W., Jardine, R.J. and Gens, A.(1987). The behaviour of soft
clays. Embankrnents on Soft Clays, Chapter 2, Bulletin of the Public Works
Research Centre, Athens, Greece.
40 Hansbo, S. (1981). Consolidation of fine-grained soils by prefabricated
drains. Proc. 10th ICSMFE, Stockholm, Vol. 3, pp. 677-682.
41 Hansbo, S., Jamiolkowski, M. and Kok, L. (1981). Consolidation by
vertical drains. Geotechnique, Vol 31, No. 1, pp. 45-66.
42 Hardin, B.O. and Black, W.L (1966). Sand stiffness under various
triaxial stresses. Journal of Soil Mechanics and Foundation Engineering,
ASCE, Vol. 92, SM2.
43 Hinton, B. and Campbell, J.S. (1974). Local global smooting of
discontinuous finite element functions using a least squares method.
International Journal of Numerical Methods in Engineering, Vol. 8, pp.
461-480.
44 Hird, C.0 and Kwok, C.M. (1986). Predictions for the Stanstead
Abbotts trial Embankment. Prediction Symposium on a Reinforced
Embankment, King's College, London.
45 Hird, C.0 and Kwok, C.M. (1 990a). Finite element studies of interface
behaviour in reinforced embankments on soft ground. Computer and
Geotechnics, Vol. 8, pp. 111-131.
46 Hird, C.0 and Kwok, C.M. (199Db). Parametric studies of the
behaviour of a reinforced embankment. Proceedings of the 4th International
Conference on Geotextiles, Geomembranes and Related Products, The
Hauge, Netherlands.
47 Hird, C.0 and Russell, D. (1990). A benchmark for soil-structure
interface elements. Computer and Geotechnics, Vol. 10, pp. 139-147.
48 Hird, C.C, Pyrah, I.C. and Russell, D. (1990). Finite element analysis
of the collapse of reinforced embankments on soft ground. Vol. 40, No. 4,
pp. 633-640.
49 Hird, C.C. and Pyrah; I.C. (1990). Predictions of the behaviour of a
reinforced embankment on soft ground, Performance cf reinforced soil
structures, British Geotechnica I Society.
50 Hird, C.0 and Jewell, R.A. (1990). Theory of reinforced embankments
Reinforced Embankments: Theory and Practice in the British Isles. pp.
115-1 39. Pub: Thomas Telford.
229
References
230
References
References
232
References
88 Poulos, H.G. and Davis, E.H. (1974). Elastic solutions for soil and
rocks. Wiley, New York.
89 Roscoe, K.H. and Burland, J.B. (1968). On the generalised
stress-strain behaviour of "wet" clay. In: Engineering Plasticity, Ed. J.
Heyman and F.A. Leckie, Cambridge University Press, pp. 535-609.
90 Rowe, P.W. (1972). The relevance of soil fabric to site investigation
practice. Gotechnique, Vol. 22, No. 2, PP. 1995-300.
91 Rowe, R.K. (1982). The analysis of an embankment constructed on a
geotextlle. Proc. 2 Int. conf. on Geotextiles, Las Vegas, Vol. 2, pp.
677-682.
92 Rowe, R.K. (1984). Reinforced embankments: Analysis and design.
Journal of Geotecnical Engineering Division, ASCE. Vol. 110, GT2, pp.
231-246.
93 Rowe, R.K., Maclean, M.P. and Soderman, 1(1. (1984). Analysis of a
geotextile reinforced embankment constructed on peat. Canadian
Geotecnical Journal. Vol. 21, pp. 563-576.
94 Rowe, R.K. and Soderman, K.L. (1985). An approximate method for
estimating the stability of geotextile-rein forced embankments. Canadian
Geotechnical Journal, Vol. 22, No. 3, pp. 392-398.
95 Rowe, R.K. and Soderman, K.L (1987). Stabilization of very soft soils
using high strength geosynthetics: the role of finite element analyses
Geotextiles and Geomembranes, Vol. 6, pp. 53-80.
96 Rowe, R.K. and Mylleville, B.LJ. (1988). The analysis of
steel-reinforced embankments on soft clay foundations. Proc. 6th tnt. conf.
on Numerical Methods in Geomechanics, lnnsbruck, pp. 1273-1278.
97 Rowe, R.K. and Mylleville, B.LJ. (1989). Consideration of strain in the
design of reinforced embankments. Geosynthetics '89 Conference, San
Diego, USA, pp. 124-1 35.
98 Rowe, R.K. and Mylleville, B.LJ. (1990). Implications of adopting an
4th
allowable geosynthetic strain in estimating stability. Proceedings of the
International Conference on Geotextiles, Geomembranes and Related
Products, The Hauge, Netherlands, pp. 131-136.
99 Russell, D. (1988). A comparison of CRISP84 and CR!SP81. University
of Sheffield, Internal Report.
233
References
References
113 Tavenas, F., Jean, P., Leblond, P. and Leroueil, S. (1983). The
permeability of natural soft clays, part II: Permeability characteristics.
Canadian Geotechnical Journal, Vol. 20, pp. 645-659.
114 Tavenas, F. and Leroueil, S. (1980). The behaviour of embankments
on clay foundations. Canadian Geotechnical J., Vol. 17, No. 2, pp. 236-260.
115 Taylor, D.W. (1937). Stability of earth slopes. Journal of the Boston
Society of Civil Engineers, Vol. 24, pp. 197.
116 Taylor, D.W. (1948). Fundamentals of soil mechanics. Pub:
McGraw-Hill, New York.
117 Terzaghi, K. and Peck, R.B. (1967). Soil mechanics in engineering
practice. Pub: Wiley, London.
118 Tomlinson, M.J. (1987). Foundation design and construction. Fifth
Edition, Pub: Longman.
119 Van Langen, H. and Vermeer, P.A. (1991). Interface elements for
singular plasticity points. Int. J. Num. Meth in Geomech, Vol. 15, pp.
301-315.
120 Watson, H., Crooks, J.H.A., Williams, R.S. and Yam, C.C. (1984).
Performance of preloaded and stage loaded structures on soft soils in
Trinidad. Gotechnique, No. 34, pp. 239-259.
121 Wroth, C.P. and Simpson, B. (1972). An induced failure at a triai
embankment. Part II - finite element computations. Proc. Spec. Conf. on
Performance of Earth and Earth-Supported Structures, ASCE, Lafayette,
Vol. 1, pp. 65-79.
122 Wroth, C.P. (1975). In-situ measurement of initial stresses and
deformation characteristics. Proc. Spec. Conf. in In-situ Measurement of
Soil Properties. ASCE, Rayleigh, North Caroline, pp. 181-230.
123 Wroth, C.P. (1977). The predicted performance of soft clay under a
trial embankment loading based on the Cam-clay model. Finite Elements in
Geomechanics, Willey, Chapter 6. (ed. Gudehus).
124 Wroth, C.P. (1984). The interpretation of in-situ soil tests
Geotechnique, Vol. 34, No. 4, pp. 449-489.
125 Yoshikuni, H. and Nakanodo, H. (1974). Consolidation of soils by
vertical drain wells with finite permeability.Soils and Foundations. Vol. 14,
No. 2, pp. 35-46.
235
References
236