Download as pdf or txt
Download as pdf or txt
You are on page 1of 424

INFLUENCE OF STRESS PATH AND ANISOTROPY

ON THE BEHAVIOUR OF A SOFT ALLUVIAL CLAY

A thesis submitted to the

University of London

(Imperial College of Science and Technology)

for the degree of

Doctor of Philosophy in the Faculty of Engineering

by

Laurence Daniel Wesley

September 1975
ABSTRACT

A laboratory study is made of the strength and

deformation characteristics of a soft alluvial clay. High


quality undisturbed block samples of the clay were obtained

from a test pit put down to a depth of nearly 5 metres at a

site on the Essex coast of the Thames estuary. The behaviour

of undisturbed samples during undrained shear is investigated

in conventional undrained tests and in tests after re-

consolidation to the in situ stresses. The influence of both

"Ko" and isotropic consolidation is studied, and extension


and plane strain tests are carried out. The undrained strength

is found to be not greatly influenced by reconsolidation to

the field stresses or by the wide range of pore water tension

values apparent in the undrained tests not involving reconsolida-


tion. This is shown to be due to the tendency of the stress

paths to converge toward a common failure point.

The undrained strength anisotropy of the soil is

investigated by carrying out triaxial compression and extension

tests on samples trimmed with varying inclinations between

vertical and horizontal. The strength variation in such tests


is shown to be related to the difference in stiffness of the

soil skeleton in the horizontal and vertical directions.

The pore water tension in the samples is investigated

in some detail in order to account for the wide range of

values obtained from the triaxial test specimens.

A number of drained triaxial tests are carried ou

starting from both an isotropic stress state and a Ko state,


Those carried out from the Ko state followed various stress
paths involving both compression and e:(Lension loading. 7\n
ii

assessment is made of the usefulness of elastic theory

(isotropic or anisotropic) in interpreting behaviour in those

drained tests and in relating it to undrained behaviour.


The relevance of the laboratory tests to field

situations is discussed and a comparison made of laboratory

values with those from field vane tests and those indicated

by the behaviour of a trial embankment.


iIi

ACKNOWLEDGEMENTS

The work described in this thesis was carried out as

part of an SRC supported research project concerned with the

influence of stress path on the strength and deformation

characteristics of soils. The writer is particularly grateful

to Professor A.W. Bishop for the opportunity to be involved

in the project and for supervision and guidance throughout the

course of the work.


A large number of people have assisted the writer

during the three years in which the work was carried out.

Thanks are specially due to the following:


Dr. Vaughan, Dr. Skinner and Dr. Chandler for many

useful discussions and valuable guidance on various aspects

of the experimental work.


Messrs D. and F. Evans, E. Harris, L. Spall, C. Gagg

and Mrs. E. Gibbs for much valuable assistance in the

laboratory, particularly in the design and construction of


the new hydraulic triaxial equipment.

Miss J. Gurr for the photographs in this thesis.

Mrs. B. Price for typing the manuscript.


Mr. R.S. Pugh with whom the writer worked closely in

coordinating the field and laboratory work involved at the

Mucking site.
Messrs D. Gudgeon and A. Reid of Einnie and Partners
who were responsible for the test pit and the many excellent

samples supplied to Imperial College.

Finally, thanks are due to my wife and children whose


wish to see the sights of London provided the original
iv

motivation for returning to Imperial College in pursuit of

this PhD. (The loan of the carving knife for use in part of

the sample trimming operation was also appreciated).


V

TABLE OF CONTENTS

Page
Abstract
Acknowledgements

Table of Contents .v
Chapter 1 Introduction 1

Chapter 2 Literature Review 4


2.1 Influence of Stress Path on Undrained
Strength 4
2.1.1 Significance of "Stress Path" 4
2.1.2 Field Stress Paths 6
2.1.3 Stress Paths in Undrained Laboratory
Tests 9
2.2 Anisotropy in Normally Consolidated
Clay 20
2.2.1 Undrained Compression Tests at
Varying Inclinations 20
2.2.2 Field Vane Tests 24
2.2.3 Bjerrum's Hypothesis and the Use of
CompreSsion and Extension Tests 26
2.2.4 Anisotropy in Terms of Effective Stress 29
2.2.5 Anisotropy of the Undrained Deformation
Modulus 31
2.3 Undrained Plane Strain Behaviour and

its relationship to Triaxial. Behaviour 40


2.3.1 Theoretical Considerations 40
2.3.2 Tests on Normally Consolidated
Remoulded Clays 42
2,3.3 Tests on Undisturbed Samples 42
vi
Page

Chapter 3 Equipment 55

3.1 The Stress Controlled (or Hydraulic)


Triaxial Apparatus 55.

3.1.1 Design of the Apparatus 55

3.1.2 Principle of Operation 60

3.1.3 Method of Application of the Controlled


Stresses 64

3.1.4 Some Practical Details 66

3.1.5 Performance of Load Cells and


Transducers 69

3.2 The Plane Strain Apparatus 87


3.2.1 Modifications 87

3.2.2 Practical details and Corrections 89

Chapter 4 Tests on London Clay 94

4.1 Scope and Purpose 94

4.2 Samples Inclined at 450 96

4.3 Consolidated Undrained Tests on Vertical


Samples with Constant Vertical Stress

and Decreasing Lateral Stress 98

4.4 Drained Tests on Vertical Samples with


Constant Vertical Stress and Decreasing

Lateral Stress 101

Chapter 5 Tests on Soft Normally Consolidated



Clay from Mucking Site 121

5.1 Geology of Site and General Comments



on the Test Programme 121

5.1.1 Geology of Site 121

5.1.2 Comments on Test Programme 124

5.2 General Description of Samples and



Soil Profile at the Test. Pit 127
vii
Page

5.2.1 Method of Sampling, Sample Identifica-


tion Details and Handling Procedure 127
5.2.2 Description of the Soil 129
5.2.3 Water Content and Atterberg Limits 131
5.2.4 Clay Fraction and Activity 133
5.2.5 Degree of Saturation 135
5.3 Undrained Measurement of Shear .
Strength 144
5.3.1 Test Details and Procedure 144
5.3.2 Test Results 146
5.3.3 Additional Tests to Investigate the
Undrained Strength Anisotropy 150
5.4 Consolidation Characteristics 192
5.4.1 Oedometer and Dissipation Tests on
Cylinder Sample T24 192
5.4.2 Consolidation Stage of Triaxial Tests 193
5.4.3 Test on Borehole Tube Samples 194
5.5 Consolidated Undrained Triaxial
Test Series 203
5.5.1 Procedure 203
5.5.2 Initial Pore Water Tension (Suction) 205
5.5.3 Results and Discussion 206
5.5.4 Results from Borehole Samples 209
5.6 Influence of Stress Path on Behaviour
in Consolidated Undrained Tests 225
5.6.1 Test Programme and Purpose of Tests 225
5.6.2 Test Procedure 226
5.6.3 Test Results 231
5.7 Membrane and Strain Rate Influence 264
Vii i
Page

5.7.1 The Membrane Correction 264


5.7.2 Strain Rate Influence 267
5.8 Further Investigation of Anisotropy 282
5.8.1 Test Programme and Purpose of Tests 282

5.8.2 Test Procedure 283

5.8.3 Test Results 285

5.9 The Pore Water Tension in the Samples 311

5.9.1 Values Measured in Triaxial Samples 311

5.9.2 Pore Water Tension Measured Directly


in Sealed Blocks 312

5.9.3 Influence of Moisture Evaporation


during Sample Trimming 314

5.10 Drained Tests 320

5.10.1 Estimation of Ko 320

5.10.2 Measurement of Soil. Skeleton Stiff-


ness in the Vertical and Horizontal

Direction 322

5.10.3 Triaxial Tests following Various

Stress Paths after "K" Consolidation 327

5.10.4 Relationship between Dzained and Un-


drained Behaviour 330

5.10.5 The preconsolidation 1,L -2ssure pc 336

5.11 The Undrained resid Strength 364


5.11.1 Tests on Undi.,;tui:bed ,)les 364

5.11.2 Tests on Undi,3turbed

Pre-cut Plane 367

5.11.3 Tests on Rew.ouLded s 363

5.11.4 Summary 371

5.12 Long Term Coll(1-1'. Test!-1 379


ix
Page

Chapter 6 Discussion and Conclusions 384


6.1 The Undrained Strength 384
6.2 Use of Undrained Strength Values from
Laboratory Tests for Analysis and
Design 391
6.2.1 Relationship between Principal Stresses
and Strength on Failure Planes 391
6.2.2 The "Normalised Behaviour" Approach
of Ladd and Foott (1974) 394
6.3 Laboratory Undrained Strength in
Relation to Field Vane Strength a. d
Strength Indicated by the Trial
Embankment 399
6.3.1 The Field Vane Strength 399
6.3.2 Field Behaviour 400

References 407
1

CHAPTER 1

INTRODUCTION

This thesis is concerned principally with a


laboratory investigation of the strength and deformation
characteristics of a soft undisturbed clay. The clay con-
cerned is a normally consolidated clay found along the north
(Essex) coast of the Thames estuary. The samples tested were
taken as part of an investigation carried out in conjunction
with the construction of a trial embankment. This embankment
was used as a full scale test of the in situ behaviour of
the soil and thus provides a direct check on the relevance
and validity of the laboratory tests. The trial embankment
and the associated drilling and testing programme were
intended to provide detailed information required for the
design of improved flood defences along the.Thames coast.
The construction and performance of the embankment are
described in detail by Pugh (1976).
The laboratory work has been directed mainly to
investigating the strength and deformation behaviour during
undrained loading. It cannot easily be divided up into
distinct sections but can be broadly considered to have been
concerned with the following aspects:
1. The influence which stress path has on the soil
behaviour.
2. The nature of the anisotropy of the soil.
3. The relationship between triaxial and plane
strain behaviour.
2

The question of the pore water tension (suction)in


the soil after sampling, and the way in which it is affected
by trimming and setting up samples in triaxial cells is
considered in some detail. Other effects such as sample
size and strain rate are investigated and an attempt is
made to measure the undrained residual strength of the soil.
In addition to the tests on the soft normally
consolidated clay, several tests were carried out on samples
of London clay. While these tests do not form part of the
main topic of this thesis, they are related to it and show
several points of interest, particularly in relation to a
series of tests performed earlier at Imperial College on the
same samples of London clay (Atkinson 1973).
The layout of the work is as follows.
In Chapter 2 a review is given of literature relevant
to the present investigation. Chapter 3 describes the
apparatus used for the laboratory tests, in particular a
new hydraulic triaxial apparatus specially designed and '
built for controlled stress testing. The performance of this
apparatus is discussed in detail and some modifications
made to the existing plane strain apparatus are described.
The tests on London clay are described in Chapter 4. The
bulk of the experimental work is contained in Chapter 5
which describes all the tests on the Thames estuary samples.
Chapter 6 concludes the thesis with a resume and short
discussion of the relevance of the laboratory measurements
to field behaviour.

A word of explanation is perhaps necessary with


regard to the system of units used throughout this thesis.
3

In general S.I. units are used except for Chapter 4 which


presents the results of the London clay tests. In this
chapter British units are used so that the results can be
compared directly with those of Atkinson.
CHAPTER 2

LITERATURE REVIEW

In this chapter an outline is given of published


work dealing with the subject of stress paths and anisotropy,
particularly in relation to the undrained strength of soft
clays. The review is not intended to be comprehensive but
covers the more important published papers and provides a
summary of the present state of knowledge in this field.
The review is in three parts, the first dealing with stress
path effects, the second with anisotropy and the third with
the relationship between triaxial and plane strain behaviour.

2.1 Influence of Stress Path on Undrained Strength

2.1.1 Significance of "Stress Path"


The term stress path is used to denote the sequence
and sign of the stress changes experienced by a soil element
in the field in some particular situation, or the stress
changes a sample undergoes during a laboratory test. The
importance of stress path has long been recognised; the most
frequently quoted example of its influence being the differing
deformations associated with the active and passive pressure
states behind retaining walls.
While it is true that the values of c' and 41 are not
greatly influenced by the stress path leading to failure, the
pore pressure (and thus undrained strength) during undrained
loading, and the deformations (including volume change)
5

during drained loading are dependent primarily on the


sequence in which the stresses are applied. This has been
emphasised by Bishop and Henkel (1957, page 20) who illustrate
the effects by means of undrained and drained triaxial tests.
The term "stress path" has perhaps been made popular in
recent years by Lambe (1963 and 1967) who has advocated
estimating deformations, particularly settlements, by the
"stress path method" involving subjecting soil samples in
the laboratory to the same sequence of stress changes as
they undergo in the field. While Lambe's method appears
sound in theory, it does not appear to have been used widely
in practice. This is partly because of the more complex
testing involved, but also because of the difficulty in
"linking up" the deformations of "typical soil elements"
to obtain the settlement of the structure at the surface,
which is generally the main concern of the designer.
The advent of the computer and finite element
techniques has provided a means of estimating deformations
in situations with almost any geometrical configuration and
has thus been an added incentive to the investigation of
soil behaviour under varying stress paths, and to the study
of soil deformation behaviour generally.
The use of the term stress path in the title of this
thesis is perhaps a little misleading. The term "stress
system" might have been more appropriate as a considerable
amount of work has been done investigating the relationship
between triaxial and plane strain behaviour, and between
triaxial compression and extension tests. However, the term
stress path can be taken in a general sense to cover this
6

sort of effects.

2.1.2 Field Stress Paths


To examine the stress paths which soil elements are
subjected to in particular field situations it is necessary
first to know the initial state of stress on the element.
With normally consolidated soils it has long been assumed
that the horizontal stress is given by the "virgin" Ko
value which can be calculated with reasonable accuracy from
the expression Ko = 1 - sin V (Bishop, 1958; Simons, 1958).
This leads to values of Ko in the region of 0.45 to 0.65.
This assumption appears to have been validated in recent
years by actual measurements of the horizontal stress in the
field (Bjerrum and Anderson, 1972; Wroth and Hughes, 1973;
Massarsch et al. 1975), all of which give values in the range
of 0.4 to 0.6 except close to the surface where higher
values were recorded in some cases.
The stress paths followed by soil elements during a
particular loading or unloading operation can generally
only be determined accurately by means of some numerical
method such as finite element analysis. The paths are
complicated by the fact that the principal stress directions
will in general be continually changing during the loading
sequence. The only elements in which the principal stress
directions will not rotate are those at the centre of
symmetrical loads or excavations. There is surprisingly
little data published on actual stress paths determined by
finite element methods as the main concern appears to have
been with ultimate deformations and stresses.
7

However a general idea of the range of stress paths


possible in the field can be obtained by considering specific
elements in typical situations as for example Lambe (1967)
has done. Examples are given in Fig. 2.1.1. This shows
very approximately the total stress paths which elements in
a normally consolidated soil will be subjected to in an
excavation or during the placement of fill. The stress paths
are shown on the two types of plot commonly used for
illustrating stress paths. For simplicity the initial pore
water pressure has been ignored. Thus the starting point 0
lies on the Ko line and the paths shown are total stress
paths starting from this point. The first two elements both
undergo a reduction in stress; with element 1 the reduction
is primarily in the vertical direction and with element 2
primarily in the horizontal direction. The second two
elements undergo an increase in stress; with element 3 this
increase is primarily in the horizontal direction and with
element 4 is primarily in the vertical direction. With
elements 2 and 4 there is no reversal in the direction of
shear but with elements 1 and 3 the direction of shear is
reversed as the two stresses become equal and then aH
becomes greater than a:v. Elements 2 and 4 can be considered
to fail in "compression" (av > aH) but elements 1 and 3 fail
in "extension" (aV < aH). The terms compression and
extension refer to the mode of deformation in relation to the
vertical axis, and will be used in this way throughout this
thesis.
The following factors should be emphasised in relation
to these stress paths:
8

(a)The paths are oversimplified and in practice the


paths may be curved and will involve some rotation of the
principal stress axis.
(b)The influence of the intermediate principal
stress is not considered. The examples given are intended
to illustrate the plane strain condition although similar
paths would apply in the case of excavations or-fills not
meeting the plane strain condition.
(c)The paths are total stress paths, which is why
they do not terminate at the effective stress failure lines.
With clays, the loading (or unloading) can often be con-
sidered undrained as there is not normally time for significant
flow of pore water to occur during the formation of the
cutting or construction of the embankment. The effective
stress path may be radically different from the total stress
path and for the two simple cases illustrated only two effective
stress paths are in fact involved. It follows from the
principal of effective stress that during undrained loading
it is only the sign and magnitude of the component av - aH
which determines the effective stress path and thus the
deformation and failure of the element. Thus total stress
paths OA and OD will result in the same effective stress
path and OB and OC will result in the same effective stress
path. This has been adequately demonstrated in laboratory
tests (eg. Parry, 1960; Bishop and Wesley, 1975).
There may of course be situations in which some
drainage occurs during construction in which case the stress
paths may lie somewhere between the drained and undrained
paths, or may consist of a series of undrained stages with
"pauses" during which drainage occurs. In the building of an
9

embankment for example rates of construction may be adjusted


to ensure a certain.amount of pore pressure dissipation as
construction proceeds.

2.1.3 Stress Paths in Undrained Laboratory Tests


The unconfined compression test and undrained
triaxial compression test have long been standard methods
for determining the undrained strength of soils in the
laboratory. However, there has been considerable controversy
as to the validity of these tests arising initially from
discrepancies between the undrained strength values from
undrained tests and those from field vane measurements.
Skempton (1953) for example discusses this point but presents
results which appear to support his view that undrained
laboratory tests give a valid measure of undrained strength
except for certain soils. In more recent years the argument
against conventional undrained laboratory tests has been
strengthened by investigations into the stress path (in
particular the starting point of the path) which the soil
follows in the laboratory compared with that which it would
follow in the field. To follow these arguments it is
necessary to consider the stress changes the soil undergoes
during sampling and extrusion prior to setting up the sample
in an apparatus.
If the sampling is carried out in an ideal (or
"perfect") manner then all that is involved is a release
of the total stresses acting on the sample. No disturbance
of the soil structure occurs and there is no gain or loss
of pore water. The release of the total stresses will
10

result in a negative pore water pressure (tension or


"suction") being set up in the sample. Assuming the soil
to be fully saturated this pore water tension is given by

Us = -Ko a'
V V- A(1 - Ko )0'

where A = the conventional parameter applicable in this


case to the release of the deviator stress
aV = the vertical effective stress in situ

U,
Hence - = Ko + A(1 Ko)
(5V

If the soil behaved elastically A would have a value of


1 and thus
-s
US 1 + 2Ko
=
(77 3
V

The effective stress in the sample (= -US) would then have


the value of the average in situ effective stress. However,
the soil does not behave elastically as has been shown by
laboratory tests. Bishop and Henkel (1953) consolidated an
undisturbed sample under conditions of no lateral yield to a
vertical stress some four times greater.than that in the
field. The deviator stress was then released simulating a
sampling operation and it was found that a small increase in
pore pressure occurred. The A value in this case was about
-0.1. Skempton and Sowa (1963) also investigated the same
effect with remoulded samples and found that the release of
the deviator stress also induced a small positive pore
11

pressure. Ladd and Lambe (1963) carried out similar tests


on two clays and found that the release of the deviator
stress induced a small negative pore pressure, the average
A values being 0.17 and 0.11 for the two clays. Adams and
Radhaltrishna (1970) obtained A values close to zero for
similar tests on a glacial lake deposit. The negative pore
water pressure and thus the effective stress in the perfect
sample is likely to be rather less than the average in situ
effective stress. The tests mentioned above suggest that
U
the ratio --67 7- will have a limiting range between about 0.35
and 0.80 but will normally be between 0.5 and 0.6. In other
words the effective stress in a "perfect" sample will be close
to or slightly above the in situ horizontal effective stress
since the Ko value is normally in the 0.4 to 0.6 range.
Actual measurements of pore water tension in un-
disturbed samples of normally consolidated clays have been
recorded on a number of occasions and the values have generally
been found to be much lower than the values suggested above
for perfect samples. The measurements have all been on
samples taken from boreholes with good quality thin walled
tubes. The tension has been determined by setting samples
up in a triaxial cell and measuring the pore water pressure
after the application of a known cell pressure. Assuming
the soil to be fully saturated the difference gives a direct
measure of the tension in the pore water at zero confining
pressure. Ladd and Lambe (1963) report that the measured
pore water tension in samples from three different clays
ranged between 1% and 43% of the value applicable after a
"perfect" sampling operation. Adams and Radhakrishna (1970)
12

found tension values between about 25% and 65% of the theo-
retical value, the lower values coming from the deeper
samples. Bjerrum (1973) indicates that with Norwegian clays
a similar loss of tension occurs during conventional sampling.
Parry and Nadarajah (1974) report no negative pore pressure
in samples of a clayey silt from relatively shallow depths.
Brand et al. (1972) found with samples from depths up to 7 m
in a soft clay in Bangkok that the pore water tension values
were very variable but did not show any increase with depth.
At shallow depths the values were slightly higher than the
value applicable to a perfect sample but deeper down were
considerably lower.
The reason for the loss of pore water tension in soft
clays is not entirely clear, although it is not surprising
in view of the difficulties involved in taking undisturbed
tube samples from boreholes in these materials, particularly
at greater depths. Clearly sample disturbance (due to
physical distortion) and the opportunities the soil has to
take up water during the sampling operation are the main
factors involved but the relative importance of each is
unknown. Bjerrum (1973) believes that the inner less
disturbed part of the sample will tend to suck in water
available from the outer disturbed zone of the sample.
Some measurements made by Schjetne (1971) during the actual
sampling operation appear to support this view as some tension
still existed in the soil after sampling but this disappeared
completely in a few hours.

It is clear from the above that in conventional


undrained tests the effective stresses at the start of the
13

test are likely to be very different from those which acted


on the sample in the ground. The stresses are isotropic
and generally substantially lower than those in the field.
This last fact is perhaps the most serious criticism which
can be made of the undrained test (unconfined or triaxial).
Two further criticisms in relation to stress path can be
made. The first is that the stress path during a test is
always the same that is am constant and av increasing, and
the second is that the test is axisymmetric whereas field
situations are more often plane strain. Despite these
"deficiencies" the test still appears to give reasonable
results in many cases and part of the reason for this is
apparent from a closer examination of laboratory and field
stress paths.
From the test results of Ladd and Lambe (1963),
Skempton and Sowa (1963) and Adams and Radhakrishna (1970),
it is possible to draw a general picture of the effective
stress paths the clay undergoes in the laboratory compared
with those- it is likely to undergo in the field. This is
done in Fig. 2.1.2. The point 0 represents the in situ
stress state. During undrained loading in the field the
stress path followed will be OA or OC, depending upon whether
av - am is increasing or decreasing (and becoming negative).
During a perfect sampling operation the stress path would
be OB. For simplicity OB is taken as the same path as OC
although this is not necessarily so as OC in the field could
be a plane strain path. In a laboratory undrained compression

test on this perfect sample the stress path would look


something like BD, the undrained strength being almost the
14

same as that for the direct path OA. In actual samples


there is a loss of pore water tension so that at the start
of the laboratory test the sample may be at some point such
as E. The stress path during the test will then look something
like EF with a consequent decrease in the undrained strength.
As indicated the undrained strength loss results primarily
from the loss of effective stress (ie loss of pore water
tension) and not from the release and reapplication of the
deviator stress.
To further clarify the way a soft clay behaves during
undrained compression tests in the laboratory Fig. 2.1.3
has been prepared. The behaviour is slightly idealised but
illustrates the results of consolidated undrained triaxial
tests with different initial consolidation pressures. This
pattern of behaviour is taken from the same references as
Fig. 2.1.3 but is verified by other sources (eg Parry 1968).
The upper part of the figure shows a plot of undrained strength
against initial consolidation pressure and the lower part
the effective stress paths during the tests. The following
points should be noted:
(a) At higher stresses the soil behaves as a normally
consolidated soil but at lower stresses behaves
as an overconsolidated soil.. This "overconsolidated"
behaviour is not surprising since any deposit
normally consolidated geologically tends to
behave as a lightly overconsolidated soil due,
to creep effects and the development with time of
bonds between the particles (Bjerrum, 1967).
In this case there is also a true overconsolidation
15

effect due to the stress reduction from point


B to E.
(b)The effective stress failure envelope is only a
straight line through the origin at the higher
stress level. At low stresses the envelope is
significantly above this straight line.
(c)The undrained strength versus initial consolida-
tion pressure is similarly not a straight line
through the origin. It flattens out at low
consolidation pressures.

The above factors mean that the strength loss which


might be expected to accompany the loss of pore water tension
is thus offset or counterbal-.anced to some extent by the -\-
change in the shape of the stress path, ie the change in
behaviour from normally to overconsolidated. Stress paths
EF and BD tend to converge as the tests proceed. It is
apparent also that the loss of strength due to loss of pore
water tension is likely to be much more serious with deep
samples than shallow ones. With shallow samples the smaller
relative tension loss and the tendency of the stress paths to
converge may mean that the undrained strength is not greatly
influenced. With deeper samples which suffer a proportionately
greater tension loss the effect is likely to be more serious.
This fact is emphasised by Adams and Radhakrishna (1970)
who state that the discrepancy between field vane and conventio-
nal laboratory strength increases with depth (and is attributed
to loss of tension) whereas there is little or no descrepancy
in the "overconsolidated" soil at shallow depth.
16

These factors all need to be kept in mind when con-


sidering the best method to use for the laboratory measure-
ment of undrained strength. Ladd and Lambe (1963) believe
that reconsolidating the sample to the in situ stresses
before undrained compression testing will lead to an over-
estimate of undrained strength due to the reduction in void
ratio during the consolidation stage. They propose a rather
tedious method for eliminating this effect. However there
is no evidence to show that the increase in undrained strength
which may result from the void ratio reduction is not more
than counterbalanced by the loss of strength due to damage
to the soil structure during sampling and reconsolidation.
Bjerrum (1973) emphasises strongly the need to reconsolidate
samples to their in situ stresses and gives results from
Norwegian clays for undrained and consolidated undrained tests
showing much lower strengths with the undrained tests.
However the depth of the samples is not given and Bjerrum
does not consider the relative importance of the strength
"loss" at different depths.
It is of interest to note that Taylor (1948, page 397)
discusses the merits of true undrained tests, and undrained
tests after reconsolidation to the in situ pressure. He
states "it is likely that the reconsolidated test
approximates the in situ strength more closely than any
other type of test."
17

0 Cr;

.0" \
.1••■ •••••• •■■•■ •••■■• /11•111M

Excavation LI 441 °Y

Kf
A
te
Sta
D Kos
N

0
0 0y+ OH
C 2

K
K = Effective stress f
f failure line

TOTAL STRESS PATHS FOR SOIL ELEMENTS IN


PARTICULAR FIELD SITUATIONS

FIG 2.1.1
18

from `perfect sampling.

from tension loss.

Strength D
loss
eko
4, 0
F In situ
stresses

Suction loss

g-1
2

elope

0-"\T = vertical stress


0 = horizontal stress

OA
possible field paths
0 BC

0 BD laboratory path with perfect sampling


E F probable laboratory path after loss of tension

EFFECTIVE STRESS PATHS DURING UNDRAINED LOADING


OF NORMALLY CONSOLIDATED CLAY

FIG 2.1.2
19

Initial Effective Consolidation Pressure ( )

Nor malty
Consol idated
/ behaviour

C■1

0-I CT '
1 + 3
2

BEHAVIOUR OF NORMALLY CONSOLIDATED CLAY IN


CONSOLIDATED UNDRA1NED TESTS

F I G 2.1.3
20

2.1 Anisotropy in Normally Consolidated Clays

A glance at the literature suggests that some wide


differences of opinion exist as to the relative importance
of anisotropy in normally consolidated soils, at least as
far as undrained strength is concerned. Parry (1971, page 651)
for example, after discussing the evidence from various
investigations, states that "it seems likely that anisotropy
can be disregarded as a likely cause of differences between
predicted and observed stability behaviour." Bjerrum (1972)
on the other hand, based on the results of compression and
extension tests states that "it is immediately apparent how
great the anisotropy is - - - the ratio of undrained strengths
observed in compression and extension tests being of the
order of 2 to 3." If this is a legitimate measure of aniso-
tropy then it could account for large descrepancies between
predicted and observed behaviour, at least in those cases
where predicted behaviour is based on compression tests.
These differences of opinion arise in part from the different
methods which have been used to measure undrained strength
anisotropy, and from the fact that often only one method has
been used on a particular clay so that it is difficult to
inter-relate the evidence from the various methods. These
methods, and the evidence they produce will be considered
in turn in the following sections.

2.2.1 Undrained Compression Tests at Varying Inclinations


Undrained compression tests (unconfined or triaxial)
on cylindrical samples cut with their axes at varying inclina-
tions to the vertical have been carried out on a number of
21

occasions. A comprehensive set of tests was done by Lo


(1965) on samplesof lightly overconsolidated Welland clay.
The results of the tests are shown on a polar diagram in
Fig. 2.2.1. It is seen that the strength is greatest when
the axis is vertical and decreases steadily with inclination
to a minimum value when the axis is horizontal. As-Bishop
(1965) has pointed out there is considerable ambiguity
associated with this type of test when attempts are made to
relate particular values of undrained strength with particular
planes. Failure may occur on planes originally vertical in
the ground, and not on planes which could form part of a
conventional slip surface (see section 5.3 and Fig. 5.3.19).
Even if this ambiguity is ignored, there is still doubt as
to the orientation of the failure plane in inclined samples,
as indicated in the lower part of Fig. 2.2.1. The planes AB
and CD are possible failure planes, and except in the case
of vertical or horizontal samples there is no a priore
reason to assume the strength associated with the two planes
is the same.
To overcome the ambiguity with respect to the
failure plane De Lory and Lai (1971) carried out tests on
block specimens (also of Welland clay) 3 in. high, by
1.5 in. wide by 4.5 in. long. They called their tests
semi-confined as they used an arrangement whereby failure
was restricted to one direction. The arrangement is shown
in Fig. 2.2.2,. Samples were tested at varying values of i
from 0 to 180°. The results of these semi-confined tests
are shown in the lower part of Fig. 2.2.2. They are plotted
first as strength versus the angle i and in this plot
22

show exactly the same trend as Lo's tests, ie maximum strength


when the axis is vertical and minimum strength when the axis
is horizontal.
The results are also plotted against the angle (a)
of inclination of the failure plane to the vertical. The
failure plane was found to always occur at an angle close
o
to 30 to the compression axis of the sample and this value
of 30° has been used in plotting this second diagram. This
angle a is taken simply as the angle between the failure
plane and the vertical without regard to the direction in
which shear is occurring. Thus a varies only between 0 and
90o. Because of the test arrangement two values of strength
are associated with each value of a, except when a = 0 and
0
a = 90 . For example a is 60o in both the test with i = 30
and with i = 90°. However there is a difference in strength
between these two tests and De Lory and Lai attribute the
strength difference to the fact that shear on the plane is
in opposite directions in the two tests. Thus, in the plot
of strength against a there are two curves, the upper curve
(case a) applying when shear occurs in the same direction as
the existing shear stress in the ground, and the other
(case b) when the direction of shear is reversed. De Lory
and Lai believe their results support the hypothesis put
forward by Bjerrum (1972 and 1973) that the soil structure
has a greater resistance to shear in one direction than
the other. This hypothesis is discussed later. However, it
is not at all established that unique values of undrained
strength (dependent only on the direction of shear) can be
associated with particular planes. It is significant that
the strength plot is symnetrical about the i = 90° line.
23

For example the sample with i = 30 and i = 150 have the same
strength despite the fact that shear occurs on different
planes. A simple explanation for this type of behaviour
would be that the values of c' and cp' are not influenced
by orientation but that the pore pressure is, the response
increasing as i increases from 0 to 90°.
It has not been established that the behaviour
revealed in the above tests is the normal pattern for soft
undistrubed clays as insufficient data from other sources
is available. Parry and Nadarajah (1974) for example found
that a horizontal sample had a lower strength than a vertical
sample but that with the sample inclined at 45° the strength
was lower again. However, it appears that Parry and Nadarajah
tested only one sample at each orientation so it is difficult
to know whether the values represent a general trend.
The question of whether the variation in undrained
strength revealed by these tests arises from anisotropy of
the c' and 4' values or from a difference in pore pressuke
response is of considerable importance and appears to be at
least partially answered in Lo's reply to the discussion on
his (1965) paper. In closing the discussion Lo (1966)
presents the results of a series of consolidated undrained
triaxial tests with pore pressure measurements carried out
on the Welland clay used in earlier tests. Samples were
tested at inclinations from i = 0 to i = 90°. The strength
pattern.in these tests was the same as in the earlier un-
drained tests, but it was found that the pore pressure versus
strain curves were identical in each test. However, the
strain to failure was progressively greater as i increased
24

from 0 to 90 so that the actual value of the pore pressure at


failure similarly increased. Loss data is insufficient
to accurately determine c' and (1)1 values but it appears that
the variation in these is relatively small and that the
difference in pore pressure at failure accounts for most of
the variation in undrained strength with inclination.

2.2.2 Field Vane Tests


By carrying out vane tests using vanes with differing
height to diameter ratios it is possible to separate out the
two components of torque, that is the shear resistance around
the cylindrical surface itself, and the shear resistance on
the upper and lower flat surfaces. It is thus possible to
calculate the shear strength on vertical planes and on
horizontal planes. Aas (1965 and 1967) has used this procedure
on six sites in Norway where the clays were soft marine
deposits of high sensitivity. The results obtained and the
method used for plotting them are shown in Fig. 2.2.3.
Plotted in this way, the intercept on the y axis gives the
value of SV and the intercept on the x axis the value of
SH/SV. It is seen that the value of Su/Sy is generally
between 1 and 2, ie the strength on the horizontal plane is
greater than that on the vertical plane. Aas shows from a
further test with a vane having blades at 45° that there is
a steady drop in strength from the horizontal to the vertical
plane. Aas suggests that the difference in strength on
different planes is related to the in situ effective stresses
acting on the planes. In normally consolidated clays the
stress on the horizontal plane is generally about twice that
25

on the vertical plane, whereas in lightly overconsolidated


clays the stresses may be approximately the same. Aas found
that the values of SH/SV closest to unity were from the two
clays which were lightly overconsolidated. There thus appeared
to be a correlation between the strengths revealed by the vane
tests and the state of stress in the ground.
The results of these vane tests do not appear to be
compatable with results of the compression tests discussed
earlier. If in a normally consolidated clay the minimum
strength is found on vertical planes, then it follows that in
tests on cylindrical specimens, the specimen cut with the
axis inclined at 30° to the vertical should have the lowest
strength as failure would occur on a plane originally vertical
in the ground.
Wiesel (1973) carried out vane tests in a soft normally
consolidated soil near Stockholm using vanes of varying
proportions, similar to those used by Aas. The results,
given in the lower part of Fig. 2.2.3 do not, however, show
the same trend as those of Aas. According to Wiesel's tests
the ratio of SH to SV is about 0.6 to 0.8, ie the strength
is lower on the horizontal plane.
It thus appears that vane tests do not produce a consis-
tent trend and appear at variance with the results of laboratory
compression tests. It should be borne in mind when attempting to
evaluate these tests that in the vane test shear on the vertical
plane occurs in the horizontal direction and not in. the vertical
direction as it would (or could) do in a compression test on a
cylindrical sample trimmed at the appropriate inclination.
This is a possible, though unlikely, cause of the lack of
26

agreement between vane and compression tests. A further


complicating factor is the state of drainage. Compression
tests are clearly undrained tests, but with field vane tests
there is the possibility of some drainage occuring during the

test.
Regardless of the importance of the above two factors,
it is perhaps unrealistic to expect the two types of test
(compression and vane) to show agreement as the pore pressures
generated by the different types of loading are likely to
be quite different.

2.2.3 Bjerrum's Hypothesis and the Use of Compression and


Extension Tests
'Bjerrum and Kenney (1967) have put forward a hypothesis
to explain the large difference in strength they found between
undrained triaxial compression and extension tests. The
tests were carried out after consolidation to the in situ
Ko stresses, and showed the strength in extension to be less
than half that in compression. They also found that in
shear box tests there was a large drop in strength when the
shear direction was reversed. When the stresses were applied
in such a way that the shear stress direction was the same
as that which operated in situ, the strength was over 3 times
greater than when the shear stress was in the opposite direc-
tion. To explain the difference Bjerrum and Kenney put
forward their hypothesis that the soil skeleton has adapted
itself to resist shear stresses acting in the direction they
acted in during consolidation, but is much less capable of
resisting stresses in the opposite direction.
27

This hypothesis is further developed in later publica-


tions (Bjerrum, 1973 and 1974, Berre and Bjerrum, 1973)
and Bjerrum uses the ratio of strength in extension to that
in compression as a measure of the anisotropy of the clay.
Fig. 2.2.4 shows the results of tests on clays at six sites,
five in Norway and one in Thailand. The ratio of strength
in extension to that in compression varies from-0.2 to 0.5.
Berre and Bjerrum (1973) believe that the anisotropy increases
as the plasticity decreases. They state that "the available
test results certainly indicate that the values observed in
compression and extension tests are related to the plasticity
of the clay and that the anisotropy expressed by the ratio
increases for decreasing plasticity index. Whether this
conclusion is justified by the data shown in Fig. 2.12.4
appears rather debatable.
If the difference between undrained extension and
compression tests is a legitimate measure of anisotropy
then these tests show much greater anisotropy than the methods
discussed earlier. However, there appear to be good grounds
for believing that this is not in fact a valid means of
measuring the anisotropy of the soil. It has long been known
that with tests on isotropic materials, the strength in
undrained extension is less than that in compression. Bishop
and Eldin (1953) carried out tests on sand samples. Their
results from tests after isotropic consolidation are shown
in the upper part of Fig 2.2.5. It is seen that the ratio
of strength in extension to that in compression varies
between 0.2 and 0.7, depending on the initial porosity.
Parry (1960) reports the results of tests on remoulded London
clay. The ratio in this case is between 0.82 and 0.88,
28

depending on the overconsolidation ratio. There is thus a


significant difference between extension and compression
tests regardless of whether or not the soil is anisotropic.
The reason for this difference lies in the different pore
pressure response in the two types of loading, and the way
in which this influences the effective stresses in the
sample. If elastic behaviour and a Mohr Couloumb failure
criteria is assumed then for a 4)1 value of 30° the strength
in extension will be about 70% of that in compression.
Parry (1971) using the simple critical state assumption that
there is a unique relationship between e (void ratio) and
p (al + Q2 + a3), and a Mohr Couloumb failure criteria,
shows that this leads to strengths in undrained extension
some 70% to 80% of those in compression.
The experimental results in Fig. 2.2.5 suggest that
the difference between extension and compression tests is
greater with non plastic materials than with clays. This
is of some significance in relation to Bjerrums assertion
that the anisotropy increases as the plasticity decreases.
It appears that the difference between extension and compression
tests may increase as the plasticity decreases even when the
material is isotropic.
The use of undrained extension and compression
tests as a measure of anisotropy thus appears a dubious
procedure unless careful account is taken of the pore pressure
response. It is of course a legitimate procedure if we are
interested only in comparing the behaviour of a soil element
beneath the centre of a circular load with that of a soil
element in the centre of a circular excavation. Bjerrum
29

however is concerned with the variation in strength round


a conventional slip.surface which is a plane strain condition.
Bjerrum surprisingly does not evaluate his results
in terms of pore pressure response and the effective stress
failure envelope, although he lists values of c' and cb'
for each soil tested. These are apparently the same in
extension and compression as single values are listed alongside
the two types of test. It would appear then that the
difference in strength in compression and extension tests
results from the differing pore pressure changes and effective
stresses during the test, rather than from differing c' and
cp' values.
It should perhaps be added that the very low ratios
of strength in extension to that in compression reported by
Bjerrum have not been found by other workers. Ladd and
Varallyay (1965) for example carried out extension and
compression tests on two soils after Ko consolidation. The
ratio of strength in extension to that in compression was
0.47 and 0.51.

2.2.4 Anisotropy in Terms of Effective Stress


To provide further information on the nature of the
undrained anisotropy in normally consolidated clays it is
necessary to look more closely at the factors influencing .
the pore pressure response during undrained loading and at
the failure envelope in terms of effective stress (ie the c',
cih' values). The data on these aspects is somewhat limited.
As already mentioned Lo's tests on Welland clay
suggested that the values of c' and cp' were not greatly
30

affected. However, in a later paper Lo and Morin (1972)


report on a study of two sensitive clays belonging to the
Champlain Sea (or "Leda") clays, and show that the effective
failure envelope is markedly anisotropic. Consolidated
undrained and drained tests were carried out on samples cut
with their axis vertical and horizontal, and at some inter-
mediate inclinations. The results of the consolidated
undrained tests are shown in the upper part of Fig. 2.2.6.
The failure envelope revealed by these tests has an unusual
"hump" in the low stress range where structural bonds apparently
provide a large "cohesive" component of strength. It is clear
that it is precisely this cohesive component which is aniso-
tropic. It is substantially less when the sample axis.
is horizontal. At higher stress levels, the bonding is
destroyed and the strength appears to be purely "frictional"

and isotropic. The behaviour of the highly sensitive Champlain


clay is perhaps an extreme example and not typical of soft
clays generally but it does clearly demonstrate that the
effective stress failure envelope need not necessarily be
isotropic.

An entirely different measure of structural anisotropy


is provided by making use of conventional oedometer tests
on samples cut with the normal_orientation and on samples cut
so that the compression axis is the horizontal (or lateral)
axis in the ground. Tests of this kind give an indication
of the relative "stiffness" of the soil structure in the
two directions. The first reported tests of this kind were
those of Zeevaert (1953), and were carried out on two samples
of Mexico city clay. His results showed two important
factors:
31

(a)The soil, skeleton is stiffer in the vertical


direction than in the horizontal direction up
to the value of the precompression pressure.
(b)The precompression pressure indicated by the
compression curves is lower for the horizontal
compression tests than for the vertical tests.

Tests on other soft clays have shown a similar


trend. Parry and Nadarajah (1974) give results from tests
on clayey silt samples taken from a site near York. Their
curves are shown in the lower part of Fig. 2.2.6. The
preconsolidation pressures are not well defined but the soil
is clearly more compressible in the horizontal direction than
in the vertical direction in the low stress range.
This difference in stiffness of the soil structure
in the two directions is of considerable significance as it
is possibly the main factor influencing the strength in
undrained compression tests at varying inclinations. The
difference in stiffness in the two directions would be expected
to make the pore pressure response dependent on the inclina-
tion of the compression axis and thus lead to a variation in
undrained strength with inclination.

2.2.5 Anisotropy of the Undrained Deformation Modulus


So far only strength anisotropy has been discussed
in detail. The question of anisotropy of the deformation
modulus is of considerable interest but the data available
on this aspect is very limited. Lo (1965) unfortunately does
not report on the deformation modulus in his tests on Welland
clay. De Lory and Lai (1971) in their unconfined compression
32

and semi-confined tests report that the strain to failure


is slightly greater in tests on horizontal samples than in
tests on vertical samples. The values respectively are 7%
and 6%, in semi-confined tests, and 15% and 12% in con-
ventional unconfined tests. These figures suggest that the
modulus difference between vertical and horizontal samples
is small and varies in the same way as the strength.
While it seems logical that the deformation modulus
should vary in the same way as the peak strength there is
some evidence to suggest that this is not necessarily the case.
Parry and Nadarajah (1974) report that in undrained
compression tests at varying inclinations carried out on an
undistrubed clayey silt the deformation modulus is greatest
for vertical samples and drops progressively to a minimum
value for horizontal samples. However, the strength in the
same tests is greatest for the vertical sample and least for
the 45° sample, with the horizontal sample having a strength
somewhat greater than the 45° sample. Boehler and Giroud
(1971) report on unconfined compression tests at varying
inclinations on samples trimmed from a remoulded clay subjected
to one dimensional consolidation. Their results are shown
in Fig. 2.2.7. The strength values show the same trend as
in the tests reported earlier with a progressive drop as the
angle i increases from 0 to 90°. The deformation modulus,
however, shows the opposite trend with the maximum value
being obtained from the horizontal samples. It appears that
there is insufficient data to be able to draw any definite
conclusions about the anisotropy of the undrained deformation
modulus in normally consolidated clays.
33
100

7.6 80
'>a;

60
_c

40

3 20


20 40 60 80 100
0/0 of strength with axis vertical

Two of three possible


planes on which
failure can occur

STRENGTH VARIATION WITH INCLINATION OF


COMPRESSION AXIS UNDRAINED TESTS ON
LIGHTLY OVERCONSOLIDATED CLAY ( after Lo,1965 )

FIG 2.2.1

34

Test
arrangement

180
Orientation of samples
and failure planes

(3= inclination of failure plane to vertical.


comp. axis •. lI

10
o
9
h 0
t O
eng G
80
tr o
fs
o
a ; 60
0 30 GO 90 120 150 18(
Angle

100
co
t.)
_c
Direction of 90 *en
shear on c >
▪ u.)
failure
plane 80 f;i rts
x
• -c
70
00
Case
60
90 60 30 0
Angle /3

RESULTS OF UNDRAINED SEMI-CONFINED TESTS


( after Delory and Lai 19 71 )

F I G 2.2.2

35
2,0
TONVERG
1.0

ASERUM 0

o o
DRAMMEN
oo 2.0 2
-
• 1,0

Pc C•1 0
MANGLERUD
0

2.0

1.0

0
1.0 1/3 2/3
0.51 D
3 H
Results from N.G.I. tests
( after Aas , 1967 )

Sv
0.5 1.0 1.5 2.0

Different curves
are from
different vane
combinations
E
q

*
ro
"
10 x

W
I

12
Results from a clay near Stockholm
( after Wiese1 51973 )

MEASUREMENT OF ANISOTROPY WITH FIELD


VANES OF VARYING DIMENSIONS

FIG 2.2.3
36

1.0

,..••■■■ 0.8
a_
;( E
0 0
w
0.6
_c _c
or o-•--
..••■•• '..
c c) .....*
a) a) ......
.•■•••

vi (/) 0.4 ......^" '''


....-
,o -----
.•■••^'
0
0

- ,
20 40 60 . 80 100
P.1.

Site P.I. Undrained Strength ( Ext. )


Undrained Strength Comp.)
Bangkok 88 0.52
Kimola 31 0.6 9
Drammen 31 0.40
( plastic)
Vaterland 20 0.35
Studentertunden 17 0.28
Drammen 10 0.21
( lean)

RATIO OF UNDRAINED STRENGTH FROM TRIAXIAL


EXTENSION AND COMPRESSION TEST AFTER
Ko CONSOLIDATION
( after Berre and Bjerrum, 19 73 )

FIG 2.2.4.
37
1.0

0.8

0.6

•••••■•

x E0 0.4
1.11 c.)
•■•■•■•

t 0.2


42 43 44 45 46
Porosity %
SAND
( after Bishop and Eldin , 1953 )

1.0

x
••■•■•••
0
-0

I I i I
0 2 4 6 8 10 12 14 16 18 20 22 ?L. 2(
Overconsolidation ratio
REMOULDED LONDON CLAY
( after Parry 1960

COMPARISON OF UNDRAINED STRENGTH FROM TRIAXIAL


EXTENSION AND COMPRESSION TESTS ON SAND AND
REMOULDED CLAY ( AFTER ISOTROPIC CONSOLIDATION )

FIG 2.2.5
38
20
/ Vertical Samples
/
/
• /
/ • - .
10 .
6-
.
.... .
0,1

6- d .
41i"
10 20 30 4
(TIC (T3-' psi
2
20
/ Horizontal Samples
/
/
/
/Vertical o
10 i samples
/

----

0 10 „ 20 30 40
4 ai psi
2
STRESS PATHS AND FAILURE ENVELOPES FROM
CONS. UNDRAINED TESTS ON CHAMPLAIN CLAY
( after Lo and Morin , 1972 )

0
..Vertical
..., ,,/
2
O
Horizontal
c4
0
'L7)
:l 6
0

cS 8

10
10 20 50 100 200
Consolidation pressure kN/m2

CONSOLIDATION TESTS ON A SOFT CLAYEY SILT


( after Parry , 1974 )

F I G 2.2.6.
39

Each point represents average of between


five to eight measurements

Compressive
Strength

150-
=30°

= 45°

100-2

Deformation
Modulus i= 60°

50_1

Compre sive Strength


(100kN /m2 ) 2 3
1 3P
50 100 150
Deformation Modulus ( 100 kN/mL )

ANISOTROPY OF UNDRAINED COMPRESSIVE STRENGTH


AND DEFORMATION MODULUS IN TESTS ON REMOULDED
ONE DIMENSIONALLY CONSOLIDATED CLAY

( after Boehler and Giroud 11971 )

FIG 2.2.7
40

2.3 Undrained Plane Strain Behaviour and its Relationship


to Triaxial Behaviour

2.3.1 Theoretical Considerations


The simplest method of gaining some idea of the
differences to be expected between plane strain and triaxial
behaviour is to consider the soil to behave elastically.
It is readily shown that in this case, assuming the soil to
be isotropic, the pore pressure response in undrained
compression (a3 = constant) is given by:

Au =(Aa1 - Aa3) for triaxial loading


Au = 1 - Aa3) for plane strain loading

The pore pressure is thus higher in plane strain compression


than in triaxial compression, and if the effective stress
failure envelope is the same in each case, then the undrained
strength in plane strain compression would be expected to.be
less than in triaxial compression. In the corresponding case
of extension loading (63 constant, al decreasing) the change
in pore pressure will again be higher in plane strain than in
triaxial loading. However the pore pressure change is now
negative since al is decreasing so that its effect will be
the opposite to that in compression loading, ie the undrained
strength should now be higher in the plane strain case.
The assumption of elastic behaviour can also be used
to indicate likely differences in the slope of the stress
strain curves. For an isotropic soil, the slope of the stress
strain curve in undrained loading is given by
41

Au1 _ 3E'
for triaxial case
Ae1 2(1+11')

where Ae1 = strain resulting from stress increase Au1

E' = drained modulus


p' = Poisson's Ratio
Au1_ 2E' 4E'
and for plain strain
Ae1 l+p' 2(1+p')

Thus according to elastic theory, the undrained


stress strain curve will be steeper by a factor of 4/3 for
plane strain loading than for triaxial loading. While it
is recognised that the soil is unlikely to behave elastically,
the above relationships provide at least an indication of
expected trends between plane strain and triaxial behaviour.
Parry (1971) has used an entirely different approach
in an attempt to relate undrained strength in triaxial
extension and plane strain compression to that in triaxial
compression. Parry makes use of the following assumptions:
(a)The soil reaches the critical state at which
there is a unique relationship between void ratio
and mean effective stress (a'
1 + u' + u 1 ).
2 3
(b) At the critical state c' = 0 and the value of
(I)' is independent of stress path.
a'
+2
(c)For plane strain , = 0.4.
a1
3

From these assumptions Parry deduces that the


undrained strength in plane strain compression should be
between 0.87 and 0.95 of strength in triaxial compression,
depending on the value of 4)1 . These figures however apply
42

to the ultimate or "critical" state, and the theory on which


they are based cannot be used to predict peak strengths if
these are not coincident with critical state strength.

2.3.2 Tests on Normally Consolidated Remoulded Clays


The only extensive series of undrained plane strain
tests on remoulded clay appears to be that reported by
Henkel and Wade (1966). These tests were carried out on
samples of normally consolidated remoulded Weald clay. A
parallel series of triaxial tests were also carried out and
in each case the samples were Ko consolidated prior to the
undrained shearing. The results of these tests are summarised
in Fig. 2.3.1 and in Table 2.3.1. The main trends shown by
the results are as follows:
(a)The undrained strength is about 8% higher in
the plane strain tests than the triaxial tests.
(b)There is a better defined peak in the plane
strain tests and the strain to failure is less
than in the triaxial tests.
(c)The pore pressure change is greater in the
plane strain tests than in the triaxial tests.
(d)Assuming c' to be zero, the plane strain tests
indicate a (1)1 value slightly higher than the
triaxial tests.

2.3.3 Tests on Undistrubed Samples


There are' two series of tests on undisturbed clays
which are of considerable interest. They are particularely
valuable as in both cases the soil has been brought to failure
43

from the Ko stress state by a compression and an extension


type of loading. The design of the plane strain apparatus
for both series was such that the magnitude of the horizontal
stress, which becomes the major principal stress during an
extension loading, could be made greater than the inter-
mediate principal stress (ie the stress acting on the end
plattens used to maintain the plane strain condition).
In the two versions of the plane strain apparatus which have
been used at Imperial College it has not been possible to
test samples with an extension type of loading as the use
of an all round cell pressure on the sample means that
the horizontal pressure cannot be greater than intermediate
principal stress and thus must always be the minor principal
stress.
The most comprehensive of the two series is that of
Vaid and Campanella (1974) who carried out tests on a clay
of medium plasticity (WL = 44, Wp = 18) known locally as
Haney clay. In this series the samples were consolidated
under Ko conditions in both a plane strain apparatus and
in a triaxial cell, and then sheared undrained by increasing
or decreasing the vertical principal stress. All samples
were consolidated to an effective vertical stress of 6 kg/cm2,
which was substantially higher than the in situ stress.
The results are summarised in Fig. 2.3.2 and in Table 2.3.1.
Each test was in fact carried out twice, using the two
possible methods of loading in each case. (For example the
compression test was carried out with constant and QV
increasing and with QV constant and aH decreasing). The
results have been plotted in Fig. 2.3.2 in terms of the
44

constant aH, variable av method of loading. The results


show the following:
(a)The strength in extension loading is
substantially less than in compression loading
for both types of test, although the difference
is less in plane strain than in triaxial loading.
The ratio of extension strength to compressive
strength is 0.71 for plane strane and 0.62 for
triaxial loading. Although this difference is
not great the tests do confirm that triaxial
extension and compression tests tend to over-
estimate the anisotropy of the undrained.
strength except for the special cases of axi-
symmetric loading mentioned earlier.
(b)The plane strain strength is higher than the
triaxial strength in both types of loading.
It is 10% higher in compression and 26% higher
in extension loading.
(c)The stress strain curves have a much sharper
peak in compression loading than in extension
loading and the strain to failure is correspond-
ingly much lower. The scale used in plotting
the data does not make possible a careful
comparison of the slope of the stress strain
curve between plane strain and triaxial loading,
but the general appearance of the curves suggests
that the slope may be marginally steeper in
the plane strain test.
(d)The pore pressure change does not vary greatly
between the two types of test, although the
45

compression tests show the same trend as in


Henkel and Wade's tests, that is slightly
higher pore pressure in the plane strain tests.
(e) The cp' values assuming c' to be zero are slightly
higher in plane strain than triaxial loading,
and several degrees greater in extension loading
than compression loading.

The second series of tests are those of Duncan and


Seed (1966) carried out on undistrubed samples of
San Francisco bay mud (WL = 88, Wp = 43). The tests in this
case were only plane strain tests so correlation with
triaxial behaviour is not possible. The samples were all
initially consolidated under Ko conditions to a range of
starting pressures, all higher than the in situ stress
level. The samples were then sheared undrained by both
compression loading and extension loading. A series of
results were thus obtained covering a range of initial
consolidation pressures. The results however showed the
same trend regardless of the value of the consolidation
pressure. Results for an initial vertical effective
consolidation pressure of 3.7 kg/cm2 are shown in Fig. 2.3.3
and the results of the series are summarised in Table 2.3.1.
They show the following trends:
(a)The general shape of the stress strain curves
are very similar to thos of Vaid and Campanella,
although the peak in the compression curve is
less distinct and the strain to failure much
greater.
(b)The ratio of strength in extension to strength
in compression is 0.75.
46

(c) The cb' value is surprisingly greater in the


compression tests than the extension tests.

It is worth noting that these two series of tests


show very similar results despite the difference in apparatus
and technique used in the tests. The procedure used by
Vaid and Campanella was more or less the same as that used
in standard triaxial testing with the vertical axis in the
ground also being the vertical axis in the apparatus. The
vertical stress is thus always applied by the rigid top and
bottom plattens. In Duncan and Seed's tests the procedure
for compression loading was the normal procedure, but for
extension type of loading the samples were set up in the
apparatus with their original vertical axis now horizontal.
Thus the major principal stress at the start of the test
(a V in Fig. 2.3.3) was applied by the flexible membrane and
the rigid plattens were supplying the minor principal
stress (ie the horizontal stress in the ground, am in
Fig. 2.3.3). This method resulted in a rather cumbersome
method of applying the stresses during the shearing stage.
Despite these differences in procedure the two series of
tests showed consistent results. The curves in Fig. 2.3.3
are not in quite the same form as given by Seed and Duncan.
They have been replotted so that the presentation is
identical to that used for Vaid and Campanella's tests.
To relate the results more readily to actual field
situations they have been replotted in Figs. 2.3.4 and
2.3.5. These figures show the vertical and horizontal
stresses and the pore pressure response when the sample is
47

brought to failure either by increasing the vertical stress


(ie element 4 in Fig. 2.1.1) or increasing the lateral
stress (element 3 in Fig. 2.1.1). Presented in this way
the variation in pore pressure with type of loading is more
readily assessed. Failing the soil by increasing the
horizontal stress involves a greater change in stress to
bring about failure than increasing the vertical stress
and thus also a greater pore pressure change. The pore
pressure parameter Af in Table 2.3.1 corresponds to the
loading sequence in Figs. 2.3.4 and 2.3.5 and it is seen
that the Af values are all fairly close to unity. This is
in considerable contrast to the tests on remoulded Weald clay
which gave A values of 1.7 and 2.0. It is of interest to
note that at large strains the values of the principal
effective stresses tend to converge for the two types of
loading.
Apart from this difference in A values the three
sets of tests show reasonable agreement although they are
perhaps at variance with theoretical considerations in that
the undrained strength in compression loading is greater
for plane strain than for triaxial loading. The pore pressure
response is slightly greater in plane strain than in tri-
axial loading as elastic theory suggests, but this is
offset by an increased q value which leads to the higher
undrained strength.
It should be remembered that these tests have been
carried out at relatively high stress levels and that at
the in situ stress level the behaviour may be somewhat
different. Conclusions drawn from the above tests assuming
48

that the value of c' is zero may not be valid if the soil
has a cohesive component of strength resulting from bonding
between the particles. The effect of such bonding is likely
to be obliterated by the use of high stress levels.
49

Table 2.3.1. Results of Consolidated Undrained Triaxial


and Plane Strain Tests on Normally
Consolidated Clays

Failure Values
Test Type G"-- Cr Strain
1 3 A
2 V yo ciSI f
lc

Remoulded Weald Clay after Henkel and Wade ,1966 )

Triaxial Comp 0.26 6 25.9 1.7


Plane Str. I/ 0.28 2 27.1 2.0

Undisturbed Haney Clay [after Void and Campanella 0974 )

Triaxial Comp. 0.268 0.35 29.8 0.94


Plane Str. I/ 0.296 0.4 31.6 0.97
Triaxial Ext. 0.168 -13 33.8 1.18
Plane Str; ii 0.211 -10.5 34.3 0.93

Undisturbed San Francisco Bay Mud ( after Duncan & Seed ,1966)

Plane St r. Comp. O. 37 3.6 38 1.12


Plane Str. Ext. 0.28 10.2 35 0.70

Note: 1. 0-/ = vertical effective consolidation pressure.


2. All samples initially Ko consolidated.
3. Undisturbed . samples all consolidated to
pressures higher than those in situ.
4. In calculating the value of Af in extension tests
from the formula A - 6'u - '° the stress
Llo1- — Ac3
07 has been taken as the major principal
stress at failure ( = vH , the horizontal stress )
50

0.6

r
— ..............
0.5
0--0- 3
1
arc
0.4
Au
c2'
03
,
Cr'
ic
, _____________
.....,
... ......
.,
/ Triaxial ----.
Plane strain

. 1 1
0 1 2 3 4 5 6
Axial Strain

UNDRAINED PLANE STRAIN AND TRIAXIAL TESTS


ON REMOULDED WEALD CLAY
( after Henkel and Wade 0966 )

FIG 2.3.1
51

Plane strain -
Triaxial ,

......

Compression
0— constant
H
CT-V increasing

J0
Extension
/
0— constant
H 1
0— decreasing /
V

...--_...--- .1
5r =8
WL -
- 44
WP = 18

16 12 8 4 0 4 12 16
Axial Strain

16 12 8 4 0 4 8 12 16
kg I cm2

■ .... ...■
f

.,....• ..=.•
.....
.....
....''
00"

if
lam••• mom= =Emma* MIN..,
'..... ........

...............
\N. •

—1

UNDRAINED PLANE STRAIN AND TRIAXIAL TESTS ON


UNDISTURBED SAMPLES OF 'HANEY' CLAY
AFTER Ko CONSOLIDATION
( after Vaid and Campanella, 1974 )

FIG 2.3.2
52
4

"EV
• ,
cn
..

bx
1
6' Compression
1 CF constant
H
CTV increasing

Extension
CF constant
H
Cr decreasing
V

Sr = 8
Wi. = 88
W43

-16 -12 -8 -4 0 8 12 16
Axial Strain
-16 -12 -8 =4 0 4 8 12 16
a NJ
g /cm2

-2
UNDRAINED PLANE STRAIN TESTS ON UNDISTURBED
SAMPLES OF SAN FRANCISCO BAY MUD AFTER
Ko CONSOLIDATION
( after Duncan and Seed ,1966 )

FI G 2.3.3
53

(.1
E
U
5 :1)
Lu

H initial
Mamma •••■•■•• al••■• ari ••■■•• ••••■•■•■ 111111•Mell• I•1■I

OH increasing
constant 0— constant
0-V
0— increasing
V

-16 —12 —4 0 4 8 12 16
Vertical Strain

UNDRAINED PLANE STRAIN BEHAVIOUR OF UNDISTURBED


HANEY CLAY STARTING FROM Ko STRESS STATE
( after Vaid and Campanella , 1974 )

FIG 2.3.4
54

Cr increasing

CT- constant 7 0— constant


E 0— increasing
U V
0,
6

-_ V (initial)

■••■••••

16 12 8 4 0 4 8 12 16
Vertical Strain

UNDRAINED PLANE STRAIN BEHAVIOUR OF UNDISTURBED


SAN FRANCISCO BAY MUD STARTING FROM Ko
STRESS STATE
( after Duncan and Seed 1966 )

FIG 2.3.5
55

CHAPTER 3

EQUIPMENT

3.1 The Stress Controlled (or Hydraulic) Triaxial


Apparatus

Conventional triaxial equipment is not at all well


suited to tests other than standard tests, that is tests in
which c 3 is held constant and a1 is increased by loading at
a constant strain rate until failure occurs. Tests in
which a1 or c 3 (together or separately) are varied in a
controlled manner can only be carried out if considerable
modifications are made. Extension tests similarly are only
possible if modifications are made.
In short, conventional apparatus is not well suited
to the stress controlled type of stress with which this
thesis is largely concerned, and the writer was faced at the
outset with the choice of either modifying existing equipment
or of designing a completely new triaxial apparatus capable
of carrying out any type of test in which the two stresses
are varied in some controlled manner. As it seemed that
future trends in research would increasingly necessitate
the use of apparatus of this type it was decided that the
design and construction of a new apparatus was preferable to
the modification of existing equipment.

3.1.1 Design of the Apparatus


Ideally, a triaxial apparatus for controlled stress
testing would be one in which the two stresses al and a3
56

can be controlled and varied entirely independently of each


other. However, such an arrangement involves practical
difficulties as the following discussion shows. The means
by which a triaxial sample is set up, consolidated, and then
tested by applying an axial load to increase (or decrease)
the value of a 1, means that a1 is actually the sum of two
independently applied stresses, that is a3 and the deviator
stress al - a3. Thus in a conventional arrangement:

(P - a 3 a)
1 = G3 A A

where P = load applied to loading ram


a = loading ram area
A = sample area
(The effect of the ram and top cap weight is ignored).
The value of a 3 can normally be varied at will and
with a suitable loading system the value of. P may also be
varied at will. However, the value of a
1 can only be
controlled directly if a3 is held constant. In any test
in which a3 is varied, the operation of changing a3 will at
the same time produce a change in the value of al, the magni-
tude of which will depend on the ratio A. In order to
maintain control of a 1 it is necessary to vary the value of
P at the same time in such a way as to compensate for the
effect on al of the change in a3.
Thus to carry out a test in which al is to be held
constant and a3 reduced it is necessary to increase P at a
certain fixed proportion of the amount by which a3 is
decreased.
57

The only way in which the value of al can be made independent


of changes in a3 is to make the loading ram area equal to
the area of the sample; that is a = A and a l = P/A. With
conventional equipment this is not really a practical
proposition due to the excessive size and weight of such a
ram and the large friction force it would involve.
In designing the new apparatus a great deal of thought
was given to the advantages of an arrangement in which al and
a would be entirely independant of each other but in the
3
end practical considerations took precedence and such an
arrangement was not incorporated into the apparatus.
The new apparatus is essentially a development of
several existing items of equipment in use at Imperial College,
in particular the creep cells described by Bishop et al.
(1973) and the plane strain apparatus built by Atkinson (1973).
The vertical load in both cases is applied by means of fluid
pressure acting on a piston and loading ram. Leakage and
friction are virtually eliminated by the use of Bellofram
rolling seals. With suitable ancillary equipment this method
of loading is ideal for both controlled stress and controlled
strain loading and was the logical choice for the present
apparatus.
The new apparatus is shamndiagrammaticallyin Fig. 3.1.1.
The upper part is very similar to a conventional triaxial cell
except that the vertical load is applied by moving the sample
pedestal upwards from below and pushing the top cap against
a stationary load cell, which records the load. The pedestal
is mounted at the top end of a loading ram at the bottom end
of which is a piston and pressure chamber. Bellofram rolling
58

seals are used to retain the cell fluid and the fluid in the
pressure chamber, and the ram travels up and down in a
"Rotolin" linear bearing. Figs. 3.1.2 and 3.1.3 show scale
drawings giving further details of the apparatus. The axial
load is applied to the sample by increasing the pressure in
the bottom pressure chamber. The loading ram has a cross arm
attached to it which in turn supports two vertical rods which
travel up and down through holes in the cell base. The vertical
movement of the ram and thus the axial strain is measured by
dial gauges (or displacement transducers) mounted on the top
of the cell. One rod would be sufficient but two were used
in order to keep the ram bal-.anced and also to enable both a
dial gauge and transducer to be mounted at the same time.
A large number of alternative arrangements were con-
sidered before arriving at this design. At first sight it
seems more logical to load the sample from above, as in
conventional tests, but the arrangement used with the load
applied from below has several advantages. In particular,
setting up the sample is very simple, as in a conventional
apparatus, and no corrections are required for loading ram
weight. The upper part of the cell which is removed for
mounting or dismantling the sample remains light in weight
and easy to handle. Also the weight of the loading ram acts
against the lower Bellofram and helps maintain a positive
pressure in the pressure chamber so that there is no danger
of damage to the Bellofram through it being turned "inside
out". There is a. danger of this occurring when the load is
applied from above as the weight of the piston and ram
pulls down on the Bellofram instead of acting against it.
59

The use of two Bellofram seals perhaps requires some


explanation since it would be possible to use only one
Bellofram seal to simply separate the cell fluid from the
pressure chamber fluid. This is the arrangement used in . the
creep cells at present in operation at Imperial College.
However there are several important advantages in using two
Belloframs, the main ones being the following:
(a)Extension tests, that is tests in which the
horizontal stress is greater than the vertical
stress, can be carried out. These would not be
possible with a single seal as the pressure on
the pressure chamber side of the Bellofram would
always have to be greater than the cell pressure.
(b)The linear bearing is not submerged in the cell
fluid or loading chamber fluid, so that the use
of oil in either of these is unnecessary.
(c)Strain measurements can be made' externally by
measuring the movement of the loading ram at any
point between the Bellofram seals.

In its method of operation the apparatus is identical


with Atkinson's plane strain apparatus. It should be noted
however that it differs in several practical details.
First, the vertical strain measurement is made
externally by means of the cross arms on the loading ram
mentioned earlier, whereas in the plane strain apparatus it
is made internally by means of an oil filled dial gauge
submerged in the cell fluid. The internal measurement has
the advantage that it is a direct measurement so that corrections
60

for load cell deflection are not necessary. The external


method has the advantage that electrical transducers can
easily be mounted in place of the dial gauges. An internal
dial gauge however was not really a practical possibility
with the present apparatus because of the large amount of
space it would require.
Secondly, the pore pressure and drainage loads from
the base of the sample pedestal are taken out directly via a
hole drilled down the centre of the loading ram and out
through slots in the spacer block immediately above the lower
pressure chamber. In the plane strain apparatus these
leads are taken from the lower loading platten and out through
holes in the base of the cell.
Thirdly the new apparatus has the load cell mounted
on a threaded shaft fitted with an adjusting nut, so that
it can be raised or lowered at any time during a test. After
mounting and consolidating the sample the load cell can be
gradually lowered to bring it into contact with the upper
loading cap. With the plane strain apparatus the position
of the load cell must be fixed before the start of the test,
that is before filling the cell with water.

3.1.2 Principle of Operation


The operator using the apparatus will generally desire
to vary al and 03 in some required manner and measure the
resulting deformation of the sample and the pore pressure
response. However the two pressures which can be controlled
directly are in fact the cell pressure (a3) and the pressure
in the lower chamber (p). The key to the operation of the

61

apparatus is thus the relationship between a1, a3 and p. By


considering the equilibrium of the loading ram we can obtain
the relationship between these pressures.
Let A = sample area
a = effective Bellofram seal area
W = total weight of loading ram, sample pedestal,
cross arm (ie all the moving parts)
p = pressure in lower pressure chamber.
Then the downward force on ram = a1A + a 3(a - A)

and upward " 11 tl
= p a
Weight of ram = W

so thatpa=a1A+ a3(a - A) + W 3.1.a

Rearranging we have

a a W 3.1.b
al = PT '
13(1 - A) - A

In the present apparatus A = 1.767 in2


a = 4.56 in2
so that a/A = 2.58

Equation 3.1.b is correct whether al is greater or


less than than a3, although of course a1 can only be made
less than a by attaching the sample top cap to the load cell.
3
It should be noted that with the Imperial College
type load cells used in the apparatus, the load cell measures
only the force exerted on it by the sample top cap and does
not measure the thrust from the cell fluid. The force acting
on the load cell ram is therefore not the same as the load
measured by the load cell.
62

If P = thrust on load cell ram


and a' = area of
then by considering the equilibrium of the top cap we have:
P + a 3(A - a') = a 1A
or P - a a' = A(a 1 - a 3)
3

The load cell measures P - a 3a' and so provides a direct


measure of the deviator stress.
For any particular test it is easy to determine from
equation .3.1.b the way in which the pressure p must be varied
in relation to a 3 in order to produce a required stress path.
To do this it is easier if we write equation 3.1.b in terms of
stress change.
ie Aa 1 = Ap A + Aa 3(1 - A 3 . 1. c

We will consider several typical cases in turn.


(a) Tests in which a 3 is constant and a 1 is increased
or decreased (ie conventional tests).

In this case the test is quite straightforward and


involves simply increasing or decreasing the value of p. The
change in al is related directly to the change in p:

Au1 = A AP

It is worth noting that with this conventional type of test


the application of al may be on a controlled stress or a
controlled strain basis. By varying p in a controlled manner
the test can be run as a controlled stress test, or
alternatively by pumping fluid into (or out of) the pressure
chamber at a controlled rate of flow the test can be run as

a controlled strain rate test.


63

(b)Tests in which a 1 is held constant and a 3 is


decreased or increased.
In this case we have from equation 3.1.c

tip = Aa3(1 -

Thus to maintain a 1 constant, the value of p must be


increased or decreased by a certain fixed proportion of the
change in a3.

(c)Tests in which the sum of a 1 and a 3 is held


constant.
In this case we have

Aa l + Aa3 =Lip(A) + Aa3(2 - =0


and Lip = Aa3(1 - 22)

Again p must be varied by a certain fixed proportion


of the change in a3.

(d)"Ko1' tests
In this case we may wish to increase al and a3
in a certain fixed proportion to each other.

ie Aa 3 = K oA a1

ApK
Thus R; a + Aa3(1 -

And by rearranging
1 - Ko
A )
tip = Aa3 1 a( Ko
Again p must be varied by a fixed proportion of a3.
64

To conclude this section it is worth noting that


the apparatus is well suited to carrying out tests in which
stresses are applied in a controlled stress or controlled
strain manner and are then held constant for a certain period
of time. It is also very simple to switch from controlled
stress loading to controlled strain rate loading. A photo-
graph• of the apparatus is given in Fig. 3.1.4 and the
bayonet type fitting used for connecting the top cap to the
load cell for extension tests is shown in Fig. 3.1.5.

3.1.3 Method of Application of Controlled Stresses


The new apparatus has not been designed to suit any
particular method of pressure application and can therefore
be used with any pressure system which can supply and
continuously vary pressure in a controlled manner. For the
tests described in this thesis the existing mercury pot
pressure sources have been used and a small, drive unit built
to raise or lower the mercury pots as required. This drive
unit is in principle the same as that used by Davies (1974)
for applying the initial stresses in his creep tests. It has
been described breefly by Bishop et al (1973).
A diagrammatic view of the drive unit is shown in
Fig. 3.1.6. It consists of a motor which drives directly
Kopp Variator A which in turn is linked directly to a two
stage gearbox reduction, which reduces the speed to that
required for driving the mercury system winch drum. From
the drive shaft between Kopp A and the first gearbox a belt
drive is taken to drive a second Kopp Variator B. This is
connected in the same way to a two stage gearbox reduction
system. We thus have two drive shafts to connect to two
65

mercury pot winch drums which constitute the two required


pressure sources. The Kopp Variators can be used to vary
the output speed by any value between 1/3 and 3 times the
input speed. Thus the speed of rotation of the second
drive shaft is linked directly to the speed of the first
shaft and can be set at a fixed ratio of this speed by
appropriately setting Kopp B. Any ratio of speed between
the second and first drive shafts can be achieved by using
suitably different gearboxes. These are available with a
very wide range of speed ratios. Thus the stress path in
any test is controlled by the ratio of the speed of rotation
of the two shafts, ie by the setting of Kopp•Variator B,
and the speed of the test by the setting of Kopp Variator A.
If the range of speeds obtainable from the Kopp Variators is
insufficient then different reduction gearboxes can be
mounted. It is also possible to make the second drive
shaft rotate in the opposite direction to the first shaft
by replacing the connecting belt drive by a gear wheel '
drive. This may be necessary to carry out tests in which
thesumofa1 and a3 is to be maintained constant.
For each drive unit callibration curves can be
prepared of the type shown in Fig. 3.1.7. The lower graph
shows the rate of rotation of the first drive shaft in
r.p.m. for various settings of Kopp A. Also shown is the
actual rate of pressure change produced in the mercury pot
system.. This graph applies only to a particular set of
reduction gear boxes and would need to be recalculated if
different gear boxes were mounted. The upper graph shows
the ratio of the output speed to input speed for the second
66

Kopp Variator. This gives directly the ratio of the rates


of pressure change provided the same reduction gear boxes
are again used.

3.1.4 Some Practical Details


The question of the influence of friction on the
operation of the apparatus needs to be considered briefly.
As mentioned earlier, the loading ram travels up and down
in a "Rotolin" linear bearing and the cell and pressure
chamber fluid are retained by "Bellofram" rolling seals.
Both the bearing and the bearing and the Bellofram seals
offer some resistance to movement and an attempt was made
by two different methods to measure this "frictional"
resistance. The first method was simply to install a dummy
metal sample and then increase the pressure in the pressure
chamber by stages and measure the resulting thrust on the
load cell at each stage. The thrust was again measured as
the pressure was reduced in stages. This was done with a
constant cell pressure of 200 kN/m2 which was the lower
limit of the range of cell pressures used for most of the
tests described in this thesis. The difference in the
thrust on the load cell during the increasing and decreasing
stages was a measure of the frictional resistance. The
results are shown in Fig. 3.1.8. The load cell thrust is
plotted against the theoretical thrust, ie the difference
between the ram and cell pressure multiplied by the Bellofram
area. The plot should be a straight line inclined at 45°
with the intercept on the x-axis equal to the weight of the
ram. This is seen to be the case but there is a small
departure due to the frictional resistance. The maximum
67

difference in the load between the increasing and decreasing


paths (ie upward and downward movement of the ram) was
just under one pound suggesting that the resistance to move-
ment in each direction was under 0.5 lb. The second method
of estimating the resistance was to measure the pressure
required to just move the ram in the upward direction and
then reduce the pressure until the ram just began to move
in the downward direction. A pressure transducer was used
to measure the pressure and a volume gauge installed in the
pressure line to detect the upward or downward movement of
the ram. The measurements were made without a sample in
the apparatus but with a constant cell pressure of 200 kN/m 2 .
Some ten to twelve determinations were made covering the
full travel distance of the ram. These measurements showed
a very consistent difference between the pressure required
to move the ram in each direction, which when converted to
a load was between 0.8 and 0.9 lb. This method thus gave
an almost identical value for frictional resistance as the
first method.
This frictional resistance is small and is generally
of no consequence when using the apparatus. The deviator
stress is determined directly from the load cell so that it
has no influence on the accuracy of measured stresses. The
only time it may be of significance would be in tests in-
volving a change in the direction of travel of the loading
ram. Even in this case it would only be of importance when
very soft soil was being tested and when particular stress
paths were to be followed.
The procedure adopted by the writer in carrying out
most tests was to set up the sample, apply the cell pressure
68

and then raise (or lower) the pressure in the pressure chamber
until the loading ram moved just enough to register a reading
on the load cell. At this stage therefore the pressure was
just sufficient to balance the weight of the ram and over-
come the frictional resistance. From this stage onwards
the pressure was raised (or lowered) as required and it was
in fact only the pressure change from the initial position
which was of importance. The stress path was controlled by
measuring the cell pressure and the deviator stress and then
making adjustments to the Kopp variators as required. The
actual value of the pressure in the pressure chamber was not
normally measured.
When the first two of the new cells were completed
(referred to as Unit 1 and Unit 2) trial tests were carried
out using dummy samples to check the performance of the
cells and the drive units for raising and lowering the
mercury pots. With the first cell (Unit 1) a test was carried
out to simulate the consolidation of a sample with a fixed
ratio of a3 to al' in this case a ratio of 0.55. The result
is shown in Fig. 3.1.9 and it is seen that the measured
stresses are very close to the intended (required) stresses.

With the second apparatus a test was simulated in which al was


to be held constant and a3 steadily decreased. The measured
values in this case are shown in Fig. 3.1.10, and the agree-
ment is again very good. In neither case was any adjustment
made during the test.
With real samples there is the added complication of
the change in cross sectional area of the sample as the
test proceeds. However, with undisturbed clays the strain
to failure is generally small and this effect is of quite
69

minor importance.
A full description of the apparatus is to be published
shortly (Bishop and Wesley 1975).

3.1.5 Performance of Load Cells and Transducers


The load cells installed in the new apparatus were the
standard Imperial College type in use in the soil mechanics
laboratory but had an increased sensitivity obtained by the
use of "maraging" steel for the triangular strain gauge
plate. Using this steel the cells had a range of up to
1000 lb and a sensitivity of about 45 mV (micro-volts) per
lb when measurements were made using the "Solartron" data
recorder. In terms of deviator stress this was about 10 mV
per kN/m2 as the sample area is close to 10 sq. cms.
The cell pressure and pore pressure were measured on
Bell and Howell pressure transducers which had a sensitivity
of about 60 mV per psi, that is about 10 mV per kN/m2 so that
the accuracy of these measurements was of the same order as
that of the deviator stress.
Some minor difficulties were encountered in using the
load cells and cal. ibrations required special care because of X
the very low stress level involved in most of the tests. In
the majority of tests the failure deviator stresses were in
the range of 20 to 40 kN/m2 , equivalent to loads of about
20 to 40 Newtons (3 to 6 lb). The three load cells made for
the three new sets of apparatus were made to the same dimen-
sions within manufacturing tolerances but when cal'Abrated it )\
was found that one of the load cells performed much better
than the other two. It was more sensitive and the loading and
unloading readings were almost identical. Also, there was no
70

zero reading shift after loading or between loadings. An


examination of the cells showed that this cell with the
superiour performance had a very small amount of "play"
between the strain gauge plate and the cell base.- With the
other two cells no play at all could be detected.
It appears that with a tight fit the groove produces
a cantilever action in the triangle so that there is less
deflection for a given load and thus less sensitivity. Also,
the friction between the strain gauge plate end points and the
base groove results in a hysteresis effect and a shift in the
zero readings. Thus, there is a choice to be had between a
tightly fitted cell with no "play" to correct for in deflec-
tion measurements, or a loosely fitted cell with significantly
better performance (at least at low loads) but with "play"
to correct for. In the writer's view the loose fit appeared
the better choice and the two tight cells were in fact dis-
mantled and new bases mounted which allowed a very small
amount of play.
The difference in performance before and after this
operation is illustrated in the lower part of Fig. 3.1.11.
The graph shows the output in mV versus the load on the cell.
The hystersis effect with the tight fitting base is seen
clearly, while with the loose fit the loading and unloading
readings are identical. Also shown on the same graph are the
readings when the cell is outside the base altogether and
simply sitting on its upper edge. In this position the
readings are almost identical with those when the cell is
installed in the loose fitting base.
A further factor in load cell performance which was
looked into was the effect of the starting position on the
71

load celireadings. In an actual test the load cell is held


by its ram (shaft) and the sample cap is pushed against it.
However, calibrations are normally carried out with the load
cell sitting on its base as the starting position. The
upper part of Fig. 3.1.11 shows the calibration of one of
the load cells carried out by both methods. In the first
method the load cell was simply placed on the bench and
dead weights placed on top of it. In the second method the
cell was held by its ram and an upward thrust applied to it by
using the ram of a triaxial cell. The pressure in the cell
was raised in stages and the pressure acting on the ram pro-
vided the upward thrust. A rotating bush at the top was used
to eliminate friction. The results shown in Fig. 3.1.11
indicate that the same cal: ibration was obtained regardless
of the starting position. This is perhaps not surprising
especially in view of the fact that the cal:_ibrations of the
cells in compression and extension were almost identical.
Because of the small amount of play in the load
cells and because they deflect significantly under load each
cell was cal- ibrated for deflection as well as for mV output ›;
versus load. These calibrations were done with the load
cells installed in the apparatus and a brass "dummy" sample
set up so that the load was applied to the load cell in the
same way as it would in an actual test. Calibrations for
two of the load cells obtained in this way are shown in
Fig. 3.1.12. Each cal ibration was carried out a number of X
times and an average curve drawn through the points. The
scatter of the points is however too small to be of any
practical importance. The deflection characteristics of the
two cells are seen to be quite different. It appears that
72

there is much more "play" in one of the cells than in the other.
It should perhaps be noted that cal .ibrations prepared in
this way may include some minor deflection or "take up"
in the apparatus itself, although with all components being
of steel and of substantial cross section it is difficult to
see where any significant deflection could occur.
Strain measurements were made by both normal dial
gauge and resistance type deplacement transducers. These
transducers were commercially produced and consisted of a
tapered wedge travelling between and deflecting metal strip
springs on which strain gauges were mounted. Check tests
were carried out on two of these transducers by recording the
mV reading at each 0.01 inch travel over a travel range of
about 0.3 inches. The results are shown in Fig. 3.1.13.
It is seen that the output does not vary with travel although
the output from one of the transducers is significantly
higher• than the other. However with an output of around
100 mV for 0.01 inch the transducers are accurate to almost
0.0001 inches, which was adequate for the tests described in
this thesis.
The load cells, pressure transducers, and displace-
ment transducers were thus all of the resistance type and
measurements were all made using a "Solartron" data recorder.
This printed readings on a paper strip and it could be set
to read at any interval between 10 secs and 2 hours. A view
of the apparatus set up alongside the "Solartron" recorder
is given in Fig. 3.1.14.
73

DIAL GAUGES
--FOR STRAIN
MEASUREMENT

LOAD - CELL

S AMPLE

CELL
PRESSURE

LINEAR MOTION
BEARING
BELLOFRAM
SEALS

DRAINAGE AND
PORE PRESSURE
LEAD

-4--= VERTICAL LOAD


CONTROL PRESSURE

STRESS CONTROLLED TRIAXIAL APPARATUS:


I AGRAMATIC VIEW

FIG 3.1.1.
74

Adjusting nut to raise


or tower load cell

Displacement gauge
transducer

Load cell

Adjusting screw

Top cap

Soil sample

Perspex cylinder

Rods to activate dial gauge


and transducer for
strain measurement.

Loading
ram Hole at centre of loading
ram to take pore pressure
leads to sample pedestal
Linear
bearing
Bellofram Seals

Inlet to apply pressure


Scale . 1 actual size to loading ram
2.54

SCALE CROSS SECTION OF


STRESS CONTROLLED TRIAXIAL APPARATUS

FIG 3.1.2
7

173
3

PART A PART

Locking
nut

Bel lof ram


cap

Crossarm

/././

B el lof ram Slot for


cap crossar m

Locking
nut

STRESS CONTROLLED TRIAXIAL APPARATUS COMPONENTS

FIG 3.1.3
PHOTO GRAPH OF THE APPARATUS

F I G 3.1.1,
77

Load cell
base

0.°
10.4
ORM/

Stainless steel
connector

Perspex t p cap

DEVICE FOR CONNECTING TRIAXIAL SAMPLE TOP


CAP TO LOAD CELL FOR EXTENSION TESTS

FIG 3.1.5
DRIVE SHAFTS TO
MERCURY POT WINCHES

KOPP VAR IATOR B


MOTOR -

POSITIVE BELT
DRIVES

TWO STAGE REDUCTION


GEAR BOX ES

0 KOPP VAR IATOR BEAR! NGS

VARIABLE SPEED DRIVES TO RAISE AND LOWER MERCURY POTS


79

1.5
KOPP B
. y

1
cu
CD

S. 0.5
0.
C
O

0 10 0 10
0 10 0
1

2 3 4
Kopp Setting

2.5 2.5
KOPP A

1.5
0
0
q-q \ 10.
4-;

*
o 1 "E
a.
z
O .-V

5:
0.5 0.5 cr


0 0 10 0 10 10 0
2 4
Kopp Setting

CALLIBRAT ION OF KOPP VARIATORS

FIG 3.1.7
80

150

( 0.9 lb )
approx. 4N
/4
— 100
Ui
0

z
. ./.

a.
rd

c
50
0

/
111
0
—J

/
-4-Weight of ram
--) /
plus sample


50 100 150
( Ram Pressure — Cell Pressure ) X Ram Area ( Newtons )

RELATIONSHIP BETWEEN RAM PRESSURE , CELL PRESSURE


AND LOAD ON SAMPLE

F I G 3.1.8
81

70
0

60 9

0
0

0
50 0
Required stress
0 path ("K:= 0.55 )
0

40 0
0

0
0

30

0
0
20 0

0
0
0
0
10
0
0

.10 30 40 50

kN / m2

TRIAL TEST WITH DUMMY SAMPLE ; UNIT 2

F I G 3.1.9
82

210

"E
z
200 ?e 0 61 00-0
0
o 0 0 0 0 0 0 0 0 0 0 0 0 0 0 c,

190
200 190 180 170 160 150
k N / m2
3

50

Required stress path


( o— constant )
1
40

cNIE
30
z
-X

1..i) 20

6-
10

0
200 190 180 170 160 150
cr- kN /m2
3

TRIAL TEST WITH DUMMY SAMPLE ; UNIT 2

0-CONSTANT 0-DECREASING
1 3

FIG 3.1.10

83
1200
/
0 Load cell resting on bench A
loaded from above.
0
1000
A Load cell clamped in frame
loaded from below.
A

— 800
0

s 600
a. zt
0
O
A
— 400
i;
G.
O
-o Load cell No 26
res 200 A
0
-J ( loose f it )
0 .

A

/
0 20 40 60 80 100 120
Newtons
.17
/i7
1000
0 Tight fitting case /1
V Loose s, 11 /

• Body resting on 0
?/'
800 upper edge of case )"

/ 0
// 0
600 /,

/
/ ' o
a / o
*S. /,
O // °
400 o
, 0
a /
,0 °
200 /'
.4 7 o
71
. Load- cell No 25
• /
7

1
0 20 40 60 80 100 120
Load ( Newtons )

PERFORMANCE OF LOAD CELLS AT LOW LOADS

FIG 3.1.11
8
120 20 40 6
Load ( Newtons )


100
inches . .

....:
,„„ -
-- ..4r
/..
0.0001

...- ....-- ,--


..
50
.;'
0 I
S.
U 4 Load cell No 26
.1 ( Unit 1 )
I

.
)1
1
....:
••- 1 1
1
200 400 600 800 1000 1200 1400
Reading ( mV )


0 40 60 8 100
100
Load Newtons )

0
o 50
0

C
0
-----. Load cell No 25
.---*'..-
..--":" ( Unit 2 )
...•••"---..
.
•ti.
.
I I
I
200 400 600 800 1000 1200
Reading (mV )

DEFLECTION OF LOAD CELLS IN STRESS CONTROLLED


TRIAXIAL APPARATUS

FIG 3.1.12.
85

90

Average = 83 .5 mV per 0.01 in.

85 •
. • • . . . . • .
e • • I • • • •

• e •

80

0 1
UN IT
75
0 01 0.2 0.3
Travel in.

125
I
Average = 119.2 mV per 0.01in


— 120 ,g,. ...,, ••■■•• •
O ..... .■.■ ■
O

• • • I

• • •

C) •
0

E 115

0
UN IT 2 •
110
0 01 0.2 0.3
Travel in.

PERFORMANCE OF DISPLACEMENT TRANSDUCERS

F I G 3.1.13
86

THE APPARATUS IN OPERATION ALONGSIDE THE


" SOLARTRON " RECORDING GEAR

FiG
87

3.2 The Plane Strain Apparatus

The apparatus used for the plane strain tests has been
fully described by Atkinson (1973), who designed the apparatus
for his tests on London clay. Only a brief outline of the
apparatus will be given here, along with details of some
minor alterations made to it to make it more suitable for the
very soft clay tested here.
The apparatus takes a sample 3" high by 11/2" wide and
8" long (ie along the plane strain axis.) The sample is
completely enclosed in a rubber membrane and is held between
rigid top and bottom plattens. Rigid end plattens are brought
into contact with the ends of the sample to maintain the plane
strain condition during testing. The whole apparatus is
enclosed in a large perspex cell which is filled with water
for the application of the all round pressure. The apparatus
is shown diagrammatically in Fig. 3.2.1. The loading system
using a lower pressure chamber and Bellofram seals is identical
to that used in the hydraulic triaxial cell described earlier.
A load cell fixed to the top plate of the apparatus measures
the vertical load and one of the sample end plattens contains
a load cell for measuring the intermediate principal stress.

3.2.1 Modifications
The minor modifications made to the apparatus included
the following:
(a) New plain top plattens were made to replace the
existing top plattens with porous stones and
connecting leads. This considerably simplified
the setting up procedure and reduced the chances

of disturbance to the sample.


88

(b)The arrangement of the drainage and pore pressure


leads from the bottom platten was altered so that
water could be flushed through the bottom porous
stone to remove air when consolidated undrained
tests were performed. Also the pore pressure
transducer was repositioned so that there was a
direct connection from the transducer to the
bottom porous stone. With the original arrange-
ment it was not possible to flush water along the
bottom porous stone, and there was the danger of
air becoming trapped in the line of valves and
connections between the pore pressure transducer
and the sample. An air trap was connected into
the volume change lead as shown in Fig. 3.2.1.
(c)The end platten mountings were altered so as to
greatly reduce the frictional force acting when
the end plattens were brought up into contact
with the sample. The slotted supports were re-
designed and small roller bearings mounted on the
end plattens so that when movement of the plattens
takes place the bearings roll on their supports,
the supports themselves remaining stationary as
far as horizontal movement is concerned. The
reduction of this frictional resistance was
essential to prevent damage to the sample during
the operation of bringing the end plattens into
contact with the sample.
Also the supports (ie the roller bearings) on the
platten containing the load cell were mounted on
the thrust plate against which the end platten
89

and load cell operate, so that any small frictio-


nal force in the bearing does not influence the
load cell readings.
(d) The vertical load cell was replaced with a high
sensitivity cell to increase the accuracy of the
vertical load measurements.

3.2.2 Practical Details and Corrections


While the plane strain apparatus was designed for
testing stiff clays such as London clay, no great difficulties
have been encountered in using it for soft clay. The most
difficult part of the test operation is in the trimming of
the sample and in setting it up on the bottom platten and
getting the membrane and top plattens in place. The sample
preparation procedure is described elsewhere. Once the sample
is set up in the apparatus it is a simple matter to put the
end plattens in place and enclose the apparatus in the perspex
cylinder.
The question of the effect of friction between the
ends of the sample and the end plattens on the measured
vertical load needs some consideration. As all the tests
performed were of short duration (ie less than 8 hours)
and the maximum stresses involved between sample and platten
were small (generally less than 10 kN/m2) i
it was possible
to minimise the effect of friction by lubricating the plattens
with silicone grease. To assess the effectiveness of this a
dummy perspex sample enclosed in a rubber membrane was set
up and after lubrication with silicone grease the plattens
were brought into contact with the sample and a stress of
approximately 10 kN/m2 applied. The frictional force acting
90

on the dummy sample was measured by removing the support


from beneath the dummy and then measuring the load required
to push the dummy in the downward direction. This load was
found to be between about 0.7 kg and 1.0 kg. There was no
detectable change in this value when the sample was left
set up for 3 or 4 days before the measurement was made.
The end plattens are mounted in such a way that during
testing the plattens move upward at the same rate as a point
at mid-height of the sample (assuming uniform strain over
the height of the sample). In other words, the end plattens
are stationary with respect to the mid-height of the sample.
With this arrangement the effect of the end platten frictional
force will be to reduce slightly the stress in the sample
from the top and bottom plattens towards the centre. To
obtain the vertical stress at the centre of the sample it is
thus necessary to subtract half the frictional force as
determined above from the force measured with the load cell.
In terms of the vertical stress the correction is of the order
of 0.4 to 0.6 kN/m2. As the strength being measured was
approximately 30 kN/m2 this correction was considered sufficiently
small to be ignored.
One further minor factor which was investigated was
the influence of the dial gauge. Measurements were made of
the force required to deflect the vertical dial gauge (which
was mounted on the load cell) when the lower platten was
moving at the speed used for testing. It was found that the
force was surprisingly large and is shown in Fig. 3.2.2; the
gauge in question being denoted gauge A. However, a spare
gauge, denoted gauge B, was also checked and it was found
that the force was much less; only about one fifth of that of
91

gauge A. The reason for this is not known, though it seems


likely that the first gauge had the wrong spring mounted in
it during assembly. The second gauge was used for the tests
described in this thesis.
Load cell

Perspex cylider
End platten
with load cell
Bevelled end platten
1 '1 •
""/
I s I
‘s ;; te• tr

Volume
gauge To pressure source for cell
pressure application

Linear bearing

Air trap To pressure source for


controlled stress testing

Pressure lines for 8


drainage and back
pressure application
Motor driven piston for strain
controlled testing
Pore Pressure Pressure
transducer chamber

THE PLANE STRAIN APPARATUS


93

7
30

0 25
0
0
Gauge A e
5
0
0
0 20

0 •
4
0 •
p,0 • 15
c.,k`.'
...'
. ea-
3 :ks.•o'(*s
(..,
,S,
i., 6
e •
0
0 c.,\IN
0 0'
De's 10
U •
`o 2 mg.

0 • U
0

Gauge B 5

0 .
0 •
• •


0 2 0 4 0.6 0,8 1.0
Deflection ( inches )

FORCE EXERTED BY RUBBER SEALED OIL FILLED


DIAL GAUGES

FIG 3.2.2
94

CHAPTER 4

TRIAXIAL AND PLANE STRAIN TESTS ON SAMPLES OF LONDON CLAY

4.1 Scope and Purpose

In this section a series of several tests on London


clay are described. These were carried out while the manu-
facture of the new triaxial cells for stress path testing was
in progress, and while waiting for the arrival of samples
from the Mucking Flats trial embankment site.
In his Ph.D. thesis Atkinson (1973) has described a
series of tests carried out on samples of London clay from . a
large excavation at the Barbican redevelopment site. Triaxial
and plane strain tests made up the series and Atkinson has
given a detailed interpretation of the results in terms of
elasticity and plasticity theory. He has shown that below
certain stress levels the behaviour of the soil can be
explained using anisotropic elastic parameters. The tests
were all carried out by consolidating the soil isotropically
to an effective stress of 30 psi (roughly equal to the mean
effective stress on the soil in the ground) and then applying
a shear stress by increasing the axial load at a controlled
strain rate. Drained and undrained tests were carried out
on both vertical and horizontal samples.
At the time the writer arrived at Imperial College
Atkinson had just completed his series of tests but still had
in hand one undisturbed block sample from the Barbican site.
This block was used to carry out several further plane strain
95

and triaxial tests. One of the objects of these tests (apart


from familiarising the writer with the plane strain apparatus)
was to try and determine whether the elastic parameters
deduced from Atkinson's tests could also predict deformation
behaviour when a different stress path to failure was imposed
on the soil. Tests were therefore carried out holding
constant the vertical stress on the sample and decreasing the
lateral stress. An attempt was also made to assess the
effect of inclination by carrying out two tests on samples
inclined at 45°. A total of three pairs of tests (6 tests)
were carried out, each pair consisting of a plane strain and
a triaxial test.
The first pair of tests were on samples inclined at
45°, and a conventional type of loading was used, that is the
cell pressure was kept constant and the vertical stress
increased at a fixed strain rate.
The second pair of tests were consolidated undrained
tests on vertical samples. The vertical stress was maintained
constant and the horizontal stress (cell pressure) gradually
reduced until failure occurred.
The third pair of tests were drained tests on vertical )\%
samples with the same loading method as with the second pair.
In each case the same starting procedure as that used
by Atkinson was followed, that is the samples were consolidated
under a cell pressure of 70 psi and a back pressure of 40 psi.
The sample was set up in the apparatus and the cell pressure
applied with no drainage permitted. This cell pressure was
left on overnight and the pore pressure in the sample measured
the following morning. The back pressure was then applied
and left on overnight and the test started the following day.
96

The symbols used to identify the samples are shown


in Table 4.1 along with a summary of the results. It should
be noted that the water contents of the samples are within
the range of values measured by Atkinson although the average
is marginally lower than Atkinson's average value. There
may thus have been some very slight loss of moisture during
the storage period. The only significant harmful effect
resulting from storage of the block samples appears to have
been the increased tendency for the soil to come apart at
the fissures. This made sample preparation a rather formidable
task involving about 6 hours of very careful trimming in the
case of plane strain samples.
Each pair of tests is described in detail in the
following sections.

4.2 Samples Inclined at 45°

These were the first tests of the series and the


plane strain test was rather in the nature of a familiarisa-
tion test on the plane strain apparatus. For this reason
perhaps not too much weight should be attached to the results
of these tests. The plane strain test may be slightly
suspect on account of a leak occurringin the membrane which
was discovered when the cell pressure was applied. This
made necessary stripping down the sample and installing a
new membrane.
The tests were carried out at a strain rate of
just under 0.4% per day, the same rate as that used by
Atkinson. Filter paper drains were used. The duration of
the tests was thus about 10 to 12 days.
97

The results of the tests are shown in Figs. 4.2.1


and 4.2.2 as curves of deviator stress and pore pressure
change against strain. In these figures the same nomenclature
has been used as in Atkinson's thesis, that is
G A = axial stress (ie = a1)

aH = horizontal stress (ie = a)

The triaxial sample failed rather prematurely on a


fissure having an inclination of 53° to the horizontal, so
that curves depart rather abruptly from the curves from the
intact material as soon as movement began on the fissure.
The most significant features of the results are the
following:
(a)The slope of the initial part of the deviator
stress curve is not very different from the slope
of Atkinson's curves for both vertical and
horizontal samples. The plane strain curve has
a slope almost identical with Atkinson's vertical
sample and the triaxial sample has a slope
marginally steeper. Thus the value of the un-
drained modulus E obtained by Atkinson (which
was the same for both vertical and horizontal
samples) appears to be applicable also to samples
at intermediate inclinations.
(b)The pore pressure for both tests is close to that
which would be obtained from a perfectly elastic
isotropic material. In the triaxial test the pore
pressure is very close to the theoretical value
of one half of the deviator stress. This behaviour
appears reasonable as in a sample inclined at 45°
98

the effects of anisotropy would tend to counter-


. barance each other.
(c) A further feature of the plane strain test was
that failure did not take place along bedding
planes. The failure plane very clearly cut
across the bedding planes and was inclined to
the vertical at about 58°.

4.3 Consolidated Undrained Tests on Vertical Samples with


Constant Vertical Stress and Decreasing Lateral
Stress

In these tests the vertical stress was kept constant


and the lateral stress (cell pressure) reduced at a constant
rate until failure occurred. The rate of decrease was
2.7 psi per day which resulted in a time to failure of
13 days, roughly the same as the time to failure in the
conventional strain controlled tests. In the pla6e strain
apparatus the vertical stress stayed constant almost of its
own accord through the action of the hydraulic loading system
but in the triaxial test a dead load system was used in which
lead shot was added morning and evening to maintain the
vertical stress constant. Thus in the plane strain apparatus
the loading was smooth but in the triaxial test it was by a
succession of step increments. The lead shot which needed
to be added to the hanger system of the triaxial test was
calculated from the load cell readings at each stage.
Sufficient shot was added to bring the vertical stress to the
correct value at the time the next reading was due to be
taken.
99

Unfortunately, in the triaxial test, a fault developed


in the load cell (or the microstrain readout device) which
was not discovered until after the test and the stress path
followed was well astray from that intended. The vertical
stress rose continually throughout the test instead of staying
constant. It was possible after the test to determine with
reasonable accuracy the vertical stress throughout the test
by measuring the friction in the loading ram 0-ring and
subtracting this from the dead load. An exact record had been
kept of the weight of lead shot added throughout the test.
The deviator stress and pore pressure are plotted
against strain in Figs. 4.3.1 and 4.3.2. The slope of the
deviator stress curves is seen to be similar to those obtained
by Atkinson for the standard type of loading. In the
following Figs. 4.3.3 and 4.3.4 the stress paths are shown
in terms of both total and effective stresses. These show
that the effective stress paths followed in the writer's
tests are not significantly different from those in Atkinson's
tests. The tests are thus a reasonable demonstration of the
fact that the effective stress path in this type of test is
independent of the total stress path. This point has been
discussed in detail by various authors, for example Henkel
and Wade (1966) and Duncan and Seed (1966).
The triaxial test sample in this case appeared to be
entirely free of fissures which accounts for the high deviator
stress at failure in this case.
Perhaps the most interesting fact to emerge from
these controlled stress tests was that there was no indication
at all of impending failure until failure actually occurred.
100

No deformation or shear planes could be observed before


failure actually took place. This is in contrast to the
strain controlled tests where shear planes could be detected
several days before the peak deviator stress was reached.
In the plane strain test failure occurred sometime between
2 pm and 3 pm but was not observed. The sample was inspected
at 2 pm but no sign of approaching failure could be de-
tected. At 3 pm the sample was again inspected and failure
was found to be complete; movement of about 0.5 inches having
occurred on the shear plane through the sample. The sample
was dismantled as quickly as possible and water contents
measured on the intact soil and on soil scraped from the
failure surface. Two measurements of each were made, with the
following results:
W %

Intact Soil 27.6 and 27.4

Shear Plane 28.6 and 28.7

There was thus a measurable rise in water content


on the slip surface but this may have taken place after
failure as some time lapsed between failure and dismantling
the sample
In the triaxial test the failure was observed directly.
The sample was inspected about half an hour after the addition
of lead shot in the morning and although no deformation or
slip planes were visible the dial gauge measuring vertical
deflection could be almost seen to be moving. Records of
deflection and pore pressure were therefore commenced
immediately at one minute intervals to see if the rate of
deformation was increasing. There was in fact a steady
101

increase in the rate of movement until failure occurred 10


minutes later. The readings during these last 10 minutes
are plotted in Fig. 4.3.5. Both the vertical deformation
and pore pressure curves change slope dramatically at a point
where the failure plane first becomes apparent and the test
"runs away". The drop in pore pressure is surprising in
view of the low permeability of the soil. It would not be
expected that pore pressure changes within the sample would
be immediately recorded in the porous stone at the base.
However, the failure plane ran diagonally from top to bottom
of the sample so that it was virtually in contact with both
porous stones. This may account for the immediate response,
though it is possible that beyond the point where the failure
plane appeared the rapid drop in pore pressure was the result
of distortion of the membrane alongside the bottom porous
stone.

4.4 Drained Tests on Vertical Samples with Constant


Vertical Stress and Decreasing Lateral Stress

In these drained tests the rate of reduction in the


horizontal stress (cell pressure) was made somewhat slower
in order to maintain the same, time to failure. The rate
used was just below 2.0 psi per day. In these tests the
decrease in effective horizontal stress was the same as the
decrease in cell pressure, whereas in the undrained test
the decrease in pore pressure which accompanied the reduction
in cell pressure meant that the drop in the effective horizontal
stress was less than the drop in cell pressure. In other
words the Mohr circle at failure in the drained test was
102

smaller than the Mohr circle at falure in the undrained test.


The results of the tests are plotted in Figs. 4.4.1
to 4.4.3. In the first figure the actual stress paths are
plotted and show that these are very close to the intended
paths. In the next two figures, values of axial strain,
lateral strain and volumetric strain are plotted against
the deviator stress. The values of axial strain are very
low in these tests so that the conventional method of
plotting values against axial strain is not very suitable.
In both tests failure occurred at an axial strain close to
0.5%. In the plane strain test the value was 0.51% and in
the triaxial test it was 0.47%. It should be noted in
Figs. 4.4.2 and 4.4.3 that the sign of the volumetric and
lateral strain is opposite to that of the axial strain. That
is it say the samples are decreasing in height but increasing
in volume and diameter (or width). The axial and volumetric
strain were determined directly and the lateral strain
calculated from the relationships:

Triaxial: eV = eA + 2eH

Plane Strain: eV = eA + eH

where eV = volumetric strain

eA = axial strain

eH = lateral strain

As mentioned previously, one of the objects of these


tests was to determine whether the elastic parameters deter-
mined by Atkinson from tests involving a conventional stress
path could predict deformation when a different stress path
was followed. Using the same nomenclature as Atkinson, the

103

appropriate relationships applicable to the writer's tests


are as follows:
(a)Plane Strain Test
aA constant

a' changing

dev = -a'
j
E'r n(1
13- p ,2) - p'(1
1 +Ai')

=-1 — H
deH E' da'
1n(1 - p' 2 )
x

1
deA E H p'(1
= do'
' 3 1 + p')

(b)Triaxial Test
aA constant

a' changing

deV 31 2 dal (n - p' - np')


= -fT 1

1 n(1 - p'
deH = - daH
E'
x H 1)

V'
deA 2 -2 da'
E'

where deb, deH and deA are respectively the volumetric,


lateral and axial strain resulting from a change of do A in
the lateral stress, and

E' = Youngs Modulus in the direction of the axis of


soil symmetry (the vertical axis)
n = the anisotropy ratio = EX/E' where E; is
Youngs Modulus in any direCtion normal to the
axis of symmetry
104

p' = ratio of two orthogonal strains, both in a plane


1
normal to the axis of soil symmetry, caused by
a stress increment in one of these orthogonal
axes
p3 = ratio of the strain in any z axis to the strain
in the x axis caused by a stress increment in
the x direction.
The values of these parameters determined by Atkinson
were as follows:
E' = 1600 lb/m2
n = 0.50

1 = 0
P 3T = 0.19

Inserting these in the above equations gives

dev
(a)Plane Strain: 777 - 0.31
‘"li 1600

deH 0.50
1600

deA 0.19
dQH 1600

dev 0.62
(b)Triaxial: dal,1 1600

deH 0.50
dc
j 1600

deA 0.38
1600
105

These relationships should thus predict the initial


slope of the stress strain curves shown in Figs. 4.4.2 and
4.4.3. To see whether this is the case the initial part of
the curves are shown a second time in Fig. 4.4.4 alongside
the predicted slopes as determined above. It is immediately
apparent that there is good agreement in the plane strain
test but rather poor agreement in the triaxial test. The
initial volumetric strain is correctly predicted in both
tests, but the axial strain is less than the predicted. For
the plane strain test the value is only slightly less but
is very much less in the triaxial test. While two tests are
too few to be able to draw any general conclusions, the
results do tend to support the view that elastic theory may
be a valid means of predicting deformations in a material
such as London clay. However a great deal more information
would be needed before conclusions could be drawn regarding
its limitations and drawbacks.
It should be pointed out that the lack of agreement
in Fig. 4.4.4 cannot be attributed to the elastic parameters
determined from Atkinson's stress paths being inapplicable
to the stress paths followed here. That is to say, adjusting
the values of E, p and n will not produce better agreement.
The behaviour in the two tests is not compatible with elastic
theory. Elastic theory and in fact intuitive reasoning
suggest that the axial strain in the triaxial test will be
greater than in the plane strain as there is a greater degree
of lateral restraint in the latter. The explanation for the
difference ,tn behaviour may partly be found by reference to
tests on London clay carried out by Davies (1974). In these
106

drained tests the vertical stress was held constant and the
lateral stress decreased. It was found that in some tests
the length of the sample actually increased in the early
stages of the test, an effect which is certainly not com-
patible with elastic theory. However not all the samples
exhibited this behaviour and there appears to be a connection
between behaviour and the "stress path" followed in setting
up the sample and applying the cell and back pressure. If
after the application of the cell pressure the pore pressure
in the sample was less than the intended back pressure then
swelling occurred immediately prior to the start of the test
on connection to the back pressure. If on the other hand
the pore pressure was greater than the back pressure, then
reconsolidation occurred on application of the back pressure.
The tendency to increase in length at the start of a test
appears to be more likely in those cases where reconsolidation
has not taken place.
In the writer's tests it was found that reconsolidation
was necessary in the plane strain tests but not in the
triaxial tests due to the fact that excess water was introduced
in the plane strain test in the "flushing" procedure used to
remove air.
The peak strengths from these tests are very
variable and emphasise the difficulties of strength measure-
ments with fissured materials. The strength is particularly
variable with the triaxial samples where the possibility
of failure on a fissure is greater and the possibility of
obtaining a non fissured sample also greater. This is
emphasised in Fig. 4.4.5 which shows the Mohr's circles at
107

failure from the three triaxial tests. Sample B6 was an


"intact" sample while sample B2 had a distinct fissure
running across it at about 60° to the plane perpendicular
to the sample axis.
108

Table 4.1 Details of London Clay Samples

Density Failure Values


Sample Water us
eA Cr- 0- CT f
Content ib/ft3 grn/cm3 psi 1 3 3
psi psi

B1 26.9 124.8 2.00 - 2.46 28.3 13.9


B2 26.2 121.8 1.95 59 0.93 21.9 19.8
B3 27.7 123.0 1.97 17 1.71 33.1 9.2
B4 26.9 125.4 2.01 42 2.59 51.3 14.8
B5 26.6 121.7 1.95 26 0.51 23.2 6.3
B6 26.8 125.4 2.01 33 0.47 29.1 1.8

Key:

A increasing
B1 Plain Strain, Undrained, 45° Incl. 0- constant , CF

B2 Triaxial II II II II

H decreasing
B3 Plane Strain, Undrained, Vert. 0A constant , CJ-
B4 Triaxial SI

B5 Plane Strain , Drained, Vert. G- constant CT- decreasing


I A
B6 Triaxial f
109

40

Shear planes
become apparent

30

o 0
o o
0
Cr — Cr
0 A H
Y ° 0
0 °°

• .

20
0

du

0

ba

10

0 1 2 3 L

Strain °I.

PLANE STRAIN TEST : UNDRAINED , 45 INCLINATION

0:3— constant , increasing.


A

FIG 4.2.1
110

~o~--------r---------~--------T---------~--------~

30~--------~---------+----------r---------~--------~

Ul
a.

Failure occurred on fissure

b~

10~--~----~--------~----------r-----~~~--------~

o 2· 3 s
Strain ~.

T RIAX IAL TEST UNDRAINED, 45° INCLINATION.


"H constant OA increasing

FIG. '.2.2
111

40

Failure

V
30

a.

20
...--......'Standard type test
with o constant , o- i ncreasing.
A
( after J. Atkinson )

6 Ib

10


0 1 2 3
Strain °/.

PLANE STRAIN TEST : VERTICAL UNDRAINED.


constant o- decreasing
A H

FIG 4.3.1
112

50
/
Failure

‘'Z'
40
/
azZ' _--- ...---- ------
..---- .
.---
V
,. V
/
/

/ •
3

Standard type test


/ with offconstant cr inc-easing
A
( after J. Atkinson ) .
/

/
/

/
/
10
/
7 TRIAXIAL TEST : VERTICAL ; UNDRAINED,
/ . cr constant ( intended )
A
o- decreasing
H
/
/
b •
I
I

0 1 3
Strain ( 6/0)

FIG 4.3.2
30

20 Effective stress paths

\1/4/D .,..x x —x
0 X
/ /
o Total stress
/
I paths /
°1 /
0 /

10 (
0
Full lines : o constant
1 /
decreasing 0
3 1
0 /
Dashed lines : a— constant /
1 0 /
cr- increasing
3 /
( after J. Atkinson ) 0\
/
1 /
0
/
a /
\/
10 20 30 40
/ /
Cr 4. Cr
A
E ' '/olJ

psi
2
PLANE STRAIN TESTS VERTICAL UNDRAINED .
30~----------------~----------------r-----------------r-----------------~------1

Eff ect ive stress


paths

III
~ 201------------------~~----------------~----_+------------_4------~--~------_+--------_;

/
/
/
/
/
10~-----------------+--------------~--~--------~--------~--~~----------_+~------~
Full lines ~ ~ 0;' constant, (intended)
OJ decreasing

Dashed lines ~ ~
I
constant
0; increasing
( aft e r J. A t kin so n )

"Tl
Cl
0 10 20 30 40
+
~
•tal
..r-. "A "H psi

TRIAXIAL TESTS VERTICAL


2
UNDRAINED
...
--"

.f'
115

0.4 20

Pore pressure

18

0.3 Failure mode 16

In

14 ;
0.

4.)

0 0.2 12 IS

.■-,_,,,,,,
be

1- In

O tlo
re plane

/
10 a,
I-
0
a.

0.1
Deformation

0 1 2 3 4 5 6 7 8 9 10
T ime ( minutes )

TRIAXIAL TEST : VERTICAL ,UNDRAINED,


0- CONSTANT (INTENDED) , 3 DECREASING.

DEFORMATION AND PORE PRESSURE MEASUREMENTS AT FAILURE

FIG 4.3.5
116
20

Intended path----)\

Failure

a 10 Actual path

CNA

6z

TRIAXIAL TEST

10 20 30
o—
A + OH psi
2

21
\\
\ •

Intended path- \\

\ Failure
0
0
10 0
0

Actual path
O
.
6< •0
0

PLAIN STRAIN TEST 0


0
0 .

0 20 30
psi

ACTUAL STRESS PATHS IN DRAINED TESTS WITH


Cr CONSTANT , Cr" DECREASING
A

FIG 4.4.1
30
ci. Failure

bx
I
bct
eA 0 II
bx
e x I
H 20

TRIAXIAL TEST: 10
VERTICAL , DRAINED, o
0— constant,
A
a- <2
H decreasing.


1.0
2.4 20 1.6 1.2 08 04 0 0,4 08
e °I.
v
30

(0
a.
....

eA
A
0
o.
I
. Failure
e •
V
64
II
eH x 6z 0
<1
0
20
0

PLANE ST RAIN TEST : 10 °


VERTICAL , DRAINED,
0
a- constant.
A
decreasing. • 0
H


2.4 20 16 1.2 0.8 04 0 0.4 0,8 1.0
e v eH 'I.
e °/.
A
119

0,8

0.'
Predicted initial
slopes

0 20 30
-!
0
flo- : 0 - - cr (psi)
3 1 3

c
'iii
'-
...- 0.'
(/)

PLAIN STRAIN
TEST
0.8
• eA
<:> eV
x e
0.8 H

-- ----- ----
0"' Predicted initial ......---
slopes
'


-
0
'0 0 10 20
110-- :
313
(J - 0-- (psi)
30

C
rd
....'-
V'l
0.,

TRIAXIAL TEST

0.8~ ________________ ~ ______________ ~ __ ~ ________ ~ ________ ~

COMPARISON OF ACTUAL AND PREDICTED DEFORMATIONS IN


DRAINED TESTS

FIG 4.4.4
SO~------~--------~--------~------~--------~--------~--~----~------~

'01---------+--------+--------4--------~

III
a.
r:;'~
" e~
,~
,\ ~
30 ~,
,~
~\~
III
84
III
C1J

-~

( /)

20 ------

e(\q\'(\.....-""': (
'- c;,~1 ~ '1
"'
C1J
.J:.
u(e ",-
f\~~ ... Q
(/) / c , ...
10
82 ( failed on fissure )

"Tl
C)
0 10 20 "30 '0 50 60 70 80
l' Normal 5 tress ps i. )
~

t1'
COMPARISON OF UPPER AND LOWER STRENGTH LIMITS ~

N
0
121

CHAPTER 5

TESTS ON SOFT NORMALLY CONSOLIDATED CLAY FROM MUCKING SITE

The bulk of the testing done for this thesis is


described in this chapter. The tests were on a soft
normally consolidated alluvial clay, samples of which were
obtained from a test pit put down to a depth of about 5 m.
This test pit formed part of an investigation programme in-
volved in the planning of improvements to the flood protection
system of levees along the coasts of the Thames estuary.
The test pit was put down prior to the construction of a full
scale trial embankment. The embankment construction and
performance are described elsewhere (Pugh, 1976).

5.1 Geology of Site and General Comments on Test


Programme

5.1.1 Geology of Site


The Mucking site from which the samples were obtained
(and where the trial banks were constructed) lies on the
north (Essex) coast of the Thames estuary close to the town
of Stanford le Hope. A long stretch of the coastline here
is protected from sea flooding by a levee some 3 to 4 metres
in height. The test pit was situated some 100 m on the land
side of this levee. The ground here is flat and the surface
only a metre or so above mean sea level. The seaward side
is inundated at high tide but consists of mudflats several
hundred metres in width at low tide.
122

Maps published by the Institute of Geological Sciences


(I.G.S.) indicate that the strata along the coast here
generally consist of recent alluvial deposits nearest the sur-
face, followed in succession by Thames Gravel, London Clay
and Woolwich and Reading Beds. The samples taken and tested
were from the recent alluvial deposits.

The upper part of these alluvial deposits consists


predominantly of clay and silty clay, but deeper down may
consist of quite thick layers of fine sand interspersed with
some thin silt and clay layers. The clay layer is of
variable depth, ranging between about 5 and 15 metres and
appears to get deeper in the easterly direction. At the
Mucking site the clay layer is between about 5 and 9 metres
thick. Below it gravels are found and at the boundary
between the clay and gravel there is a thin peat layer.
Some thin peat layers also occur at shallower depths in the
clay.
This alluvial material is presumed to have been
deposited over the last 10.000 years during the gradual rise
of the sea level from its low levels during the last
glaciation period. According to Akeroyd (1972) the Thames
estuary evidence suggests that the mean sea level has risen
some 11 m since the Neolithic period, about 5000 BC and
some 1.6 to 2.6 m since the Roman period. Greensmith and
Tucker (1973) suggest that the sea level rise has been about
30 m over the last 9000 years of which 12 m has occurred over
the past 6000 years. These figures are more or less in
agreement with an earlier analysis made by Walbancke (1968).
123

Detailed information on the actual mode of deposition


does not appear to be available though it is understood that
a study of the area is at present being undertaken by the
I.G.S. It is clear however that deposition took place
under salt water conditions and that the main source of the
material deposited was the river Thames. It would be
expected therefore that the fine fraction of the material
deposited would originate mainly from London clay beds.
Evans (1965) has given an account of the sediments
found in the Wash and made suggestions regarding their mode of
deposition which may be relevant to the formation of the
Thames clays. Evans suggests that deposition occurs, mainly
in the zone limited by high and low tide levels. This zone
will generally be fairly flat so that deposition occurs over
a wide area. The velocity of the tidal currents falls off
as they move in over these intertidal flats so that coarser
material is deposited out near the low water level and finer
material near high water mark.. This mud flat bounded by
high and low water levels may have streams running across it
during low tide which may cut out quite deep channels. The
position of these channels will vary so that considerable
"reworking" of the material may occur during the deposition
period. While this mechanism is partly hypothetical it does
appear to fit in with evidence from the Mucking site. At
Mucking the silt content increases with depth and "shear
surfaces" are found in the clay which may result from minor
slips occurring possibly along the banks of former drainage
channels across the area.
124

5.1.2 General Comments on Test Programme


One or two comments are perhaps appropriate at this
stage on the laboratory test programme. Two factors which
tended to create difficulties in the testing were the
variability of the material and the very low strength and
stress levels involved. The variability of the material was
an unpredictable factor and could only be "eliminated" to
some extent by doing more than one of each type of test and
averaging the results. This was a rather time consuming and
somewhat inefficient practice as in some cases the results of
the tests would turn out to be almost identical. However, in
other cases the results would vary considerably and the
repetitive testing was justified.
The low strength and low stresses involved in normally
consolidated clays at shallow depths are illustrated in
Fig. 5.1.1. This shows a plot of in situ vertical stress and
undrained strength against depth. The vertical stresses are
given for a range of soil densities and the undrained strengths
have been calculated assuming the relationship:
Cu
= 0.11 + 0.0037PI
pi

Limiting values are given for PI of 25 and 75.


A glance at the soil mechanics literature shows that
very little testing has been carried out in this low stress
and low strength range. Many investigators involved with soft
clays have carried out their laboratory tests on samples re-
consolidated to stresses much greater than the in situ field
stress (eg Duncan and Seed,l966; Vaid and Campanella, 1974).
This appears to have been done for practical reasons, ie to
125

make experimental erros smaller in relation to the strength


of the material being tested. This procedure is questionable
as it seems likely that when tested at elevated pressures the
soil does not behave as it would at in situ stress levels.
For this reason most of the tests in this thesis have been at
stress levels close to the in situ values. This meant that
considerable care was necessary in cal'Abrating load cells
and transducers and in the actual techniques used for making
measurements. The use of internal load cells of high sensiti-
vity eliminated the ram friction factor which is often a
source of error when triaxial tests are carried out on very
soft samples.
126

kN/m2
O__----__-------5o~----__------~10~O~----__----~1~50
G.W. L •

51-....

.
E

101---~

.5 10 15 20
lb/ in 2

STRESS LEVELS AND UNDRAINED STRENGTH IN


NORMALLY CONSOLIDATED CLAY

FI G 5.1.1
127

5.2 General Description of Samples and Soil Profile at


the Test Pit

5.2.1 Method of Sampling, Sample Identification Details


and Handling Procedure,,
Tests described in this chapter were nearly all
carried out on large "block" samples taken from the test pit
put down alongside the trial embankment. They are referred
to as cylinder samples as they were taken in large cylinders
10 in (25 cm) in diameter and 12 in (30.cm) high. They had
a wall thickness of 0.25 in (6.2 mm) and a sharp cutting edge
at one end. The wall thickness was turned down to 0.17 in
(4.2 mm) for a distance of 3 in behind the cutting edge in the
hope that this would reduce sample disturbance. The cylinders
were pushed into the soil using a dead weight pressing on a
plate resting on the upper rim of the cylinder. When full the
cylinders were dug out by hand. The top and bottom surfaces
were carefully trimmed flush with the ends of the cylinders
and the ends then sealed with rubber pads and metal plates
held in place by four tie bars. Aluminium foil was placed
between the rubber pads and the ends of the samples.
A total of 36 cylinders were obtained, details of
which are given in Table 5.2.1. Generally two samples were
obtained from each depth but from one level (3.18 m - 3.48 m)
ten were obtained, to enable an extensive testing programme
to be carried out on material likely to have the same pro-
perties. The deepest samples were from just under 4.7 m
which was the maximum depth to which the test pit was taken.
The samples were stored until required in the constant
temperature (20 - 21°C) storage room at Imperial College.
128

With rubber and metal plates tightly clamped at each end


the possibility of any drying out occurring was remote.
For extruding the samples the same procedure was
adopted as is normally used for extruding 4 in dia. U4
samples. Special retaining plates, tie bans and an extruding
plate were made which could be fitted to a conventional
vertical hydraulic U4 extruder. The cylinder was thus held
in a vertical position and the sample extruded by pressing
the extruding plate up from below. It was found that the
samples generally extruded "cleanly" with a smooth outer
surface all the way round. However there were some cylinders
in which corrosion or some form of reaction between soil and
cylinder resulted in drag at the outer edges and slight
disturbance of the samples.
The principal difficulty in handling the samples arose
in cutting up the large blocks into smaller pieces suitable
for triaxial, plane strain, or consolidation testing. Most of
the testing was on 3 in x 11/2 in triaxial specimens so that the
procedure normally followed was to extrude only 4 in of the
cylinder at a time and complete the tests on this 4 in
layer before extruding a further 4 in of sample. The diffi-
culty arose in separating this 4 in layer from the remainder
of the sample. A horizontal• cut was made with a wire saw
using the rim of the cylinder as the guiding surface, but there
was a strong tendency for the soil to stick together again
as the gap behind the wire saw closed up. This could be
overcome to some extent by placing spatula blades in the saw
cut as it was formed, so that the two surfaces could not
stick back together again. After the horizontal cut was
129

made, at least one and sometimes two vertical cuts were made
in the 4 in layer so that it could then be lifted off in
sections. The sections not for immediate use were sealed and
stored until required. Sealing was carried out by placing
carefully cut pieces of sheet plastic on the soil surfaces
and then covering the sample completely with wax.
The remainder of the sample was stored in the cylinder
with the rubber pads and metal plates again clamped at the
ends. Before beginning the entruding operation the bottom
surface of the soil sample was covered with a circular piece
of sheet plastic cut to the full size of the sample and a
circular piece of plywood was placed between this plastic and
the extruding plate. This enabled the extruding plate to be
removed from the cylinder and a ring of wax poured around the
edge of the plastic sheet to ensure no loss of moisture during
further storage.

Unless otherwise stated all cutting and trimming of


the soil was carried out using a wire saw. The material was
found to be of just the right consistency for wire saw cutting
and this was used for trimming of all triaxial and plane

strain samples.

5.2.2 Description of the Soil


To start the laboratory testing programme eight
cylinders were selected more or less evenly spaced over the
full depth of the test pit. These cylinders were used for
identification tests and undrained strength measurements in
order to get a general picture of the soil properties, and
particularly their variation with depth.
130

The description of the soil obtained from careful


examination of these samples is given in Fig. 5.2.2. The soil
can generally be described as a soft grey clay over the full
depth investigated but is somewhat darker in colour near the
surface due to a higher organic content. This difference in
colour becomes more apparent after the soil has been left
exposed to the air for some time. When first extruded, the
soil from the two or three cylinders nearest the surface may
still be light grey in colour but on exposure to the air
becomes slowly darker. However the organic content is small
and after a depth of about 2 m becomes almost insignificant.
It appears to be then confined to traces of dark material
remaining in what appear to be very fine roat holes running
vertically through the soil. The root holes are larger and
more pronounced near the surface and become progressively
finer and fewer in number with depth. However they are still
present in the samples from the greatest depth.
These root holes constitute the main structural
feature of the soil apparent from visual examination. There
is no trace of any horizontal bedding despite careful examina-
tion of the soil both in its natural state and after drying
out.
A further structural feature of some importance was
the presence of very distinct "fissures" or slip surfaces in
the soil above a depth of about 2.2 m. In the writer's
view these fissures were definite slip surfaces on which
shearing involvihg substantial relative displacement had taken
place. A photograph of one of these surfaces, found in
cylinder T18 (1.38 m - 1.68 m), is shown in Fig. 5.2.1.
131

This shows fairly clearly the nature of these surfaces and the
fact that movement appears to have occurred on them. However,
these surfaces disappear completely after 2.2 m. Near the
lower part of cylinder T13 (1.98 - 2.28 m) a very distinct
shear surface was found running right across the sample at
an angle of about 15° to the horizontal and 10 cm up from
the bottom of the sample. Above this plane the soil contained
a large number of similar slip surfaces but below it no trace
of any could be found; and all samples from below this depth
were completely free of fissures. It was clear that on this
slip surface near the bottom of cylinder T13 substantial
movement had taken place as the soil above the surface was
different in colour from that below, the difference becoming
more pronounced as the soil dried out.
The cause of these shear surfaces is not known with
certainty. The most likely explanation is that they have
resulted from slipping which occurred along the banks of minor
water courses running through the soil during or shortly
after deposition.

5.2.3 Water Content and Attenberg Limits


Water content and Atter.berg limit measurements made
on each of the eight cylinde.rs are listed in Table 5.2.2 and
plotted against depth in Fig. 5.2.2. These determinations
were made on the soil on which 8 in x 4 in undrained triaxial
tests (described later) were carried out. The values are
thus not entirely representative of each cylinder as a whole.
It can be seen that there is considerable variation in the
Attenberg limits and in the relationship of the water content
132

to them. The soil is of very high plasticity for the first


2.5 m but after this there is a sharp drop to a minimum value
at about 3.5 m, below which the plasticity rises again
slightly.
It is of interest to note that in all cases the
Atterberg limits plot above the A-line on the Casagrande
plasticity chart except for those from the cylinder (T8)
closest to the surface. In this case they plot somewhat
below the A-line, which is presumably due to the considerable
organic content of this sample. With the deeper samples the
organic content appears to be insufficient to have any
significant effect on the Atterberg limits.
The variation in water content and Atterberg limits
within each cylinder was of some interest and this was in-
vestigated by making a large number of water content determina-
tions on a vertical and horizontal section through cylinder
T25. A vertical cut was made through the middle of the top
10 in of the sample and this vertical face was then divided
into a grid of 1 in squares. A second cut was made 0.5 in
from the face and parallel to it so that the specimens used
for the water content measurements were 1 in x 1 in x 0.5 in
in size. A similar grid was made on a horizontal section of
the bottom 2 in of the cylinder and water contents measured
in the same way. The results are shown in Figs. 5.2.4 and
5.2.5. They show that the variation in water content within
this one cylinder is very large. The variation is greatest
on the vertical section where the water content shows a general
increase with depth, from a minimum value of 44 % at the top
to 67 % near the bottom. On the horizontal section the maximum
133

variation is from about 59% to 74%; there apparently being


a rather sharp rise in the bottom 2 in of the cylinder where
the horizontal section was taken.
To check whether the composition also varied a second
Atterberg limit test was carried out on material taken from
near the bottom of the cylinder. This test gave the following
values:
L.L. = 62
P.L. = 27
P.I. = 35

The values from cylinder T25 in Table 5.2.2 were from


the upper 9 in of the cylinder as it was this portion which
was used for the 8" x 4" undrained test. A comparison of the
values shows that there is considerable variation in the
composition of the soil, and the changes in water content
may simply be reflections of the changes in composition.
The variation in water content in itself is therefore con-
sidered to be unlikely to indicate directly changes in other
properties such as undrained strength or compressibility.
It should be mentioned that these Atterberg limits
and the clay fractions mentioned in the next section were
all carried out without any pre-drying of the soil.

5.2.4 Clay Fraction and Activity


Clay fractions (less than 0.002 mm) were determined
using the normal hydrometer method. The values are given in
Table 5.2.2 and plotted alongside the Atterberg limits in
Fig. 5.2.2. For the top 2.5 m the clay fraction is about
60% but below this falls to a minimum of 30% at about 3.5 m.
134

In Fig. 5.2.7 the plasticity index is plotted against the


clay fraction, and shows the activity to be about 1.1.
This compares with a value of 0.95 for London clay reported
by Skempton (1953). The proportion of sand sized particles
was very small in all the cylinders, ranging between 2% and
7%.

5.2.5 Degree of Saturation


Although there was no reason to suspect that the soil
was not fully saturated a check was made during later tri-
axial testing by measuring the pore pressure response to a
series of all-round pressure increments. These were applied
to 3 in x 11/2 in samples set up in triaxial cells. Fig. 5.2.6
shows the results from three samples trimmed from cylinder
T29 (3.18 - 3.48 m). Within experimental accuracy the pore
pressure increase is seen to be identical with the magnitude
of the cell pressure increment, confirming that the soil was
fully saturated.
In the lower part of Fig. 5.2.7 values of the wet
density are plotted against water content. These have been
determined from the vertical undrained triaxial test specimens
described later and are from all the cylinders listed in
Table 5.2.2 except for the shallowest cylinder (T28)
which contained a large amount of organic material and was
probably only partially saturated. The points all lie close
to the line representing a degree of saturation of .100%
and a specific gravity of 2.7. A direct measurement of
specific gravity was not made.
135

Table 5.2.1 Details of Test Pit Cylinder


Samples

Number
Depth of Identification
m. Samples Number

0.3 — 0.6 1 T8
0.6 — 0.9 1 T7
0.78 1.08 2 T 9 and T16
1.08 —1.38 2 T 10 0 T17
1.38 —1.68 2 T18 is T19
1.68 — 1.98 2 T11 N . T 12

1.98 — 2.28 2 T13 i T14


2 . 28-2.58 2 T 20 " T 21
2,58 —2.88 2 T 22 n T23
2,88-3.18 2 T34 " T35
3 .18 — 3.48 10 T24 to T33
3.38 — 3.68 1 T42
3.48 —3.78 2 T36 and T37

3,78 — 4,08 2 T38 0 T39
4.08 —4.38 2 T40 H T41
4.38 — 4,68 1 T 43

Table 5.2.2. Properties of Cylinder Samples Used for Undrained Tests

Water Density Atterberg Clay Undrained Shear


Sample Depth Content Limits Fraction Strength k N/ m2 Sensitivity
No m gm/cm3
L.L. P.L. R I. %
Undisturbed Remoulded

T8 0 30- 0 60 673 - 138 61 67 61 50.9 62 1


T9 0.78-1.08 76.3 1,50 111 40 71 61 15.1 6.3 2.5
T 18 1.38-1.68 85.8 1.51 114 37 77 62 15.0 3.9 4
T 13 1.98-2,28 73.8 1.5 7 113 39 74 66 12.4 7.4 2
T 22 2.58-2.88 79.9 1.53 94 32 62 54 13.6 3.1 4.5
T 25 3.18-3.48 51.1 1.73 54 25 29 30 18.3 3,5 5,5
T 38 3.78-4.08 63.5 1.62 66 26 40 34 19.3 2,1 9

T43 4.38 -4.68 72.5 1.57 73 29 44 37 20.8 2.4 9


Note: The undisturbed strength has been jaken as half the failure
deviator stress in the 8 x4 tests.
The remoulded strength has been determined with the laboratory vane.

.74

=
=

= PP

p--- n
.
Nr.- '`"'• -°".. or:f.

'
167"7- ,--- .

°` - •
• ,

'11N... '-‘ - •
40 _
,, • ^
4.-_,-1"e- -
-,41 vY- --z*=01Siimpmer--L f- 404.,,

7';

PHOTOGRAPH OF SHEAR SURFACE FOUND IN CYL. T18


( DEPTH: 1. 38 1.63 m )
FIG 5.2,1
Description
Clay Fraction ( °I. )
of Soil Water Content and Atterberg Limits ( 0/. )

0 20 40 60 80 100 120 0 20 40 60 80 100


0-
Dark grey clay : containing
considerable organic matt er
o

Greyish brown clay : almost no


organic matter
o
1 Grey clay : contains fragments 1
of decomposed wood

Grey clay : contain s black o o


E zones of organic matter

2- 2
o o
Light grey clay: t race of
organic matter.
0 o
o

3 3
0 o

Grey clay : slightly silty


0 WL
virtually no organic matter
o A wP
4- 4
• w
o

5- 5
-n

GI
CYLINDER. SAMPLES t ATTER BERG LIMITS AND CLAY FRACTION


Undrained Shear Strength kN m2 Liquidity Index Sensitivity

5 10 15 20 25 0 0.2 0.4 0.6 0.8 1 2 4 10

n n
O 8 x 4 Triaxial ( V )
If .N .
3 x 1i- ' ( V ) ----
4,.. 3IIx 1-2-
iN " ( H ) /

& Plane Strain /

A +I

+e \ 4
\
-.- 0 6\
\
\
.
\\r)
IC:'
\

\+ 0

\
A i \0
\
\
E'Z'S 01 J

CYLINDER SAMPLES : UNDRAINED STRENGTH AND SENSITIVITY


140

.4 25 cm
>

43.7 45.1 50.8 50.9 47.4 50.2 47.3 503 46.8 45.4
50

42.4 .44.6 43.3 49.0 48.1 49.8 51.0 51.6 ,49. 6


50,7
45 6°

482 45.4 .52.2 52.7 .44.5 47.4 .54.6 .56.5 52.6


...........".7.50.0
.......... .
25 cm

53.1 .54.3 .54.3 .52.4 55.0 53.5 49;1 54.8 .54.3 53.8

54.8 5y.7 55.8 .56.4 .54.8 56 56.3 54.1


60.9

59.6 .585 58.5 593 601 6Q3 58.4 58.8 57.1


60 61.1

.61.3 60.8 61.6 .63.4 .63.0 63.7 6.7.1%4N\ .64.5 60. 58.4

60.4 513..8 61.8 64.0 .69.8 63.6 67.0 64.3 60.1 58.5
65

60.0 5 8.3 60.3 59.4 61.1 61.5 51.4 60.4 59.2 58.4
s

.59,3 56.6 1 61.4 .6 2.5 60.6 58,3 58.3 56.8 58.6 589
i

SAMPLE T 25 WATER CONTENTS ON VERTICAL


SECTION THROUGH CYLINDER

FIG 5.2.4
141

~---------------------------25cm

.62.2 .63.3 .
62.3

.
64.6 . .
63.8 .
65.0

•67.2 ~4.8 ~3.4

~7.8

~7.2 p8.1 ~6.4 .65.&

~9.3 .
67.5
.
65.2 68.4 66.3

66.2

.68.0 p8.3 .66.7 .66.0 .65.2 Ji6.5

~7.7
G .66.3 .65.6 .&7.0 .67.7

~6.3.

SAMPLE T 25 WATER CONTENTS ON HORIZONAL


SECTION THROUGH CYLINDER

FIG 5.2.5

142

300 •

/
/
ot/'
/
/,
//
20
/0
je / // ,
z 7/-
/
//
t//

/P?
15
U)t.n .x,'•
//
/,/
/0
10 1 2C//
//
//
// • Sample a.1
-
.11-/P x Sample a.2
0 //j/ 0 Sample a.4
a.
/
,.)G0
/ /
//
//

4/
//

0 100 200 300
Cell Pressure kN irriz

CYLINDER SAMPLE T 29 3.18 — 3.48m PORE PRESSURE


RESPONSE DURING CELL PRESSURE APPLICATION

FI G 5.2.6
100.
143
/
/
/
/
80 /
•/
/•
/
/.
60 /
/
x /
a)
-D
c ../
— 40 •
/
>, /.
U
47.
/
20 /
0
a- Activity = 1.14


20 40 60 80 100
Clay Fraction % .

19

1.8 CSC
1/4SN
.I% , \,: r,

% tb
N 1.7 —ro.

E
rn •

1.6 e


• ••• •

•:-•
*or.
. ■..... ••■
••.%,
ft "••

1.4
40 50 60 70 80 90 100
Water Content (1/0

TEST PIT CYLINDER SAMPLES: ACTIVITY AND


DENSITY / WATER CONTENT RELATIONSHIP

F I G 5.2.7
144

5.3 Undrained Measurement of Shear Strength

In this section the results of a series of more or


less conventional undrained strength tests are presented.
These tests form a "starting point" for the overall investiga-
tion of undrained strength and provide information on varia-
tions in strength with depth and with sample orientation.
The only aspect in which they are unconventional is that
they include plane strain.tests; these were carried out
partly for direct comparison with triaxial tests and partly
for the investigation of anisotropy.

5.3.1 Test Details and Procedure


On each of the cylinder samples described earlier
(listed in Table 5.2.2) a series of undrained strength
measurements has been made. The series consisted generally
of the following tests on each cylinder sample.
a. One 8 in x 4 in triaxial test
b. One plane strain test
c. Four 3 in x 11/2 in dia triaxial tests.

These tests were carried out on samples trimmed with


the normal orientation, that is with the compression axis
the same as the vertical direction in situ. In addition to
the above, with several of the cylinders tests were carried
out on 3 in x 11/2 in samples trimmed with the axis horizontal
and at 45°. Generally four of each type was carried out.
The tests were all carried out with an applied cell
pressure equal to the total vertical in situ stress. The
strain rate used was 0.2% per minute, this rate being chosen
145

to be as close as possible to normal commercial rates for


undrained tests while still permitting adequate time for
taking readings at reasonably frequent intervals. The load
cell readings at this stage of the testing programme were
being made manually using a "Budd" strain indicator. For
the earlier part of the tests readings were taken at strain
intervals of 0.2%, the intervals being increased as the tests
proceeded.
The procedure followed in extruding and handling the
samples for these tests was of necessity somewhat different
from that described earlier for tests involving only
3 in x 11/2 in samples. It was found that in this case the
most satisfactory procedure was to extrude some 9 in of
the sample and then to make a vertical cut down the centre
of this extruded portion. To assist in making the vertical _
cut one of the cylinders was specially cut into two symmetrical
halves. One of these half cylinders was placed around the
extruded sample so as to provide firm guides on which to
make the vertical cut. A horizontal cut was then made right
across the bottom of this extruded portion and one half could
then be lifted free from the rest of the block. This half
was used for preparing the 8 in x 4 in sample and some of the
3 in x 11/2 in samples. The other half was stored in one of
the half cylinders just mentioned. Plastic sheets were
placed over the soil surface and the whole was then sealed
with wax.
The plane strain sample and the remainder of the
triaxial samples were cut from this-second half. For pre-
paring the plane strain samples a special metal trimming
146

mould with spacers and "templates" was used. The templates


were thin rigid metal plates 8 in long by 3 in wide and were
used for supporting the soil during the trimming operation.
The mould and spacers enabled the soil to be cut cleanly
and accurately with the wire saw to the required dimensions
of 8 in x 3 in x 11/2 in. The most delicate part of the plane
strain sample preparation operation was in enclosing the sample
in the rubber. membrane. This was done by first folding the
membrane down below the bottom platten and then placing the
sample on the platten. Two thin plastic plates 11/2 in wide by
about 4 in high were then. placed up against the sample ends
to protect these from damage as the membrane was pulled up
over the sample. Screwing down the cover plate to clamp
the membrane in place on the top platten required care but
did not present any problems. It was not possible to remove
all the air trapped between the membrane and sample, but when
the cell was filled with water this air was forced right to
the top of the sample and remained trapped between the membrane
and the top platten. It was thus unlikely to have had any
effect on the test result. To ensure that the tests were
undrained the bottom porous stone was completely sealed off
with waterproof tape.
To investigate the sensitivity of the soil the re-
moulded strength was measured using the laboratory rane.

5.3.2 TesteResults
The results of the tests are given in full in Table
5.3.1 and shown as stress strain curves in Figs. 5.3.1 to
5.3.18. In these figures the curves are for vertical samples
unless specifically stated otherwise (ie labelled 45° or
147

horizontal). Table 5.3.2 presents a summary of the results,


giving averaged values from each of the sets of identical
3 in x 11/2 in triaxial tests. These averaged values are plotted
against depth in Fig. 5.2.3 alongside values of liquidity
index and sensitivity. It was not possible to trim samples
from the shallowest sample (T8) as this was from the upper
crust which was partially dried out and contained large numbers
of cracks. A rough estimate of its undrained strength was
obtained using a hand vane.
In attempting to evaluate these results it should be
remembered that only the bottom four cylinders were free of
fissures (shear surfaces). Cylinders T9, T18, and T13 all
contained large numbers of shear surfaces and these un-
doubtedly influenced the results to a marked degree. Sample
T13 in particular (the deepest sample in which fissures were
found) had a very high incidence of shear surfaces and this
partly accounts for the "flat" stress strain curves and lower
peak strengths from this cylinder. Two samples were from
this cylinder with existing shear surfaces running across
them inclined at about 30° to the compression axis of the
sample. The resulting stress strain curves are given in
Fig. 5.3.5. The strength of these two samples is about the
same as the lower bound of the curves for the vertical
samples shown in Fig. 5.3.3. Somewhat surprisingly the
average strength from the horizontal samples from this cylinder
is somewhat greater than that of the vertical samples. In
the writer's view this is fortuitous as the peak strengths
are erratic and depend primarily on the incidence and orienta-
tion of the shear surfaces.
148

For evaluating the results the top three cylinders


have therefore been largely ignored. The main trends which
emerge from a study of the remaining four cylinders can be
summarised as follows:
(a) There is no significant size effect (at least as far
as strength is concerned) between the 3 in x 11/2 in samples and
the 8 in x 4 in samples. With Cylinder T22 the - 3 in,x 11/2 in
samples gave the greater strength but with the other three
samples the reverse was true. The difference, in each case
was small. The 8 in x 4 in samples do however show a slightly
steeper stress strain curve than the 3 in x 11/2 in samples.
(b) It is difficult to draw any conclusion regarding the
difference between the plane strain tests and the triaxial
tests. On average the plane strain tests gave strength only
marginally lower than the triaxial tests and the slope of the
stress strain curves was very similar. The peaks of the plane
strain curves however tended to be somewhat. sharper than the
triaxials.

(c) The 3 in x lk in triaxial tests show that there is a


much wider strength variation in some samples than others.
Cylinder T25 (Fig. 5.3.9) in particular shows a large strength
variation while Cylinders T38 (Fig. 5.3.10) and T43
(Fig. 5.3.12) show only a very small variation. There is no
indication of a relationship between water content and
strength. With Cylinder T43 a total of seven vertical
3 in x 11/2 in tests was carried out giving a narrow range of
2. The water content
deviator strength from 37.2 to 40.8 kN/m
variation was from 64.2 to 78.1% but there was no consistent
correlation between strength and water content.
149

(d) The plot of strength versus depth in Fig. 5.2.3 shows


a minimum strength at about 2 m and a steady increase with depth
below this. There does not appear to be a correlation between
strength and plasticity as the sharp drop in plasticity index at
about 3 m is not reflected in the strength values.
(e) There is a consistent trend between sensitivity and
liquidity index as Fig. 5.2.3 shows. The natural water content-
moves progressively closer to the liquid limit as the depth
increases and this is reflected in the increasing sensitivity
which reaches nearly 10 after a depth of 4 metres.

(f) The tests on samples at varying orientations show a


consistent trend similar to that presented by Lo (1965). The
anisotropy data is summarised in Table 5.3.3 and in Figs. 5.3.16
to 5.3.18. The strength of the horizontal samples varies between
64% and 78% of the vertical strength. The stress strain curves
show that the initial slope is very similar although it may be
marginally lower with the horizontal samples than the vertical
samples. The vertical samples however show distinct peaks
followed by a significant fall off in strength whereas the
horizontal samples give an almost "flat" curve once a maximum
value has been reached;the 45o samples give intermediate
behaviour. It was generally found that with the vertical
samples distinct failure planes became apparent shortly after
the peak strength had been reached but that with the horizontal
samples deformation was of the bulging type without shear planes
becoming apparent.
150

5.3.3 Additional Tests to Investigate Undrained Strength


Anisotropy
The practical significance of the strength anisotropy
revealed by triaxial tests at varying orientations is somewhat
uncertain as has been pointed out by Bishop (1966). For the
analysis of practical problems it is generally necessary to
relate the strength measured in undrained compression tests to
the strength on particular planes. In his analysis of the
influence which anisotropy has on slope stability Lo (1965)
assumes that failure occurs on planes which in the field would
form part of normal plane strain slip surfaces. However, it is
equally possible that in triaxial specimens cut with the axis
horizontal, failure could occur on planes which were vertical
planes in situ (with the relative motion in the horizontal
direction), as is the case with failure on the cylindrical
surface of a conventional vane test. These two limiting
possible failure modes are shown in Fig. 5.3..19.
To investigate this aspect a further series of undrained
triaxial and plane strain tests was performed on cylinder sample
T39 (3.78 - 4.08 m). Triaxial tests were done on specimens
with inclinations varying between vertical and horizontal and
plane strain tests on vertical, 45° and horizontal samples.
With these plane strain tests the sample is restricted so that
failure can take place in one direction only, ie that re-
presented by case (a) in Fig. 5.3.19. Where material quantities
allowed, two tests were carried out on each type of specimen.
In addition two further identical plane strain tests were
carried out on samples which are here designated "end on"
(E.0.) samples, for want of a better description. These have
151

been cut so that the plane strain axis is the vertical axis
in situ and the compression axis is the horizontal direction.
These various orientations and the terminology used in de-
scribing them are illustrated in Fig. 5.3.20. With the E.O.
samples the mode of failure can only be of the type represented
by case (b) in Fig. 5.3.19.
The results of the tests are shown in full in Figs. 5.3.20
to 5.3.30 and summarised in Tables 5.3.4 and 5.3.5. It is seen
from the results that both the triaxial tests and the plane
strain tests reveal the same trend as the earlier tests ie a
progressive decrease in strength as the inclination to the
vertical axis increases. The peak strengths are plotted against
the inclination i in Fig. 5.3.28. There is considerable
experimental scatter but the general trend is unmistakable and
the plane strain values are not significantly different from the
triaxial values. The averaged ratio of horizontal to vertical
strength is 0.68.
The stress strain curves also show a similar trend to
that obtained in the earlier tests. The peak of the curve is
more pronounced in plane strain tests than in triaxial tests
and in both cases it becomes progressively flatter as the inclina-
tion to vertical increases. With the plane strain tests failure
nearly always occurred on a single distinct plane which ran
diagonally almost from corner to corner of the sample. However
with the triaxial tests the planes were less distinct and only
visible in the vertical samples. The plane strain samples were
examined carefully after failure and the modes of deformation
and failure plane angles are shown in the same figures as the
stress strain curves (Figs. 5.3.24 to 5.3.27).
152

Of particular interest are the tests on the "end on"


samples given in Fig. 5.3.27. A comparison of this figure with
Fig. 5.3.26 shows that the stress strain curves and peak
strengths from the end on samples are almost the same as those
from the horizontal samples. Thus, when the compression axis
in a test is the horizontal direction, the strength is the same
regardless of whether the axis of plane strain restraint is
the vertical direction (as in E.O. tests) or the other hori-
zontal direction (as in horizontal plane strain tests). It is
also the same when there is no constraint as the triaxial tests
show, at least within the accuracy of the present tests.
It is not really possible to relate the compressive
strengths measured in these tests to undrained strength on
particular planes. This aspect is discussed further in a later
section. The tests appear to imply that the undrained strength
anisotropy does not result from intrinsic strength differences
on different planes, but from differing pore pressure responses
dependent upon the direction of the compression axis. To
summarise this anisotropy data a polar plot in Fig. 5.3.31
shows the data from the present tests alongside that from the
three cylinders tested earlier.
The strain rate used for this series of tests was the
same as that used for the later consolidated undrained tests.
This was approximately 0.3% per hour and thus much slower than
that used for the earlier tests. This accounts for most of the
difference in strength between this cylinder and the earlier
cylinders.
In Figs. 5.3.29 and 5.3.30 the values of the inter-
mediate principal stress are plotted. These are shown as curves
of a - a and plotted on the same graphs. In each case the
3
153

values are averaged from two tests. The value of a 2 - a 3


varies in the same way asal - a3 and averages about one quarter
of a1 - a 3 except for the "end on" samples in which case it is
significantly higher.
154

Table 5.3.1 Undrained Tr i axial and Plane Strain


Tets on Cylinder Samples

Failure Values
Cyl Sample Incl in - Water Density
Deviator
No Type ation Content Strain
Ivo 3 Stress
a-m/cm
k N/ m2

T9 PS V 78.7 1.48 32.8 2.3


8in TX V 76.3 1.50 30.1 2.2
3in TX a V 80.4 1.47 38.3 3.0
n b V 78.4 1.49 35.8 2.7
" c V 82.6 1.52 24.8 2.5
" d V 80.8 1.48 32.3 2.0

T18 PS V 91.5 1.46 23.7 2.0


8in TX V 85.8 1.51 29.9 2.5
3in TX a V 86.8 1.51 26.8 3.5
" b V 84.9 1.50 27.8 5.0
" c V 90.7 1.L8 25.8 3.5
" d V 91.8 1.48 24.7 2.5

T 13 PS V 75.9 1.54 29,0 3.5


Bin TX V 73.8 1.57 24.8 4.0
3 in TX c V 73.8 1.59 28.7 4.5
" d V 79.3 1.51 20.8 8.0
it e V 76.6 1.54 22.0 8.0
" f V 79.6 1.53 24.2 3.0
.. g H 78.9 1.55 23.1 4.0
n h H 77.2 1.55 29.4 3.5
., j H 79.7 1.54 31.5 8.0
" k H 81,3 1.52 32.9 4.5 •

T22 PS V 83.0 1.49 28.7 2.0


8 in TX V 79.9 1.53 2 7.1 1.6
3 TX a V 81.1 1.5 2 31.2 2.0
- c V 82.4 1.51 2 6.4 2.0
.. e V 83,7 1.51 3 2.7 3.0
/I i V 82.4 1.52 33.4 2.7
155

Table 5.3.1 ( contd. ) Undrained Triaxial and Plane


Strain Tests on Cylinder Samples

Failure Values
Cy l Sample !flan- Water Density
Deviator
No Type ation Content Strain
Stress
gm/cm3 kN/m2

T22 3in TX f 45 85.7 1.51 26.8 4.0


.. 30.7 2.6
(contd.) h 45 82.6 1.52
H j 45 84.8 1.49 21.6 4.5
., 1 45 78.0 1.54 2 7.2 3.5
" b H 80.9 1. 52 27.1 5.0
4 d H 81.1 1. 52 22.2 1.8
" g H 84.6 1.51 25.8 4.0
'' k H 80.5 1.52 21.6 4.5

T 25 PS V 51•8 1 .6B 35.4 1.8


8in TX V 53.0 1 .70 36.5 2.0
3 in TX a V 57.9 1.63 37-2 2.7
" c V 51.3 1.70 43.5 2.5
" d V 54.7 1.67 27.6 2.7
" e V 530 1.68 32.5 2.5
" f V 62.4 1.61 37.2 2.5

T 38 PS V 62.3 1.64 29.3 2.0


8inTX V 63.5 1.62 38.6 2.5
3 in TX a V 61.7 1.63 36.0 2.5
" d V 61.0 1.63 37.7 2.7
.. g V 59.3 1.65 36.6 2.5 •
" f ,V 61.9 1.63 39.0 2.5
- " b H 68.7 1.59 24.3 4.0
' c H 58.3 1.66 22.9 7.0
- e H 72.8 1.56 26.5 4.0
" h H 71.8 1.57 24.4 3.0
156

Table 5.3.1 ( contd. ) Undrained Triaxial and Plane Strain


Tests on Cylinder Samples

Failure Values
Cy l Sample Inclin- Water Density
No Type ation Content gm Deviator
Stress Strain
0/0 /cm
kN/m2
T 43 PS V 64.0 1,60 35.9 2.5
8 inTX V 72.5 1.57 .41.5 3.0
3 inTX a V 64.2 1.61 40.8 3.5
" b V 75.2 1.55 39.4 3.0
..
c V 67.4 1.60 38.3 3.0
" d V 71.9 1.57 39.0 3.5
" h V 72.1 1.57 37.4 . 3.5
ti m V 78.1 1.54 38.7 3.5
" q V 77.7 1.54 37.2 3.5
" g 45 69.8 1.60 32.4 2.7
" k 45 68.9 1.59 31.2 3.5
" p 45 77.8 1.55 33.5 3.5
" r 45 77.4 1.55 32.E 3.5
is e H 68.3 1.60 29.8 4.5
" f H 76.3 1.56 29.5 3.5
I. H 76.3 1.56 29.3 5.0
n
11 s H 79.3 1.56 29.0 5.0
157
Table 5.3.2 Undrained Triaxial and Plane Strain Tests on
Cylinder Samples: Summary of Results

Cyl Sample
I
Inclin- Water Density
Failure Values
Deviator
N Type ati on Content gm/cm3 Stress Strain
Vo %I
kN/m2
T9 PS V 78.7 1.48 32.8 2.3
,, 8in TX V 76.3 1.50 30.1 2.2
3in TX V 80.6 1.49 32.8 2.6

T18 PS V 91.5 1.46 23.7 2.0


„ 8in TX V 85.8 1.51 29.9 2.5
n 3 in TX V 88.6 1.50 26.2 3.6

T13 PS. V 75.9 1.54 29.0 3.5


" 8in TX V 73.8 1.57 24.8 4.0
- 3in TX V 77.3 1.54 23.9 5.9
n 3 in TX H 79.3 1.54 293 5.0

T22 PS V 83.0 1.49 28.7 2.0


„ 8 in TX V. 79.9 1.53 27.1 1.6
II
3 in TX V 82.4 1.52 30.9 2.4

1 •u TX 45 82.8 1.52 26.6 3.5
„ " TX H 81.8 1.52 24.2 3.8

T25 PS V 51.8 1.68 35.4 1.8


8in TX V 53.0 1.70 36.5 2.0
" 3in TX V 55.9 1.66 35.0 2.6

T 38 PS V 62.3 1.64 29.3 2.0


" 8 in TX V 63.5 1.64 38.6 2.5
. 3 in TX V 61.0 1.63 37.3 2.5
" 3 in TX H 67.8 1.60 24.5 4.5

T 43 PS V 64.0 1.60 35.9 2.5


u 8in TX V 72.5 1.57 41.5 3.0
4 3in TX V 72.3 1.57 38.7 3.3
n " TX 45 73.5 1.57 32.6 3.4
a " TX H 75.0 1.57 29.4 3.5
.
158

Table 5.3.3 Ratio of Horizontal to Vertical Strengths


in Undrained Triaxial Tests

Ratio of
Sample Depth Vertical
Horizontal
L . L. P. I • Strength
to Vertical
m. kN/m2 Strength
T22 2.58 94 62 30.9 0.78
-2.88

T 38 3.78 66 40 37.3 0.64


—4.08
T 43 4.38 73 44 38.7 0.76
—4.68
159

Table 5.3.4 Cylinder T 39 ( 3.78 - 4.08m ) Undrained


Triaxial Tests at Varying Inclinations

Sample Water Density Failure Values


Orientation Content
gm / cm3 cr - a- Strain
°I6 1 3
kN/m2

V 64.3 1.61 29.5 2.75


V 62.2 1.63 29.4 3.0
15 63.7 1.62 29.3 2.25
15 62.7 1,65 27.1 2.0
30 65.2 1.62 25.3 3.0
45 66.1 1.62 23.8 . 2.25
45 65.8 1.62 24.9 2.5
60 59.1 1.66 23.5 3.5
75 60.4 1.65 20.4 3.0
75 65.1 1.64 20.5 3.0
H 68.1 1.62 18.3 2.25
H 72.2 1.59 22.0 3.5

Note : Sample orientation is measured by the inclination


of sample axis to vertical
160

Table 5.3.5 Cylinder T39 ( 3.78-4.08m ) Undrained


Plane Strain Tests at Varying Orientations

Failure Values
Sample Water Density
Orientation Content o- - o- Strain o- -o-
grn/c 3 1 3 2 3
kN/m2 % kN/m2

V 64,1 1.60 34.0 1.5 10.2


V 62.7 1.61 29,5 1.75 7.4
45 59t8 1.63 22.2 1.75 4.9
45 69.8 1.59 26.2 1.75. 6.8
H 67.4 1.60 21.3 2.5 5.1
H 62.9 . 1.61 20.0 2.5 4.3
E.O. 63.3 1.60 20.8 2.25 7,3
E.O. 60.0 1.62 20.4 2.0 6.9
161

40~------------~--------------~------------~--------------~

eN
E
z 20~--~~~------~--------------~--------------~------------~
~

III
III

-Q.I
'-
( /)

-'-
o
",
">Q.I 'O~~------------r---------------r---------------r-------------~
Cl
A P lane Strain

o 8 " '/.. 4 II Triaxial


" ,,, I(
• 3 x 1Z-

o 1 2 3 I.. s 6 7 8
Strain .,.

CYLINDER SAMPLE T9 0.78 - 1.08 m

UNDRAINED COMPRESSION TESTS

FIG 5.3.1
162

40~--------------~--------------~------------~~------------~

30~------------~~~~~--------r---------------r-------------~

t\l

--Z
.3::
E

IJI
.~ 20r----+---H~----r_--------------~~~------------~------------~
'-
......
(f)

'Or-~~----------r-------------~r------------

Plane Strain

Triaxial
II I /I
• 3 x 1~ "

o 2 3 5 6 7 8
Strain

CYLINDER SAMPLE T18 1. 38 - 1.G8m


UNDRAINED COMPRESSION TESTS

FIG 5.3.2
163

40~------------~--------------~--------------~--------------~

30r--------------+---------------r--------------4-------------~

cs
E
z
.:s;

20
III
III
Q)

-I-

t f)

-
I-
o
n:I
>
~ . ml-;~------------~--------------~------------~~------------~
, Ib Plane St rain
II 1/
G) 8 x 4 Triax ial
N \ /1
• 3 )( 1-2 "

o , 2 3 s 6 1 8
Strain ~o

CYLINDER SAMPLE T 13 1.98 - 2.28m

UNDRAINED COMPRESSION TESTS

FIG 5.3.3
164

2 3 4 5 6 7
Strain °I.

CYLINDER SAMPLE T 13 1.98 — 2.28m


HORIZONTAL 3"x SAMPLES ( UNDRAINED )

FIG 5.3.4
165

40

30

'E
z
_v

tn) 20

47)

O
70
Ci
0

10


0 1 2 3 4 5 6 7 8
Strain

'I

CYLINDER SAMPLE T 13 1.98 — 2.28 m


,t
3 x 2 SAMPLES CONTAINING DISTINCT SHEAR

PLANES ( UNDRAINED )

FIG 5.3.5
167

40

30

> 10

1 2 3 5 7
Strain 5/.

CYLINDER SAMPLE T 22 2.58 — 2.88 m


• N
3.x 1k SAMPLES AT 45.INCLINATION UNDRAINED

F I G 5.3.7
168

40

• 30

`4E

20

CJ

10

1 3 4 5 6 7
Strain 'I.

CYLINDER SAMPLE T 22 2.58 — 2.88 m


/. i
3 x 1k " SAMPLES : HORIZONTAL ( UNDRAINED )

FIG 5. 3 . 8
171

40

30

*4E
z

20


1 2 3 5 7
Strain 0/0

CYLINDER SAMPLE T 38 3.78 — 4.08m

3u x 127 HORIZONTAL SAMPLES ( UNDRAINED

FIG 5.3.11
173
50

40

30

c'E

In

'7" 20

10


0 1 2 3 4 5 6 7 8
Strain

CYLINDER SAMPLE T 43 4.38 — 4 .68 m

3 x 1k" VERTICAL SAMPLES ( UNDRAINED

F I G 5.3.13
174
SO

40

30

E
z

(7) 20

O
1;
Gl
C

10

0 2 3 4 5 6 7
Strain I,.

CYLINDER SAMPLE T 43 4.38 — 4.68 m


3"x 1i" SAMPLES INCLINED AT 450 UNDRA1NED

FIG 5.3.14
175
50

40

30

E
z

cu
20
1:4

0
10


0 1 2 3 5 7 8
Strain
CYLINDER SAMPLE T 43 4.38 — 4.68m
o N
3 x 1* HORIZONTAL SAMPLES ( UNDRAINED)

FIG 5.3.15
176

40

0
30
o
A

A '''
t
n
O
A

A
0 A

0
0 20 0

1.7)

10 • Vertical Samples
0 45' v•

B Horizontal "


0 1 4 7 8
Strain */.

CYLINDER SAMPLE T 43 4.38 — 4.68m


UNDRAINED TESTS ON 3"x 1ill SAMPLES AT DIFFERENT
ORIENTATIONS ( AVERAGED CURVES )

FIG 5.3.16

177

40

30

0 0
0
E
A
A

c.

10 • Vertical

0 45°

A Horizontal

1 2 3 4 5 6 7 8

Strain 0 /0

CYLINDER SAMPLE T 22 2.58 — 2.88 m


UNDRAINED TESTS ON SAMPLES ( AT DIFFERENT
ORIENTATIONS ( AVERAGED CURVES )

FIG 5.3.17
178

40

30

A A 4A- 6r A
G a .
z c.
a
r..
A

4.
s-
0

..•1
O

10 • Vertical Samples
/
A Horizontal N


0 1 2 3 4 5 6 8
Strain

CYLINDER SAMPLE T 38 3.78 — 4.08 m


UNDRAINED TESTS ON VERTICAL AND HORIZONTAL
,
3 x 17 SAMPLES (AVERAGED CURVES )

F I G 5.3.18
179

( a ) Inclined planes with strike at right angles


• to specimen axis ( dip variable )

(b) Vertical planes ( strike variable )

POSSIBLE LIMITING POSITIONS OF FAILURE PLANES


IN COMPRESSION TESTS ON CYLINDRICAL
SPECIMENS CUT WITH AXIS HORIZONTAL

FIG 5.3.19
Vertical Sample

Horizontal Sample

PLANE STRAIN TESTS

" End On" Sample

Vertical

( i = 0 )

Horizontal ( i = 90° )

/
T RIAXIAL ----(
TESTS
-t--
I
0
CYLINDER T 39 ORIENTATION OF SAMPLES USED FOR UNDRAINED TESTS

FIG 5.3.20
181

40

Vertical

30

0. 20
tri

Horizontal
6
1.
rL

10


1 2 3 4 5 6 7 8
Strain

CYLINDER T 39 3.78— 4.08m UNDRAINED TRIAXIAL TESTS AT


VARYING INCLINATIONS

FIG 5.3.21
186

. 40

64

30
64

(szE 20

z 0

10

0
ro


2 3 4 5 7 8
Strain

CYLINDER T 39 3 .78 — 4.08m UNDRAINED PLANE STRAIN TESTS


ON HORIZONTAL SAMPLES

F I G 5.3.26
187

40

30

68

20
E

tn

Cri

1 2 3 5 7 8
Strain

CYLINDER • T 39 3.78 — 4.08m UNDRAINED PLANE STRAIN TESTS


ON END ON SAMPLES

FIG 5.3.27
188

35
O

30 •• •

O
25 •
E • •
z 6 •
o
-e- •
20 -e- *---------d—
rn "End On 4" •
s _ Plane Strain
Tests
Corn pressive

15

10
Und rained

• Tr iaxial •

o Plane Strain
5

0
15 30 45 60 75 90
Inclination to Vertical ( = i )

CYLINDER T 39 3.78 — 4.08 UNDRAINED TRIAX IAL AND PLANE


STRAIN TESTS AT VARYING ORIENTATIONS

F I G 5.3.28
191

Cyl. T3 .9 • (triaxial )
01
( plane strain )
Cyl T22 4- ( triaxial )
T38 X •
T43 A

50

xo.
50 100
% OF STRENGTH WITH AXIS VERTICAL

POLAR DIAGRAM SHOWING RESULTS OF UNDRAINED


TESTS AT VARYING INCLINATIONS

FIG 5.3.31
192

5.4 Consolidation Characteristics

Only a limited investigation has been carried out on'


the consolidation characteristics of the soil. This consisted
of an oedometer and dissipation test on one of the cylinder
samples plus oedometer tests on four tube samples taken from
boreholes near the trial pit.

5.4.1 Oedometer and Dissipation Test on Cylinder Sample T24


As much of the triaxial testing was to be carried out
on samples from the 3.18 to 3.48 m level (from which 10 cylinders
were obtained) it was desirable that measurements of Cv value be
made at this depth. For this purpose an oedometer test and
dissipation test were carried out on specimens trimmed from
cylinder T24. The oedometer test was on a specimen 4 in in
diameter by 2 in high with drainage from the upper surface of
the sample only. At the base- of the sample the porous stone
was connected directly to a pressure transducer so that the
pore pressure could be measured as consolidation occurred. In
this way the Cv value could be calculated from both the rate at whic
settlement occurred and from the rate of pore pressure dissipa-
tion. The dissipation test was carried out on a specimen 4 in
in diameter by 4 in high; drainage being permitted from the top
of the sample while the pore pressure was measured at the base.
With both tests the pressure was doubled at each loading
increment. With the oedometer test the increment was added
every 24 hours but with the dissipation test it was necessary
to allow several days for consolidation to be complete before
the next increment was applied.
193

The results of these tests are given in Table 5.4.1


and in Fig. 5.4.1. The upper part of the figure shows a
conventional plot of void ratio against pressure and the lower
part a plot of the CV value against pressure. These e-log p
curves show an apparent "preconsolidation" pressure (ip) of
about 50 kN/m2 , which is not quite double the vertical in situ
effective stress of about 27 kN/m2. This sort of behaviour
appears to be normal for geologically "normally consolidated"
clays (Bjerrum, 1967; Parry, 1970), and may be attributed to
several factors. In this case the most probable cause is a
combination of secondary creep and the growth of structural
bonds of some sort between the particles. The possibility that
true overconsolidation (ie a previous stress level higher than
that at present operating) due to dessication has occurred at
some stage during the formation of the soil cannot be ruled
out.
The C values are seen to be surprisingly high at the
V
lower stresses but decrease rapidly as the stress level increases.
This may be due to the presence of the fine root holes in the
soil mentioned earlier. These would provide drainage channels
at low stresses but may be completely closed up as the stress
level rises. At the in situ stress level the C value is
V
about 2 x 10-2 2
cm /sec.

5.4.2 Consolidation Stage of Triaxial Tests


Further data on the CV value at the in situ stress
level can be obtained from the consolidation stage of samples
set up for triaxial testing. For a number of triaxial tests
(described in later sections) samples were set up and were
consolidated to the mean in situ effective stress of 20.7 kN/m2.
194

Volume change measurements were made during the consolidation


stage and from these the degree of consolidation has been cal-
culated and plotted against root time, as shown in Fig. 5.4.2.
The values plotted here are from different cylinder samples but
from the one depth (3.18 - 3.48 m). This plot indicates a value
of "t
100
" of about 100 min. which corresponds to a C
V value of
-2 cm2/sec. This is in good agreement with the value
3 x 10
from the oedometer and dissipation test.

5.4.3 Tests on Borehole Tube Samples


Oedometer tests were also carried out on four tube
samples taken from two boreholes close to the trial pit.
These tubes were thin walled Shelby type tubes with an internal
diameter of 4 in. Details of these samples with water contents
and Atterberg limits are given in Table 5.4.2. The, results are
presented in Figs. 5.4.3 to 5.4.6. It is seen that the e-log p
curves are similar to those from Cylinder T24 shown in Fig. 5.4.1.
The CV values, however, are all much smaller although there
is the same trend of decreasing values with increasing stress
level. These lower CV values may be due in part to the higher
plasticity of the soil, though this explanation is somewhat
dubious as sample S10 had a P.I. value almost the same as
Cylinder T24 (assuming the P.I. in Table 5.2.2 for Cylinder T25
is applicable to all the cylinders from this depth). Disturbance
of the soil was probably greater in the tube samples than in
the cylinders and this may also have contributed to the lower
CV values.
195

Table 5.4.1 Results of Oedometer and Dissipation


Tests on Cyl. T 24

Oedometer Test:

Pressure AL Settlement Pore Pressure


Readings Readings
L e
tgo CV
kN/m2 5 t500 W
min cm2jsec min ce/sec

0 0 1.345 — — — -
15 0.9 1.325 2.25 9.1x102 1.0 9.1 x102
30 1.5 1.309 7.84 2.5x10 2 2.75 2.0x10 2
60 4.6 1.237 30.3 6.3 x10-3 14 6.1x10 3
121 11.6 1.073 96 1.8 x 10-3 58 1.3 x103
241 16.7 0.943 130 1.2x10 3 70 9.6 x10 3

( Sample height = 2 in )

Dissipation Test: •

Pore Pressure
Pressure Readings
AV e
kN /m2 V t cv
50
min cm2/sec

0 0 -1.388
25 1.74 1.346 18 3.6 x102
50 3.48 1.305
100 8.66 1.182 145 4.2 x10-3
200 14.9 1.032 600 9.9 x10-4
400 21.3 0.890 830 6.9 x 10-4

( Sample height = 4in )


Table 5.4.2 Details of Borehole Samples on which Oedometer and
Triaxial Tests were Carried Out

Atterberg
Borehole Sample Depth Density Water Limits
No . No Content
m. gm/cm3 L.L. P.L. R.
%

52/1 S4 1.10 1.51 88.2 109 53 56

52/1 S 10 2.90 1.71 50.3 59 31 28

48 S7 2.40 1.48 89.8 102 48 54

48 S13 4,35 1.50 87.8 89 42 47



19
1.4

0
1.3

Dissipation
test
4,---"----
1.2

.•-•-'-'jr
0 Oedometer test

-o
0

10

0.9

___
8 10 20 40 60 100 600
Pr essure kN / m2
101 ci, I
I

Dissipation test :
p.p. readings •

Oedometer test :
p.p. readings o
2.
settlement readings X
( root time plot )

—2
10

• .
Oedometer___■ Dissipation
test Ar...---- test

> 2

0
-3 Y.
10
8
6

SAMPLE T 24 : COMPRESSIBILITY AND Cy MEASUREMENT

F I G. 5.4.
198

.
10 \
20
\..:
30 \\
40

\.
50 \"
0 ...\
0

C ■ ,-\

:\
1
•\
0
70 ■ •
\ .•-,
4-
0
\ \ .•
w 80
1
• •
•v`••
a)
cn
w
1 \
\
, •

90 \ \ \
\
•\ \
100 1 1
2 4 6 8 10 12 14 16 18 20
I I\ 1 1
4/ Time (min)

VOLUME CHANGE DURING CONSOLIDATION ( TO


MEAN IN SITU EFFECTIVE STRESSES )
VERSUS 4-91-1 -

FIG 5.4.2

199
2.4

2.2

2.0

o 1.8

1.6
-o
0

1.4

1.2 •
10 20 50 100 200 500
Pressure kN/ m2
10 20 50 100 200 500
5

U
3

7
"E
5

2

-4
10


7

BOREHOLE 52 / 1 SAMPLE No4 1.1 m


OEDOMETER TEST

FIG 5.4.3
200

2.5

2.0

0
rt
C
1.5
V
0

to
10 50 100 500

Pressure kN / m2

-2 10 50 100 500
10

2
"EU

—3
10

-4
10

BOREHOLE 48 SAMPLE No 7 2.4 m


OEDOMETER TEST

F I G 5.4.4
201

1.5 ■■•

1.4

-o
—6 1.0

10 20 50 100 200 500


Pressure k Ni m2
10 20 50 100 200 500
5


"E 5

3 •
>
U
2

—2
10

BOREHOLE 52 / 1 SAMPLE No 10 2.9 m

OEDOMETER TEST

F G 5.4.5
202
2.5

2,0
0
Tri

1.5
0

1.0
10 50 100 500

Pressure k N / m2

-2 10 50 100 500
10

3
U)

2
"E

-3
10

7

U 3

0
-4 •
10 •

BOREHOLE 48 SAMPLE No 13 4.35 m


OEDOMET ER TEST

FIG 5.4.6
203

5.5 Consolidated Undrained Triaxial Test Series

Before commencing a more detailed investigation of the


undrained strength a series of conventional consolidated un-
drained triaxial tests was carried out on specimens trimmed
from Cylinder T29 (3.18 - 3.48 m). The object of these tests
was primarily to determine the effective strength failure
envelope of the soil but useful information was also obtained
on the influence of consolidation on the undrained strength.

5.5.1 Procedure
The procedure followed in the tests was more or less
standard but several points should be noted. A back pressure
of 200 kN/m2 was used for all tests. Before setting up the
samples the cell pressures and back pressure were set to the
required values using the pressure transducer attached to the
cell base for pore pressure measurement during the test. The
back pressure was not essential as the soil was fully saturated
but it was a safeguard against the development of air bubbles
in the system and it increased slightly the accuracy of the pore
pressure measurements, as it elevated these measurements to
that part of the transducer range where the cal' ibration was
truly linear.
Standard rubber membranes were used but no filter
paper drains were installed to attempt to accelerate pore pressure
equalisation. There were two reasons for not using filter paper
drains. Firstly, the high CV values of the soil indicated a
high permeability in which case the effectiveness of the drains
would have been questionable. Secondly the soil strength was
204

very low so that the correction for the effect of the filter
papers would have been significant. Rather than attempt to
assess the magnitude of this effect it was decided not to use
filter paper drains. The question of the rubber membrane
correction is discussed in a later section.
The specimens were set up with the bottom porous stone
connected directly to the pore pressure transducer; the cell
pressure was then applied and the specimen left until the pore
pressure became constant. This normally did not take longer
than an hour or two. The connection to the back pressure
source was then opened and the sample left to consolidate.
At the lower stress levels the samples were allowed 24 hours
to consolidate but at the higher levels one or two extra days
were allowed. The effective stresses to which the samples were
initially consolidated (cy) are given in Table 5.5.1. The main
concern of the tests was in defining the failure envelope in
the stress range close to the in situ stresses so that most of
the samples were consolidated to pressures between 0 and about
60 kN/m2. In this range 10 samples were tested and only five
at stresses between 50 and 300 kN/m2.
The strain rate for these tests was calculated from
the CV value given in the previous section. Taking the CV value
as 3 x 10-2 cm2/sec. the time for 95% equalisation is 13 min.
(using the method given by Bishop and Henkel, 1962). The
tests were run at a rate of 0.33% per hour with readings taken
initially every 20 minutes, ie at strain intervals of just over
0.1%. Thus, provided the above CV value is acceptable this
rate was sufficiently slow to allow 95% equalisation before the
first reading was taken. It is probable that at the higher
205

consolidation pressures the Cv value becomes progressively


lower and thus for these tests there may be some error in the
initial pore pressure readings. However as failure did not
occur until a strain of at least 3% (9 hours duration) the
condition of 95% equalisation would have been fulfilled before
failure was reached.
The question of the strain rate influence is considered
in more detail in a later section.

5.5.2 Initial Pore Water Tension (Suction)


As discussed in Chapter 2 the initial pore water tension
which would exist in the samples if the sampling operation were
perfect is given by

Us = - Ko a'
V - A(1 - Ko) a' (see section 2.1.3).

In the present case the value of al', can be calculated


from the following assumptions:
Depth = 3.33 m
Soil Density = 1.52 gm/cm3
Depth of Water Table = 1 m
This gives a = 26.7 kN/m2

In tests desribed later the value of A was found to be


0.19. Assuming a value of K of 0.55 gives U = 17.0 kN/m2.
o
The actual value of U can be obtained by measuring
the pore pressure in the sample after applying the cell pressure.
The values obtained in this way in this series of tests are
given in Table 5.5.1 and show a very wide scatter above and
below the theoretical value. The reason for this scatter is
not known definitely but some possible explanations are advanced
in a later section (5.9) dealing with further measurements of the

pore water tension.


206

5.5.3 Results and Discussion


The results of this series of tests are given in Table
5.5.1 and in the following figures. Typical stress strain
curves are given in Figs. 5.5.1 and 5.5.2 and the Af values in
Fig. 5.5.3. All the Mohr's circles are plotted in Fig. 5.5.4
and typical stress paths over the full stress range are shown
in Fig. 5.5.5. The stress paths from the lower stress level are
shown in greater detail in Fig. 5.5.6. In Figs.5.5.7 and 5.5.8
the undrained compressive strength and the initial stress strain
modulus are plotted against the consolidation pressure.
An analysis of these results brings out the following
main points:
1) The pore pressure response is very dependent on stress
level as can be seen from Figs. 5.5.1 and 5.5.2. With specimen
h, consolidated to 3.4 kN/m2 the pore pressure change actually
becomes negative at failure, while with specimen i, consolidated
to 304.6 kN/m2 the pore pressure change is significantly higher
than the deviator stress. The soil thus behaves as an O.C.
clay at low stresses and as a N.C. soil at high stresses.
The Afvalues from all the tests are plotted against effective
consolidation pressure in Fig. 5.5.3. It is seen that the Af
value becomes unity at a consolidation pressure of about
100 kN/m2. The Af values above this stress level are dependent
to some extent on the length of time the specimens are left to
consolidate before testing. The time allowed in these tests
was sufficient for pore pressure dissipation but not sufficient
for secondary consolidation to have ceased. Significant creep
was still occurring when the undrained loading was commenced,
and this fact would contribute to the pore pressure response.
Hence the A values greater than unity would
f
207

probably be less if longer consolidation periods were allowed.


2) The failure envelope is clearly not linear (Figs. 5.5.4
to 5.5.6). At the lower end of the stress range the envelope
flattens out to give a significant c' intercept. Two straight
lines have been drawn giving the c' and (1)1 values indicated in
the figures. It is not suggested that the failure envelope
strictly consists of two distinct straight portions; this has
been done for convenience only and there is probably a gradual
steepening of the envelope as the pressure increases until it
finally reaches the "virgin" line passing through the origin.
The additional strength in the lower stress region
appears much greater than that which could be attributed to the
slight overconsolidation effect revealed in the oedometer
tests. It appears to be the result of some form of structural
bonding between the particles.
The trend from normally consolidated to ocervonsolidated
behaviour is clearly seen in the stress paths, in Fig. 5.5.5.
3) The plot of undrained strength versus initial consolida-
tion pressure in Fig. 5.5.7 shows a relatively flat portion
which begins to rise at 6 values somewhat in excess of the in
situ vertical stress. At the higher stress levels the points
should lie on a straight line through the origin as the soil
behaviour approaches that of normally consolidated soil. That
this is not the case with the values plotted in Fig. 5.5.7 is
probably attributable to the rather short duration of the con-
solidation stage of the tests. Had a much longer period been
allowed the undrained strength would probably have been sub-
stantially higher. The values of c' and (j)' are unlikely to have
been affected.
208

The undrained strength versus consolidation pressure


curve is similar to that given by Parry (1968) for a soft clay
in Tasmania. With Parry's tests the effect may have been
somewhat accentuated by the multistage test procedure used.
During the first stage the structural strength may be destroyed
leading to slightly lower strengths in the next stages than
might have been the case had separate samples been used.
The implication of Fig. 5.5.7 is that the undrained
strength is not greatly dependent on the initial suction in the
sample (provided this does not fall to a very low value) or on
whether the sample is reconsolidated to the field stresses.
This contrasts with Bjerrum's (1973) conclusions based on
results from Norwegian clays.
4) The initial modulus versus consolidation pressure
shows a trend similar to the undrained strength although the
trend is not well defined at higher stress levels due to some
rather erratic scatter in the values.
5) There is some evidence from these tests that the be-
haviour during the undrained loading is not greatly influenced
by the stress history prior to the start of loading. The stress
changes and the volume changes during consolidation, at least
at the lower stress levels, were very dependent on the initial
suctions which varied over a wide range. However the pattern
of behaviour shown in Fig. 5.5.6 is surprisingly consistent
despite the lack of a consistent trend in the stress changes
during consolidation.
209

5.5.4 Results from Borehole Samples


On two of the borehole tube samples used for oedometer
tests, consolidated undrained triaxial tests were carried out.
The procedure used was the same as for the above tests, but a
consolidation time of one week was allowed for each specimen.
Effective consolidation pressures of 20, 50, 100, 200 and
300 kN/m2 were used. The results are given in Table 5.5.2 and
in Figs. 5.5.9 to 5.5.13. The consolidation pressures used
were clearly not sufficient to be able to define accurately
the failure envelope in the lower stress region which is the
region of greatest interest for practical design purposes in
this case. Apart from this fact the results are in good
agreement with those above from cylinder sample T29 and the plot
of A values (Fig. 5.5.13) is almost identical to that in
f
Fig. 5.5.3.
Table 5.5.1 Cylinder T 29 ( 3.18 3.48m ) Consolidated Undrained Tria x ial Tests

Consolidation Stage Failure Values


Sample Water Density Initial Initial
No Content Suction Modulus
CT I AV CF- CT-
1 3
Strain Au A
% g m/c m3 k Nim 2 % kN/m2
kN/m2 kN/m2 % kN/m2
g 48.4 1.73 16.0 1.7 1.62 25.2 2.4 0.1 0.004 2040
h 46.9 1.75 20.0 3.4 1.51 29.4 3.7 0.7 0.002 2570
p 60.2 1.63 13.0 8.1 0.65 26.6 3.8 4.1 0.15 3350
a 48.4 1.72 30:3 10.7 1.32 36.1 2.4 3.3 0.09 3 250
q 61.1 1.6 3 14.6 12.9 1.1 29.9 2.0 7.2 0.24 3400
l 50.6 1.72 18,4 18.4 0 32.7 3.1 12.5 0.38 3 240
c 46.3 1.74 16.6 25.5 -2.0 34.1 2.7 17.9 0.52 3 240
m 53.6 1.70 17.3 31.6 -1.5 36.1 3.1 21.6 0.60 4 400
e 44.4 1.76 29.8 40.0 -0.8 43.0 3.1 27.3 0.63 6 400
f 46.4 1.74 35.8 55.8 -2.2 50.4 3.7 36.6 0.72 7100
d 48.3 1.74 8.8 70.0 -8.1 57.1 5.1 523 0.90 7600
b 48.4 1.72 26.9 100.1 -9.3 • 72.7 6.4 72.0 0.99 5560
o 48.4 1.74 15.6 152.1 -11.9 101.1 8.1 10 8.5 1.0 7 9200
j 48.3 1.73 9.9 207.2 -16.4 139.4 11.1 149.4 1.0 7 13000
i 476 1.74 13.5 304.6 -18.1 170.5 9.0 23 3.4 1.3 7 15700


= Effective Cons. Pressure
211

Table 5.5.2 Borhole Samples: Consol i dated Undrained


Triaxial Tests

Water Density Consolidation Stage Failure Values


Content IV Strain
vo o-' V
gm/cm3

Borehole 48 Sample No 13 f 4 . 35 m )

81.1 1.52 20 1.6 32.7 2.3 13.2


81.7 1.52 50 8.7 39.2 6.0 35.8
83.3 1.53 100 18.3 64.3 9.7 72.6
85 3 1.53 200 21.9 101.6 10.5 144.6
84.9 1.52 300 26.5 167.1 11.5 208.0

Borehole 52 Sample No 10 t 2.9m

46.0 1.76 20 24 33.6 6.6 11.8


48.5 1.74 50 6.4 46.5 7.1 34.8
44.6 1.76 100 7.9 69.3 7.1 71.3
87.7 1.52 200 24.5 125.2 7.6 129.9
89.2 1.52 300 28.2 169.1 9.8 212.0

Note: All pressures are in kN/m2


214

1.5
0
a.... _
.....-
.....-• -.......
....•■•
.........

.....= te r— •
1.0
...."'"
31'
e

/•

0.5 . 07.
0/
i

t
0 401

—0.5
0 100 200 300

Effective Consolidation Pressure kN / m2

CYLINDER 129 CONSOLIDATED UNDRAINED TRIAXIAL TESTS :


VALUES OF A
f

FIG 5.5.3
150

L , 3'L
C -

et,
Is \ I cc‘
<D\
100 C,

6- ci t.- 8 kN/m2
QS = 25°
I
50


0 0 50 100 i / 15Q 200 250 300
cc + CI
kN/m2
• 2
U,
CYLINDER T 29 CONSOLIDATED UNDRAINED TRIAXIAL TEST: EFFECTIVE STRESS PATHS
30

, -61 !
'A) -
1

1 '5 'c

20

E 7.
i - '0°4 1 °‘
G

--
..-
--
10 --
...
...-
...
...-
..-
.

1 Ili
0 10 20 1 / 30 40 50
Cr -I- Cr a
r 1 3 kN/rn
2
a)
CYLINDER T 29 CONSOLIDATED UNDRAINED TRIAXIAL TESTS ; EFFECTIVE STRESS PATHS
AT LOW STRESS LEVELS

100
c‘4E

80 7
VI
i
Compressive Strength

-C_. ....". .....• 0 ----...... .....'


...■ ...••• ....--
...... .../
60
C....■
CS
CU ......"'
M .....■ .00. "".....•••'-'
.....0.- ''''''.
40
• _ ....-• ."110
-- -0---•-• •-•
.- • v-
d, 20
E'
C

0 10 20 30 40 50 GO 70 80 90 100 110

Effective Consolidation • Pressure kN/m2

CYLINDER T 29 UNDRAINED STRENGTH VERSUS CONSOLIDATION PRESSURE


( 3/ X 12 TRIAXIAL TESTS )
220
• 200

150

5 50
a
a)

0 2 6 8 10 12 14
St rain

240
c = 20 k N / m2 •
1, 50 is 19
N I, 100 ii
E ii 200 ii
74
A
200 — i. 300 v +
z

L
A

at,
A

a) e
0100 A
1.,
cT)
A
a- *------X---K
f
,
A

a, 50
L
0 te
al
a_ 0
0 0----■40-"--Ch---°

f
r o
op
0
a

BOREHOLE 48 SAMPLE 13

CONSOLIDATED UNDRAINED TRIAXIAL TESTS

F I G 5.5.9
100

sa•■•
N
E

50 0

0
c
c 5.5 kN/m2
0 chi= 27°

0 50 . , 100 icn 9nr


m k N / m2
2
BOREHOLE 48 SAMPLE 13
011
CONSOLIDATED UNDRAINED TRIAXIAL TESTS

0
222

4 10 12 14
St rain

240 0—c 1 = 20 kN / rr 2. •
" 50 sa 0 •
u 100 aa x
" 200 " A
200
E
z

150
a)
cr)
c

100 ie . r-------
.'811-------.6

z
a)
7

a) zt.

50 ,.
a)
e
, , . 0
. 0

, , 0
O
0
P
0 ,..."

BOREHOLE 52 SAMPLE 10
CONSOLIDATED UNDRAINED TRIAXIAL TESTS

F I G 5.5.11
100

N
E
■ 0

50

IN •
o cl= 7.0 kN / m2
Shi =26°
----------...---.-c)

50 100 150 200


kN / m2

-n
BOREHOLE 52 SAMPLE 10
C)
CONSOLIDATED UNDRAINED T RIAXIAL TESTS
01
En

N
224

1.5

• S.
0

0
1.0

A 0
f

0.5
0

0
No 10 0
No 13 •

—0.5 1
0 100 200 300
Effective Consolidation Pressure ( kN/m2 )

BOREHOLE TUBE SAMPLES: CONSOLIDATED UNDRAINED


TRIAXIAL TESTS : VALUES .OF Af

FIG 5.5 .13


225

5.6 Influence of Stress Path on Behaviour in Consolidated


Undrained Tests

5.6.1 Test Programme and Purpose of Tests


The test programme described in this section consisted

of a series of six different types of triaxial test. These were

designated A to F as follows:
Series A: Undrained Compression.
TI
B: Isotropic consolidation, followed by un-
drained compression loading.

C: Consolidation to Ko stresses, followed by un-


drained release of deviator stress, and un-

drained compression loading.

D: Consolidation to Ko stresses, followed by


undrained compression loading.
It
E: Isotropic consolidation, followed by un-
drained extension loading.

F: Consolidation to Ko stresses, followed by


undrained extension loading.

All samples were 3 in by 11/2 in vertical samples, trimmed from

Cylinder T29.
The tests had three main objects as follows:

(a)To compare direct undrained behaviour with behaviour


after consolidation to the in situ stress level.

(b)To compare behaviour after isotropic consolidation


with behaviour after consolidation to the in situ

Ko stress state.

(c)To compare behaviour in compression loading with


that in extension loading.
226

5.6.2 Test Procedure


The procedure used for the samples consolidated isotro-

pically did not differ from that used for the consolidated un-

drained tests described earlier, except that the time allowed

for consolidation was three days rather than the one day
allowed earlier. The tests were carried out at the same con-

trolled strain rate,that is 0.33% per hour. This rate was

used for the shearing stage of all the tests described in this
section. The tests were taken to a strain of 8% so that the

test duration was about 24 hours. Readings, which were all

recorded automatically, were at 20 min. intervals initially


but at 2 hour intervals after the peak deviator stress had

been reached. The effective consolidation pressure for these

tests was taken as the average of the vertical and the hori-

zontal effective stresses in situ.


To determine this horizontal stress it was necessary to

assume a value for Ko in the field as no direct measurements of

this were possible. The value of Ko applicable at the time of


deposition of the soil may be calculated with reasonable

accuracy from the formula: Ko = 1 - sin 4' which in the present

case, assuming = 32° (Fig. 5.5.4) gives Ko = 0.47. A test


on an undisturbed sample of the present soil (described in section

5.10.1) gave a value of Ko of 0.45 once "virgin" behaviour


had been reestablished at higher stress levels. Little is known

about the way in which the lateral stress changes with time

after deposition, although it seems probable that secondary

compression and the development of structural bonds would

influence it in some way. Tentative measurements by }3jerrum

and Anderson (1972) suggest that the lateral stress may in-

crease somewhat with time. Parry (1974) believes the existing


227

K value in the soil can be determined from the apparent


o
overconsolidation ratio of the soil, and states that a Ko

value of unity would be expected in a soil exhibiting an


overconsolidation ratio of 2. Ladd (1965) presents results of

laboratory tests which suggest that a Ko value of unity would

not be reached unless the overconsolidation ratio exceeds

approximately 5, at least for soils with P.I. values in the

low to medium range. While there is a direct relationship

between Ko and the overconsolidation ratio for remoulded soils

tested in the laboratory it does not necessarily follow that


this relationship can be applied to a field situation. The

overconsolidation ratio in the laboratory is a direct result


of the stress history but a soil in the field may show an

apparent overconsolidation effect due to secondary consolidation

or to the growth of structural bonds ("hardening") between the

particles. A direct relationship between Ko and overconsolida-

tion ratio then no longer exists. For the present soil the over-

consolidation ratio is between 1.5 and 2.0, indicating the


possibility of a Ko value higher than the value of 0.45 quoted

earlier.
Actual measurements of the in situ lateral stress

have only become possible recently and the data so far is very

limited. Measurements by Bjerrum and Anderson (1972), Wroth and


Hughes (1973) and Massarsch et al (1975) give values of Ko

generally between 0.45 and 0.65, with some evidence that the
value is greater close to the surface.

In the light of the factors discussed above an assumed

value of K of 0.55 has been used for the present tests. The
o
assumed effective horizontal stress is thus 14.7 kN/m2 and the

mean effective stress is 20.7 kN/m2. This "mean" has been taken
228

as the average of the vertical and horizontal stress, not the

true mean as defined by 1/3 (ai + 2cT which would be slightly

lower (= 18.7 kN/m2). The samples consolidated isotropically


were all consolidated to this "mean" value of effective stress

rather than to the vertical effective stress, as this seemed a

more logical basis on which to make comparisons of behaviour

after Ko consolidation and isotropic consolidation. The samples


consolidated to the Ko stress condition were all consolidated
to the same assumed in situ stresses, ie a vertical effective
2
stress of 26.7 kN/m and a horizontal effective stress of 14.7

kN/m2.
The procedure used for the samples consolidated to

these in situ Ko stresses requires considerable explanation.


In theory this process should simply involve reversing the

stress release which occurred during sampling, provided that the

strain and pore pressure changes which occurred then are re-

versible. During a "perfect" sampling operation no volumetric

strain is involved, only an increase in height of the sample


and an accompanying decrease in diameter, so that returning the

sample to its in situ state should involve a decrease in

height, an increase in diameter and no change in volume. In

practice however, there are two complicating factors. The


first is that the soil does not behave elastically, so that
even a perfect sampling operation is not reversible. The second

is that the soil normally loses part of the pore water tension

which it should retain after "perfect" sampling. Because of


these two factors it is generally not possible to return the

soil to the in situ stress state without some volume change

occurring.
229

In consolidating a sample back to the field stresses

some choice has to be made as to the best sequence in which

to apply the stresses. Several possibilities exist:

(a)Undrained application of the required cell pressure


and deviator stress followed by consolidation.

(b)Application of the required cell pressure followed


by consolidation, then application of the required

deviator stress. The deviator stress could be

applied undrained followed by consolidation or

applied slowly so that drainage occurs as the stress

is applied.

(c)Gradual controlled increase in both a 1 and a3 in a


fixed ratio equal to the required Ko ratio, the rate

being chosen so that full drainage occurs as the

stresses are applied.

The first possibility is not normally an acceptable

method as the undrained application of the stresses could cause

large strains or actual failure of the sample if the initial


pore water tension (and thus effective stress) in the sample is

very low. The second method also does not appear very

desirable as it involves a reversal of the direction in which

the horizontal strains are occurring. During consolidation

under the all round pressure the sample diameter will decrease,

while during application of the deviator stress the diameter

will increase. The last method appears the most logical and
has been adopted for the present tests. Because the samples

all retained some pore water tension, it was necessary at the

start to increase ui with 6 constant until the required ratio

was reached.
230

The procedure in detail was as follows:

(a)The sample was set up and a known cell pressure of

approximately 200 kN/m2 was applied. After sufficient

time for equalisation (24 hours) the pore pressure

was measured.

(b)The cell pressure was raised until the pore pressure

reached exactly 200 kN/M2 which was the back

pressure used for these tests. The drainage connec-

tion through the volume gauge to the back pressure


source was then opened.

(c)With the cell pressure held constant the deviator

stress was increased until the value of a3/6' was


0.55, the assumed K in situ. The magnitude of the
o
deviator stress applied in this stage depended

directly on the initial value of cs (ie the pore


water tension).

(d)The deviator stress and cell pressure were then


increased at a constant rate and in a fixed ratio

so that the value of a3/a'


1 was maintained at 0.55,
until the±.equired values of the field stresses

were reached. The sample was left under these


. stresses for 3 days before testing.
Steps (c) and (d) were carried out slowly so that drainage •

could occur as the stresses were applied. The time required

was between. 1 and 2.days, depending on the initial tension.


In the remainder of this section this method of con-

solidation is referred to as "Ko" consolidation, the inverted

commas signifying that it is not true Ko consolidation as no

direct control is exercised over the lateral strain in the

sample. It would perhaps be better termed constant stress


231

ratio consolidation.
It should be noted that true Ko consolidation (zero
lateral strain) is neither a desirable nor practical method of
returning the sample to the stress state in which it existed
in the ground. If true Ko consolidation were used with the
present samples, then the value of Q3 operating when al
reached the required in situ value would vary widely depending
on the value of the initial pore water tension in the soil.
To eliminate some of the uncertainty resulting from
natural variations in the material, each type of test was
carried out at least four times. Cell pressures and pore
pressures in these tests were always measured using transducers,
2
which had an output of about 12 mV per kN/m so that the pressure
measurements were accurate to about ±0.1 kN/m2. The deviator
stress measured with the load cell was probably slightly less
accurate than this as the cal:ibration of the load cells was
not as consistent as that of the transducers.

5.6.3 Test Results


The results of. these tests are presented in Tables
5.6.1 to 5.6.3 and in Figs. 5.6.1 to 5.6.21. Where necessary
to avoid ambiguity, the results have been plotted in terms of

u A and aH where

uA = axial (or vertical) stress.

aH = horizontal (or lateral) stress.

The normal convention has been used whereby a decrease


in length is taken as a positive axial strain but a decrease
in volume is taken as a negative volumetric change. Tables
5.6.1 and 5.6.2 give the results of isotropically consolidated
232

and "Ko" consolidated samples respectively, and Table 5.6.3


shows the strains and pore pressure changes which occurred
during the undrained release of the deviator stress in the

series C tests. Figs. 5.6.1 to 5.6.3 show the stress paths

and deformations involved in consolidating the samples to the

in situ Ko stresses. Figs. 5.6.4 to 5.6.10 show the deviator


stress and pore pressure change versus strain curves from the

six sets of tests, and Figs. 5.6.11 to 5.6.16 show the corres-

ponding effective stress paths. Figs. 5.6.17 to 5.6.21 show the

stress strain curves and the stress paths after averaging the

results from each set of four tests. The results will be


discussed more or less in the order in which they are presented

in the figures.
Fig. 5.6.1 shows the stress paths followed in consoli-

dating three of the samples to the in situ "Ko" stresses. The


pore water tension values indicated by the triaxial samples in
this series were almost as variable as those found in the

earlier consolidated undrained tests (section 5.5) and thus

the starting point for the consolidation process varied over a

wide range of initial all round effective stress values.

Fig. 5.6.1 shows the stress paths at the upper and lower limits

of this range and at the centre. Sample f4 was at the centre

of the range, e4 at the lower limit and c3 at the upper limit.

Samples e4 and f4, and in fact all the samples consolidated

in this way except c3, were consolidated by first increasing

a' until the Ko stress ratio of 0.55 was reached, after which

a' and a' were raised together. The stress paths in Fig. 5.6.1
1
thus consist of a horizontal portion (to points P and E)

followed by an inclined section along the line representing


233

Ko = 0.55. With sample c3, however, this procedure was not


possible as the initial effective stress was already greater
than the required value of 63. The path followed in this case .
involved reducing cs3 and increasing aj simultaneously to
point D, followed by a small increase in al with c3 held
constant. This procedure was adopted partly for the sake of
convenience in operation of the apparatus.
Also shown in this figure is the stress path the soil
follows during an undrained release of deviator stress from the
Ko value (path AB) followed by undrained reapplication to the
same value (path BC). This "perfect" sampling operation and
undrained reapplication induces a small positive pore pressure
of about 2 kN/m2 so that the sample returns to point C rather
than point A.
Fig. 5.6.2 shows the relationship between axial and
volumetric strains occurring during consolidation of the
three samples shown in the previous figure. These are typical
of the series as a whole. The points D. E, and F correspond
to the same points in Fig. 5.6.1. Sample c3 naturally under-
goes the smallest strains, with the axial strain being sub-
stantially greater than the volumetric strain. With samples
f4 and e4 the curves consist of an initial portion (during
which a3 is constant) in which the axial strain is greater
than the volumetric strain, followed by a second portion (at
constant stress ratio) in which the volumetric strains exceed
the axial strain. Thus, during that part of the consolidation
process in which AayAal = 0.55 the sample diameter is in fact
decreasing. This may seem surprising, but it should be re-
membered that the soil is unlikely to behave as a "virgin"
234

sample until the stress level is substantially higher than the

in situ stress level and also that the "virgin" Ko value is

about 0.45 as mentioned earlier. These factors are further

discussed in a later section (5.10) dealing with drained


tests.

Fig. 5.6.3 shows the total volumetric and axial

strains from all the samples subjected to "Ko" consolidation

plotted against the initial effective stress (suction) in the


samples. Those samples with low initial suction undergo

large volumetric strains, over double the axial strains,

while those samples with initial suctions around 12 kN/m2


undergo equal volumetric and axial strains. This behaviour

suggests that volumetric strain in the samples at these

stress levels is very sensitive to stress changes in the hori-

zontal direction, ie the soil skeleton may be much less stiff


in the horizontal than the vertical direction.

The deviator stress and pore pressure curves in

Figs. 5.6.4 to 5.6.10 show clearly the need to perform more


than one test of each type in order to be sure that the re-

sults can reasonably be considered to be representative of the

soil as a whole. The scatter is probably no more than that

to be expected from any natural soil of this sort, and with


some of the tests it is surprisingly small. Series C, E, and

F in particular show good agreement between all the curves.

The undrained results in Fig. 5.6.4 show the most


scatter, although the result from specimen al should, in the

writer's view, be rejected as an anomolous result as none of

the results of the earlier undrained tests (section 5.3)

gave values as low as this. For some reason this sample had

a very low initial suction value which at least partly accounts


235

for the low undrained strength. A comparison of the results


in Fig. 5.6.4 with those in Fig. 5.6.5 shows that consolidating
the samples to the mean in situ stress has very little effect
on either the strength or the slope of the stress strain
curve, a fact which is not surprising in the light of the be-
haviour revealed by the consolidated undrained series of
tests described in section 5.5.
Fig. 5.6.6 shows the results of undrained compression
loading after the undrained release of the Ko deviator stress.
This release induces a drop in the pore pressure of about
2.3 kN/m2 so that the pore pressure at the start of the com-
pression loading is negative, since the curves are plotted
with respect to the pore pressure at the end of consolidation.
The strength of the samples is only marginally greater than
that in the previous two series of tests (A and B) although
the slope of the stress strain curve is somewhat greater.
The pore pressure change and strains occurring during the re-
lease and reapplication of the deviator stress are shown
plotted on a larger scale in Fig. 5.6.7. The values here have
been averaged from thefour tests; a summary of the values
from each test is given in Table 5.6.3. It is evident that
neither the deviator stress nor the pore pressure changes
are reversable even during unloading and reloading at this low
stress level. The parameter A during release of the deviator
stress is about 0.19, so that the pore pressure induced is
negative, and not positive as found by Bishop and Henkel (1953)
and Skempton and Sowa (1963),
The compression tests, starting from the Ko stress
state, presented in Fig. 5.6.8, show little change in strength
236

but a significant change in the slope of stress strain curve.

The initial part of the curve is clearly steeper than the

initial part in any of the previous tests.

The extension test results in Figs. 5.6.9 and 5.6.10

show that the strength is almost identical regardless of whether

the sample is initially consolidated to an isotropic or a

K stress state. The slope of the curves starting from - the


o
Ko state is, however, slightly steeper than those starting from
the isotropic state. Failure in these samples tested in
extension invariably took the form of symmetrical local

"necking" of the samples, usually near mid height. of the


sample and always within the middle one third. This "necking"

became apparent shortly after the peak values of deviator


stress had been reached. In calculating the deviator stress,

the normal area correction has been used so that the stress

strain curves are strictly only accurate up to (and perhaps

slightly beyond) the peak values of deviator stress.

It is of interest to note that it is not possible to

detect any relationship between the behaviour during shearing

and the volume change Or strains the samples underwent during


the consolidation stage. This is true of both the isotro-

pically consolidated samples and the "Ko" consolidated samples.

For example, specimens b2 and c2 had almos-tidentical initial

suction and volume change during consolidation, but had a


relatively large difference in undrained strength. Specimens

a2 and b2 had very different suctions and volume change but


their strength was almost the same. Similarly samples e4 and

f4 underwent quite different volume changes during consolidation

but their undrained strength was very similar. It appears that

any small influence which the prior strains and stress history
237

may have had on behaviour after consolidation has been ob-

scured by material variations in the soil properties.

The stress paths from the compression tests are shown

in Figs. 5.6.11 to 5.6.13. The paths in the first two figures

agree well with the behaviour shown in the earlier consolidated

undrained tests (Fig. 5.5.6). All of the stress paths are sur-

prisingly linear, particularly those from the "Kon consolidated

samples shown in Fig. 5.6.14. The stress paths suddenly change

direction only when the samples are close to failure. The


failure line shown in all these figures is that obtained from

the consolidated undrained tests presented earlier in Fig. 5.5.6.


The stress paths from the extension tests (Figs. 5.6.15

and 5.6.16) show a small, though fairly constant, degree of


curvature over their full length, with a change of direction

only when failure is reached. These stress paths show remarkably

good agreement within each set of four tests, the small scatter
apparent in Fig. 5.6.16 being due as much to small errors in the

initial consolidation stresses as to differences in behaviour


during the shearing stage.

To summarise the results Figs 5.6.17 to 5.6.21 have been


prepared. These show the results after averaging the deviator

stress and pore pressure from each set of tests. The procedure

used in obtaining these curves was to average the values of


deviator stress and pore pressure at each strain interval.

This means that the values of peak deviator stress and pore

pressure in these averaged curves are slightly less than the

true value obtained by.simply averaging the peak values from

each test, since these peaks do not occur at the same strains.

The averaged values of peak deviator stress from series


B to F are as follows:
238

Series Deviator stress kN/m2

B 31.8

C 32.7

D 33.4
E 19.7
F 19.2

The value from series A is not included because it

depends greatly on whether the questionable result from specimen


al is included or not. Excluding specimen al the average is

32.8, and including it the average is 29.7.

It is of interest to compare the undrained strengths

from this series with that obtained in the earlier undrained

tests on a different cylinder from the same depth (Cylinder T25

described in section 5.3). The average deviator stress from the

3 in x 11/2 in tests was 35.0 kN/m2. However the earlier tests


were at a strain rate of 12% per hour compared with the 0.33%

per hour used for the present tests. If an allowance is made

for this difference in strain rate on the basis of tests de-

scribed later in sectibn 5.7, then the undrained strength from


Cylinder T25 would be very close to 30 kN/m2 and thus slightly

less than the strength measured in the consolidated undrained

(series B) tests above. It appears clear, then, that the

undrained strength is not greatly affected by the four possible

initial stress conditions represented by the tests A to D.

The samples tested in compression from the Ko stress state

(series D) gave the highest value of undrained strength which

was at most 10% higher than the value obtained in direct

conventional undrained testing.


239

The slope of the stress strain curves however, is

clearly influenced by the initial stress condition. The un-

drained tests and the consolidated undrained tests show little

difference but the samples tested after undrained release of


the K
o stresses (series C) show a significantly steeper stress
strain curve than the direct consolidated undrained tests

(series B). The series c stress strain curve should in theory


be the correct stress strain curve for an undrained compression

test performed on a "perfect" sample. The samples tested in

compression directly from the Ko state show an even steeper


initial portion, although the stress strain curve as a whole
is more curved than those from the other tests.

As already mentioned the strength in extension tests

is only marginally affected by the initial stress state, though


the slope of the stress strain curve starting from the Ko state

is somewhat greater. The ratio of the undrained strength in

extension to that in compression is 0.60 for the isotropically


consolidated samples and 0.57 for the Ko consolidated samples.

A comparison of the stress strain curves from both the com-

pression and extension•tests in Fig. 5.6.19 shows that the


initial slope is very similar except for the compression tests
starting from the Ko stress state, in which case the slope is

definitely steeper.

Figs. 5.6.20 and 5.6.21 show the averaged stress paths


from all the tests except the direct undrained tests. It is

seen that there is a definite tendency for the stress paths to


converge towards a common failure point, which partly accounts

for the small strength differences between each type of test.

The failure point of the samples tested in extension suggests

that the failure envelope for extension tests is virtually the

same as that for compression tests.


Table 5.6.1 Cylinder Sample T 29 ( 3.18- 3.48m ) Triaxial Tests on
Isotropically Consolidated Samples

Consolidation Stage Failure Values


Series Sample Water Density Initial
No Content Suction 0- i Av cr-
A H-0 Strain Au
Vo gm/cm3 kN/m2 kN/m2 cyo %
kN/m2 kN/m2

A a1 65.1 1.63 1.5 20.4 2.5 0.6


bi 59.8 1.65 10.0 - - 30.1 2.5 4.6
cl 59.3 1.65 18.3 - - 34.0 1.9 9.9
d1 5 7.3 1.65 16.6 - - 34.4 1.8 8.9
a2 1.65 5.6 20.7 -3.0 30.1 3.0 13.7
inInin1.0 (.0
1:0

3 b2 1.64 20.7 20.7 -0.2 31.2 3.1 13.6


CT)
N CI

c2 1.65 1 8.3 20.7 -0.1 34.0 1.7 12.5


cn

d2 1.65 1 6.3 20.5 -0.3 34.4 2.0 12.5


co
. 0,1

!1 1.63 10.8 20.7 -1.4 3 2.2 2.5 13.4


i2 1.63 10.9 20.7 -1.5 2 8.9 3.0 13.3
b4 58.5 1.63 18.3 20.7 -0.7 -19.1 -3.8 1.2
13 67.8 1.63 6.7 20.7 -3.6 -21.2 -4.4 2.0
14 63.8 1.62 8.4 20.7 -2.5 -19.2 -3.0 2.6
j4 60.0 1.64 14.4 20.7 -2.0 -19.4 -7.1 0.8

0"-c = Effective Cons. Pressure


Table 5.6.2 Cylinder Sample T 29 ( 3.18 -3.48m ) Triaxial Tests on
" 0" Consolidated Samples

Consolidation Stage Failure Values


Series Sample Water Density Initial
No Content Suction o
gm/cm2 0.; AV AL Strain Au
kN/m 2 ' % % % kt\I /m2

b3 58.7 1.66 13.0 15.0 12.4 -0.5 0.6 33.5 2.1 6.6
C c3 60.5 1.65 16.3 14.7 12.0 -0.1 0.5 34.2 1.5 6.2
d3 60.3 1.64. 12.3 14.9 12.7 -0.8 0.7 30.6 1.5 7.6
g4 61.0 1.60 8.5 15.0 12.0 -1.2 1.0 32.5 1.7 7.5
b4 59.6 1.65 14.3 14.3 13.4 -1.8 1.0 31.3 1.9 6.5
c4 59.6 1.56 5.0 14.1 12.3 -1.4 1.0 34.2 1.2 7.5
D
d4 59.3 1.66 3.1 14.6 12.5 -2.4 1.1 33.0 1.4 7.6
e4 59.4 1.64 2.9 14.5 13.3 -2.5 1.1 3 4.8 1.6 7.7
f4 59.5 1.63 11.3 14.7 12.5 -0.8 0.7 33.8 1.6 7.0
e3 60.5 1.64 4.5 15.2 11.5 -2.6 0.8 -19.8 -6.3 -3.4
F f3 60.9 1.62 5.0 14.6 11.6 -2.5 1.0 -18.2 -5.8 -1 6
g3 62.3 1.60 6.1 14.7 11.0 -1.9 0.9 -19.8 -4.7 -2.3
h3 62.6 1.64 7.8 14.9 13.2 -1.0 0.7 -19.0 -4.2 2.5

All pressures in kN /rri
242

Table 5.6.3 Axial Strain and Pore Pressure Change


During Undrained Release of , Ko Deviator
Stress ( Series ‘`C /i Tests )

Axial
Au 4u
Sample Strain
AtCF-0-1
1 3
kN/m2

b3 —0.22 — 2.3 0.19


c3 —0.22 — 2.7 0.22
d3 —0.22 — 2.5 0.20
g4 —0.24 —1,9 0.16
c3

Initial
Suctions
15 / \

In situ
Stresses

10 /.
/
f

/ kN/m2
0 10 15 20 25 30

CYLINDER T 29 STRESS PATHS FOLLOWED IN CONSOLIDATING SAMPLES TO INSITU


K: STRESSES
1.2

.---

1.0 ___... •-------. ..-----


P.---- 0 ------ -
. ------
........-9*
f
. .
4

0.8

C
0
E".
-4+
0.6

/..° •
or
.
0.4
LL.

/: D .- . .0'14 .
0.2 I ...0

41.1 ••••••
0
1
.

t; 0
.., .....--

0.2 Oh 0 6 08 1.0 1.2 1.4 1.6 1.8 2.0 2.2

Votu metric Strain e
v

CYLINDER T 29 VOLUMETRIC AND AXIAL STRAIN DURING CONSOLIDATION TO


INSITU STRESSES

Axial Strain o

strain
cn Volumetric
, "
0
O?-

Dlateral
0-,
C
-
D
"CS
1...
0
NI

Decrease in dia. Increase in dia.


during cons. during cons.
--44--
ic +
tr
lume ..........—
1 4,----P
Vo "*.•••..., 0
.......... I
7:3 0 i
C 0 ..'........ 0 0
C 0 ---
I


0 2 4 6 3 10 12 14 16 18 20

Initial Suction kN / m2
-n
C) CYLINDER T 29 VOLUMETRIC AND AXIAL STRAIN DURING K a CONSOLIDATION
246

35
di
1

30 ,....,._„*- __„

!
b1

"E 25
N

20 , ,

15

.5 10
D

cl

0 „--0----
....,----,,--- 0--
10 —0- _ le--
i

9-°-- \d 1

bl
5

<1
a1
./ -0--
\ --- ,
0

2
02 3
Strain Vo
CYLINDER T 29 ``A” SERIES TESTS UNDRAINED

F I G 5.6.4
24

b3
_---
gL

d3

0
g4
10 d3

0/0 3
Strain

CYLINDER T20 "C uSERIES TESTS: K 0 CONSOLIDATION,


UNDRAINED RELEASE OF DEVIATOR STRESS FOLLOWED
BY COMPRESSION LOADING

FIG 5.6.8
249

15

0
N
E 0
N
z
10

O
5

O
O

N
E
z <
.------*
!..................!........0

<
0 0.2 0.3
Strain

CYLINDER T29 UNDRAINED UNLOADING AND



RELOADING FROM K o STRESSES
FIG 5.6.7
250

35 -,
f4 e4
,_-ezr-ir"-- ---
-1

,--..

------m___ ----------,
30 ----. 7.--------„,
b4

25

E 20

15

L.
10

• 0
10 d4

5 ,./ ,---D-
, /,', f4 e4 b 4

0 2
Strain
CYLINDER T 29 "D" SERIES TESTS ; K o CONSOLIDATION
FOLLOWED BY UNDRAINED COMPRESSION LOADING

FIG 5.6.9
251

5
i4 i3
0\
"---------------
0
<1 :::-1-2"4
h4 j4

—5

-5
E

-10

=-15
6 j4
,,,t h4
---..., 0--_____________ •
--4
.A------0
.-----.-../3-7-•,---■CZ.----O---
-Z,--- 72-
---°------
-20
i4 - 13

-2 5 _I

3 4 S 7 8
Strain (-ye) %

CYLINDER T 29 "E if SERIES TESTS t CONSOLIDATED' UNDRAINED


EXTENSION TESTS

FIG 5.6.9
252
f

e3

10

15

10


0
N
E
N

-15
h3 f3
/
,-...,;------...._..._04-.--..„-:•-,-, 0— t,___—•_s-4,__--e_------
------
____,-_________".3--------)
-2 0
7---o----L -- ------a---ca„ jr-----9 . __..-q
--4,-____:: __,-,•_
g3 e3
I
2 3 4 8
Strain (-ye)
"F"
CYLINDER T29 SERIES TESTS; "K: CONSOLIDATION,
FOLLOWED BY UNDRAINED EXTENSION LOADING

F I G 5 .6,10
20
..----
) `......-
-,--

et
./......
15 ../....
..,<P..,
7,2'
■•• r ,

• I ..--''''' el ,,r r
• .../.... 1
..,"'''.
---"'..

b ci1
10 ,/
,...„,--) /
../.7 ,
.. -
/7.
144 `c:il , cl
5

i .
/ P
/
.i.l.'7 0
=I.)
, .,,

20 2 5 30 35 140
Cr; +
kN / m2
2
tt
CYLANDER T 29 A SERFS TESTS STRESS PATHS
15 20 25 30
1 -r (T
3/ kN / m 2
2
CYLINDER 7 29 "H // SERIES TESTS STRESS PATHS

F 1 (3 5,6.12
2.55

20

15

10

I Cr )
O

IN

15 20 25 30
Crl- Cr kN/m2
2


CYLINDER T 29 C SERIES IESTS STRESS PATHS

F I G 5.6.13
256

20

15

N
E

10

0
15 20 25 30
0-1 +
kN/ m2
2

CYLINDER T 29 D SERIES TESTS STRESS PATHS

F I G 5.6.14
257

61

D<

—10

15
10 / 15 20
a:As + .( k N / m2
2

CYLINDER T 29 E SERIES TEST STRESS PATHS

F I G 5.6.15
25

_--- -----
5 _.....----
----
_--- ----
o -,-
ii:k a __-
Rai .----
K o --- -----
_--- .--- .--- .---

..--- ----

cv
E
*N. g3 .
z
_N..
h3
—5 16"
1 cv f3

,._, _____. f,,4 • .


.......,
—.10 EC.A.......,
-...._ -......,

Pcrib>
'
-`4re
...., -....,
-.•/?,-------
efo,-0 p,
-,
-,._, -....,
-...,,
--_,
.....„..
—15 -.....,
5 10 15 20
G7A, + kN / m 2
2
CYLINDER T29 "F" SERIES TESTS STRESS PATHS

F I G 5.6.16
2S
35

—0------
, ---------7.:lr----- •
....,-----n-,4-.-c...,-'1
:
3-''.---..-
30

25

E
N. 20

B C
Series

C il A

0 0 0

10
0

0 „------
.y-----
10

„,..„,„,,,,,,-.1'.; A ' A 0
2_,......,7... ,---":.:Q4)=1.
.

0,0 ,r1
..E.,_\
5

<1
1
0
0,,0
trai
Sn

-3

CYLINDER T29 UN DRAINED TRIAXIAL COMPRESSION


TESTS ( AVERAGED )

F f G 5.6.17

260
51

10
0 E Series : Isotropic Consoliciation
o F If . :
K0
II

C•3
E

0
-10
0

-15

'°L••••=,<A,...

,
---- 0---...... 0,
----'4:1--.::., _ ,-,.,
, --- , ''. —," ---------07-7----.7::::==./425:71s..‹
,..;------t -------------"- -
-20


0 2 3 4 5
1 Strum ( —ye )

CYLINDER T 29 UNDRAINED TM/V;(1AL EXTENSION


TESTS ( AVERAGED )

F I G 5.6.18
261
35
_.......r,-----0---
O
o
30 0,-
® Isotropic Cons.

0 $1
K0
1
.
0

25

20
.
c\IE 15

z
10

a)
5 0

0
.
L
0 0
a
>
a) —5
0

.
0

-1 0
0
0
.
-15

-2 0


-1. 5 -05 0 0. 5 1 1,5
Strain
CYLINDER T 29 UNDRAINED TRIAXIAL TESTS: COMPRESSION
AND EXTENSION STRESS STRAIN CURVES
AVERAGED VALUES )

F I G 5-5.19
262

tb" Series
20
o U Series

o "Cry Series

15

10

Stresses

Isotropic Stresses

15 20 25 30

2 k N /m2

CYLINDER T 29 : UNDRAINED T BIAXIAL TESTS : EFFECTIVE STRESS


PATHS IN COMPRESSION TESTS ( AVERAGED )

F i G 5.6.20
264

5.7 The Membrane Correction and Strain Rate Influence

5.7.1 The Membrane Correction


Because of the very low strength of the material being

tested the question of the influence of the rubber membrane


stiffness on the measured soil strength values was of some

importance. According to Bishop and Henkel (1962) the correc-

tion required with 3" x 11/2" samples using standard membranes


2
is 0.6 psi (4.1 kN/m ) at a strain of 150. This figure is based
on tests on remoulded clay and on simple theoretical considera-

tions involving the compression modulus of the membrane.


The writer's tests were almost all on undisturbed soil

and the strain at failure averaged between 2% and 2.5%. If

the correction is assumed to be proportional to strain then the


2
correction at 2.5% strain should be about 0.7 kN/m . With
2
failure stresses generally lying between 25 and 40 kN/m this
. represents a correction of 2% to 3% of the measured strength.

An attempt to provide some verification of this figure

was made by conducting a series of undrained tests on 3" x 11/2"


samples taken from cylinder sample No T27, some of the samples

being tested with membranes and some without. The samples with

membranes were tested in the conventional manner with a cell

pressure of 210 kN/m2 (30 psi). The samples without membranes


were tested using liquid parafin as the cell fluid in place

of water and the same confining pressure of 210 kN/m2. The

strain rate used was 0.2% per minute. The technique of

using liquid parafin appeared satisfactory though there was

always the danger that parafin would enter the sample through

small root holes or other discontinuities in the sample. The

results suggest that in general this did not occur. Five pairs
265

of samples were tested, the results being shown in Figs. 5.7.1

and 5.7.2. The two sets of results are seen to be almost

identical except for one sample from the second set which

appeared to fail prematurely. An examination of the sample

showed that the failure had been influenced by the presence of

a large shell in the sample, and parafin probably entered the

failure plane immediately causing the rapid decrease in

strength. If the results are taken at their face value and the

peak strengths averaged then the samples with membranes had an

averaged deviator stress of 38.2 kN/ni2 and those without membranes


2
a stressof 37.4 kN/m . If the sample in which the shell in-.
fluenced the failure is ignored then the average strength from

the samples without membranes is almost identical to that with

membranes.

These tests suggest that the membrane influence is very


small and may be less than the 0.7 kN/m2 mentioned earlier.

Because of the natural variation in the strength of the soil a

very large number of tests would be required to be able to

arrive at more reliable average strengths and accurately define


the membrane influence.. In the writer's view, the marginal

increase in accuracy did not justify the effort involved.

Further evidence on the question of membrane influence

is provided by comparing the results of the 8" x 4" undrained

tests and the 3" x 11/2" undrained tests carried out on the
cylinder samples described in section 5.3 of this thesis. On

each cylinder sample one 8" x 4" test and at least four

3" x 11/2" tests were carried out. The strengths of the 8" x 4"

samples were not significantly different from the strengths

of the 3" x 11/2" samples. If the membrane influence were

significant the measured strength of the small samples should


266

be greater than that of the large samples. Since this is not


the case it appears that the membrane influence is very small.
The above comments refer to compression tests, and in
the case of extension tests the membrane influence is likely
to be slightly greater due to the larger strains at which the
peak deviator stress was reached. This strain in the exten-
sion tests averaged between about 4.0% and 4.5%, compared with
2.5% to 3% in the compression tests. Thus the correction
should be some 50% greater with the extension tests than the
compression tests. However, this statement requires some
qualification, particularly in relation to those tests in which
the samples were consolidated to the "Ko" stress state before
shearing began. During consolidation these samples underwent
a height decrease of between 0.04 and 0.08 cms, that is a
strain of between 0.5 and 1.0%. Thus the strain at failure
with respect to the initial starting position is correspon-
dingly greater in the compression tests and less in the extension
tests than the values measured during the shearing stage of the
test. In these tests starting from the Ko position the strains
at failure with respect to the initial starting position are
almost the same in both extension and compression tests and the
membrane correction is likely to be very similar.
The overall conclusion to be drawn from the above
factors appears to be that the membrane correction to be applied
to the measured peak deviator stresses is most unlikely to be
more than 2% to 3% and in the presentation of the results in
this thesis the correction has been ignored. It should be
borne in mind however when seeking to make absolute comparisons
between laboratory strengths and field measurements for example.
267

5.7.2 Strain Rate Influence


A limited number of tests have been carried out to in-
vestigate the influence which strain rate has on the results of
undrained laboratory tests. These tests consisted of consolidated
undrained tests at four different rates of strain. The rate
used for most of the undrained tests in this thesis was 0.33%
per hour (0.0056% per min), so tests were carried out at this
rate and at rates approximately 10 and 100 times faster and at
10 times slower. The rates and the number of samples tested
at each rate are listed in Table 5.7.1. All the samples were
initially consolidated under an isotropic effective stress of
20.7 kN/m2 , the time allowed for consolidation being three
days.
The results are sumarised in Table 5.7.1, and the
deviator and pore pressure change versus strain curves are given
in Figs. 5.7.3 to 5.7.6 in order of decreasing strain rate.
There is considerable scatter in these results although the
general trend, of decreasing strength with decreasing strain
rate is clearly established. The Question arises as to whether
the pore pressure measured at the two higher strain rates can
be considered reliable. In theory at least insufficient time
occurs before the first readings are taken for adequate equalisa-
tion of pore pressure. However, a (-e parison ofiF ig.. 5.7.4
and 5.7.5 shows that the pore pressure curves are very similar
and not significantly affected by the tenfold increase in
strain rate. Even with a further tenfold increase in strain
rate (Fig. 5.7.3) the pore pressure i.A-:ves are very little
affected. The pore pressure versus Lrain relationship is
surprisingly similar. in Figs. 5.7.12 n 5.7.6.
268

In Fig. 5.7.7, averaged values of the pore pressure


parameter A are plotted against axial strain. These do not
show a particularly well defined trend although the value of A
increases somewhat with decreasing strain rate. The fact that
the A values do not vary greatly in the course of a test suggests
that there is little unequal pore pressure distribution through-
out the samples. Fig. 5.7.8 shows the averaged stress paths
from each of the four sets of tests. These indicate that the
strain rate has an effect on both the failure envelope and on
the pore pressure response to the deviator stress.
Fig. 5.7.9 shows the average strength values plotted
against time to failure, with the latter plotted on a log
scale. It is clear that the strain rate has a very marked in-
fluence on the undrained strength, and that the effect appears
more marked at the higher strain rates. Over this range of
strain rates the average strength increase per tenfold increase
in time to failure is of the order of 15%, which is much larger
than the values obtained from comparable tests on remoulded
clays. Bishop and Henkel (1962) for examples give results
from tests on remoulded Weald clay and Boston blue clay.
These show an increase in strength of only 5% per tenfold in-
crease in strain rate. Berre and Bjerrum (1973) however, give
results of undrained tests on Drammen clay which show a trend
rather similar to that in the present tests. Fig. 5.7.10
shows the average curve given by Berre and Bjerrum, on which
the results of the writer's tests are also shown. These have
been recalculated from the curve in Fig. 5.7.9. It is seen
that there is reasonable agreement between the two sets of
data although Berre and Bjerrum's curve is flattening out more
269

rapidly as the time to failure increases. Some further evidence


on the influence of strain rate in undrained tests or un-
disturbed samples is given by Conlon and Isaacs (1970) who
carried out tests on a glacial clay from Welland, Ontario.
The time to failure varied between 3 minutes and 1000 minutes,
and data from six sets of tests was obtained. There was a ;>c
lot of scatter in the results but the strength increase between
a failure time of 1000 m and 10 m varied from 18% to 44%
with an average value of 29%. Thus for each tenfold increase
in strain rate the average strength increase was just under 15%.
It is of interest to note also the results of vane
tests carried out by the Swedish Geotechnical Institute and
reported by Wiesel (1973). These were carried out at varying
depths and at a considerable range of strain rates. The results,
which showed a consistent trend over the full depth range, are
summarised in Fig. 5.7.11. The rates of testing investigated
are somewhat higher than.in the previous figure but the rate of
strength increase with strain rate is somewhat less. The strength
increase is about 12% per tenfold increase in strain rate.
The results of these tests are perhaps surprising in view of
the complicating fact that in field vane tests drainage may
be occurring as the tests proceed.
270

"---j)---)
40 •-,-0,.,...........„
..".
.•••...,...
o

<1
;.........1:ta.....m...6..........a......9.

---,a,.......,

30
N
E

II
U)
o.)
20
I

10


2 3 5
Strain

CYLINDER T 27 UN DRAINE D 3"x 1-


2 TRIAXIAL TESTS
1"
US ING STANDARD MEMBRANES

F I G 5.7.1
271

.,,
s.*'.'. ,......,..

40
..........7..
..„114,.. ...,...0

0.....°.'....ICIP,....e
•■■........

1
4
30

N
E F.
I/

Large she l
on

failure plane

0
ul
0

-(1/1 20

1
t

0
6
.+

10

0 1 2 4 6

Strain

a
CYLINDER 127 UNDRAINED 3"x 1 T RIAXIAL TESTS USING
L I QUID PARAFFIN AS CELL FLUID AND NO MEMBRANE

FIG 5.7.2
272

'Table 5.7.1 Strain Rate Effect in Consolidated


Und r (lined Triaxial Tests

Number Strain Time Average Average


Series of Rate to Deviator Failure
Tests Failure Stress Strain
% /min
m. k N / m2

1 4 0.79 1 3.8 42, 8 3,1

2 3 0.074 41 34.8 4.2

3 4 0.0059 513 31, 4 2.6

4 3 0.00048 6160 26.5 3.0


Note All samples consolidated to 20.7 kN / m2.


Time to failure calculated assuming strain
to failure of 3 percent.
273

50

40
t)
E
N
z
30

10

6
Strain


1 2 3 4 5 6
20
N
E
-4^............ ....4.-e, .rr.....4
z
10

t
0

CYLINDER T 27 CONSOLIDATED UNDRAINED TRIAXIAL


TESTS ( STRAIN RATE = 0,78 % / MIN

F I G 5.7,3
274

35

B.,,.

__-.-.------ -- '-'-'-

3Q

25

N
E 20

15
7

L4'
10

LT
a
5 --n--.----e,.....—

a„..--IT-
-------4

t-
0 A
1
10
N
E


2 3
Strain
CYLINDER T 27 CONSO I_ I DATED UNDRAINED T RIA)(IAL
TEST ( STRAIN RATE = 0.074 % / MIN )

S= IG 5.7,4
275
3

n.
,.....--- ..--4Z,--..--,,—• '44-7"---
11 -------t:—.
,,,,_
--,--,,,,. , 4 ‘,.,
0
------` ^ '.

30
.....-...."—.

tc

25

If

20 1 9-

15

10

5
..7=—.
..—
--.---_.
--- ----

0
10

5 i


1 2 3
Strain

CYLINDER T27 CONSOLIDATED UNDRAINED TRIAXIAL


TESTS ( STRAIN RATE = 0.0059 % / MINI.

FIG 5.7,5
276

30

25 3

'----------------Z,

(L'----------O--_

20

N
E
15
z

10

,
,__-*----°--------"

-e --

10

N
E


0 1 2 6
Strain

CYLINDER T 27 CONSOLIDATED UNDRAINED


TRIAXIAL TESTS ( STRAIN RATE = 0,00048 / MIN.

I G 5.7.6
277

1.0
o it = 3.8 min.
x ' = 41 ii
0.9
. ' = 513 .,

11, I. = 6160 11
0.3

0.7

0.6

41 A
A GC
0.5 A-----A-2\ A

A NX
x o

0.4 €° 1-,_0X
37
-(31
X,
--, 40x- -xa x X -----------x
0
0
Off"

0.3

0.2

0.1


2 3
St rain

ST RAIN RATE INFLUENCE : VALUES OF A FROM


TESTS AT DIFFERENT SPEEDS

FIG 5.7.7
L7 8

20
/..
...---"
.---'
..----

....". ..■ .../


x
--
1
.. .-• "X.-^* ''....
...---.'
15 A
.■ .■
../". ../..

N ,...
/ ../..
j
E
N
Is
\
10

A
\
N

0 t f -- 3,8 min

x II = 41 H

® H = 513 II

t:h II 616 I0 "


0
17 20 25 30 32

kN /rn2
2

STRAIN RATE INFLUENCE : ST R ESS PATHS FROM


TESTS AT DIFFERENT SPEEDS

F I G 5.7.8
45

4
E

z
35

L.
.Dfl

L
0

a)

20
1 10 100 1000
Time to faiture ( minutes )
STRAIN RATE INFLUENCE : PEAK STRENGTHS FROM CONSOLIDATED UNDRAINED TRIAXIAL
TESTS ( AVERAGED VALUES )
1.6
0 Writer s Tests
I
- Average curve for N. G. L
1.4 tests on Drammen clay.
..... ( after Bjerrum & Berre 1973 )
~ ,
1.2 -
c
E
" i'o.,
r-.. 1"-1I J. I I I I I, I! I

0
~
~
~~
II
~
1M
0 --
-0-
4+-

I
~ .........
........ .......
~ "'(.~-..
i
I
I
I
1.. 0 i-
-
I
I ------.... -----,
tr-
t

0
(Y')

,
.1

---- _1- --1)---..


-. I
I Ii ~.

I
'
I I
O.S 1-
l~
0 iI
! I I I Il~
~ - I
i ' !
i
i I I II
I I
I I I
III
I I I I , II I
I
O r
.0 1I ! ! I ! I
i
I I , II
1 10 100 1000 10000
T',me to failure minutes )

STRA!N RATE INFLUENCE: C0 tv1 PAR ISO N WITH NORWEGIAN GEOTECH NICAl
Ui
INSTITUTE TESTS
1.6

after Wiese( 1973 )

1.2

to

0.8

-n 0.01 01 1 10 100
Time to failure I tf ) mins.

INFLUENCE OF RATE OF ROTATION ON IN SITU VANE TEST RESULTS : TESTS BY


SWEDISH GEOT ECHNICAL INSTITUTE
282

5.8 Further Investigation of Anisotropy

5.8.1 Test Programme and Purpose of Tests


In an earlier section (5.3) it has been shown that the
soil is distinctly anisotropic with respect to the undrained
strength when this is measured by carrying out compressive
tests on samples with varying inclinations. There was a steady
reduction in undrained strength as the inclination of the sample
axis to the vertical increased from zero to 90°. It was also
shown that the trend was the same regardless of whether the
specimens were tested in triaxial compression or plane strain
compression. Pore pressure measurements were not made so
that the cause of the anisotropy was not established. Aniso-
tropy of undrained strength may arise either from variations in
the effective stress failure envelope (ie in c' and/or (1)1 )
or from variations in the pore pressure response and thus in
the effective stresses in the sample as loading proceeds.
In this action a series of tests are described which
were carried out in an attempt to ascertain which of the above
two possibilities was *responsible for the undrained strength
anisotropy revealed earlier. At the same time it was hoped
to define more precisely differences in behaviour between
triaxial and plane strain loading. The tests in section 5.3
sugaested that there was little difference in either strength
or deformationmodulus between triaxial and plane strain
loading, although the plane strain tests showed a sharper peak.
This series of tests was carried out on samples trimmed
from cylinders T28 and T27, both from a depth of 3.18 to 3.48 m.
The tests, which were all isotropically consolidated, undrained
tests, consisted of the following:
283

(a)Vertical and horizontal triaxial and plane strain


compression tests.
(b)Vertical and horizontal triaxial extension tests.
(c)Triaxial extension and compression tests on inclined
samples with i = 15, 30, 45, 60 and 75°.

5.8.2 Test Procedure


All the samples were initially consolidated to an iso-
tropic stress of 20.7 kN/m2 , the mean in situ stress. The time
allowed for consolidation was 3 days and the strain rate during
shearing was 0.33% per hour as in previous tests. The pro-
cedure for the triaxial tests was thus identical to that in the
tests described in section 5.6. The procedure for the plane
strain tests was, as far as possible, the same as that for the
triaxial tests. The only difficulty associated with the plane
strain tests was in removing air from the lower porous stone
where pore pressures were to be measured, and in particular
from the groove in which the porous stone is loacted. This
de-airing problem arises because the plane strain sample is set
up between the top and bottom plattens and enclosed in the
membrane prior to setting it up in the apparatus itself. This
is a highly advantageous arrangement as far as minimising
disturbance to the sample is concerned but means that during
setting up water may drain from the bottom porous stone and
allow air to enter. The procedure adopted for removing air
was as follows. After trimming and setting up the sample in
the apparatus, the cell was filled with water and the cell
pressure applied. This was left overnight and the pore pressure
measured the following morning. The pressure in two of the
mercury pot pressure sources was then adjusted so that one gave
284

a pressure slightly above and the other a pressure slightly

below this pore pressure. These pressures were then connected

to the tubes leading to each end of the groove containing the


bottom porous stone. Water thus flowed into one end of the

groove and out the other, carrying with it any trapped air,

and did not have a significant effect on the stress state in the
sample. A perspex trap was installed in the system to catch

any air driven out. This procedure appeared to be effective as

in most cases a small amount of air was immediately driven out


of the system. After removal of air the samples were allowed

to consolidate against a back pressure of 200 kN/m2. The time


allowed for consolidation was 3 days and during this period a

further attempt at removing air was made by flushing water


through the bottom porous stone in the same manner as previously.

Before commencing the shearing stage of the test a check was

made on the effectiveness of de-airing by raising the cell

pressure and measuring the pore pressure response, with all

drainage leads closed. The B parameter was almost invariably

found to be unity. The fact that some air was initially

- trapped in or around the bottom porous stone means that there

is some error in the measured values of initial suction and in

the volume change during consolidation. An examination of the

suction values and volume changes listed in tables given later


(5.8.1 and 5.8.2) confirms this as the suction values are, on

average, higher for the plane strain samples than the triaxia.l

samples, and the volume changes somewhat lower.

These plane strain tests are open to some criticism on


the ground that isotropic consolidation of the samples involves

strain in the horizontal direction and thus in the direction of

the plane strain axis. The sample is restrained along this


285

axis at the upper and lower faces by the loading plattens so


some non uniform distortion of the sample occurs during con-
solidation and the stresses in the sample at compretion of con-
solidation are likely to be somewhat uneven, at least near the
ends of the sample. However, the volume changes during consoli-
dation are relatively small (of the same magnitude approximately
as in Atkinson's tests on London clay) and it is thought that
this effect is of minor importance.
Apart from air in the lower porous stone, a certain
amount of air is always trapped between the membrane and the
sample. This did not appear to present any difficulties. As
the water level in the cell rose, this air was forced to the
top of the space between membrane and sample and ended up in
the space between the membrane and top platten. After about
24 hours this trapped air usually disappeared, presumably by
diffusing through the membrane and going into solution in the
water filling the cell.
Where material permitted, four tests of each kind were
carried out in an attempt to take into account normal test
scatter and natural variations in the material.

5.8.3 Test Results


The test results are set out in Tables 5.8.1 to 5.8.5:
and in Figs. 5.8.1 to 5.8.14.. The first three tables summarise
the results of the triaxial and plane strain tests on vertical
and horizontal samples; the deviator stress and pore pressure
versus strain curves are plotted in Figs. 5.8.1 to 5.8.6.
Fig. 5.8.7 shows a plot of failure values and Fig. 5.8.8 gives
averaged stress paths from each type of test. Fig. 5.8.10
shows curves of intermediate principal stress from the plane
286

strain tests and Fig. 5.8.11 shows A values from vertical and
horizontal triaxial compression tests. Tables 5.8.4 and 5.8.5
and Figs. 5.8.12 to 5.8.14 give the results of the triaxial
tests at varying inclinations.
An examination of Figs. 5.8.1 to 5.8.4 shows that there
is more variation in the results of this series of tests than
in the previous series described in section 5.6. Because of
this scatter these tests were not as productive as had been
hoped. They provide a consistent picture as far as peak strength
values and pore pressure changes are concerned, but there is
too much scatter in the shape of the stress strain curves to
be able to draw definite conclusions about the relative diffe- -
rence in the slope of the curves between triaxial and plane
strain loading and between vertical and horizontal loading.
The results are discussed in detail below in the order in
which they appear in the figures.
A comparison of Fig. 5.8.1 with 5.8.3 and 5.8.2 with
5.8.4 shows that the behaviour during plane strain loading and
during triaxial loading is very similar. There is very little
difference in either Peak strength values or in the pore
pressures, although the pore pressure in the plane strain
tests appears to be marginally higher than in the triaxial
tests. It is of interest to note that there is a definite
difference in pore pressure response between vertical and
horizontal samples, both in triaxial loading and in plane
strain loading. At the same value of strain the pore pressure
in the horizontal samples is consistently about 20% higher
than in the vertical samples. There is thus not a unique
relationship between pore pressure and strain as has been
reported for other soils by Lo (1966) and Parry and Nadarajah

(1974).
287

The extension tests on vertical samples shown in


Fig. 5.8.5 are almost identical to those reported earlier in
Fig. 5.6.9. The extension tests on the horizontal samples,
given in Fig. 5.8.6, show that the pore pressure response is
different and that the strength is slightly higher than in the
same tests on vertical samples.
The strength values from these tests, after averaging,
are given in Table 5.8.3. In the third column the strength is
expressed as a proportion of the strength in vertical triaxial
compression tests. The ratio of strength in horizontal com-
pression to that in vertical compression is just over 70%. This
value is obtained both from triaxial tests and from plane strain
tests and agrees well with the values obtained earlier from
direct undrained tests (Fig. 5.3.31). This variation in
strength should thus represent the maximum difference in un-,-
drained strength to be expected in any field situation in which
the loading conditions are plane strain, making the assumption
that the strength is not greatly influenced by whether the
samples are initially under an isotropic stress state or a Ko
state. The extension tests on vertical samples give a strength
of only about 56% of the strength in vertical compression and
thus appear to overestimate the anisotropy in undrained strength
by some 15%.
In Fig. 5.8.7 the failure values are plotted from
all the tests. These appear to all lie on the same failure
line although there is a tendency for the plane strain points
to be slightly higher than the triaxial points. In this aspect
the tests show a similar trand to that in tests by Henkel and
Wade (1966) and Vaid and Campanella (1974). The best fit line
through the points in Fig. 5.8.7 is slightly below that obtained
288

earlier from the consolidated undrained series (Fig. 5.5.6).


The values of c' and (p' are as follows
c' = 5.7 kN/m2
= 28.5°
With hindsight, it appears that the line in Fig. 5.5.6
represents an upper limit rather than an average from that
series of tests, and the above values may be more representative.
In Fig. 5.8.8, averaged stress paths are shown from all
tests on vertical and horizontal samples. These demonstrate
clearly that the differences in strength are primarily the
result of different pore pressure response, rather than diffe-
rences in the effective stress failure envelope. It is of
interest to note in passing that the mode of failure in triaxial
extension tests on horizontal samples was distinctly different
from that in the same test on vertical samples. The modes of
failure are shown in Fig. 5.8.9. With the vertical samples,
symmetrical necking always occurred, as mentioned earlier.
With the horizontal samples failure clearly occurred on
distinct planes inclined at about 25° to 30° to the plane
perpendicular to the sample axis. These failure planes were
originally vertical planes in situ. The same mode of failure
was observed in all four samples tested in this way and appears
to suggest that the shear strength (ie the effective stress
failure envelope) associated with these planes is less than that
on alternate planes on which failure could occur. However the
effect is very minor and not of practical significance. The
different modes of failure may in fact be related more to
differences in the stiffness of the soil skeleton in the
vertical and horizontal direction than to differences in strength.
With compression tests the mode of failure in vertical
289

samples was very variable. With some samples failure planes


were visible but with others the deformation was fairly uni-
form. With compression tests on horizontal samples, deformation
appeared to be uniform throughout the samples and even at large
strains no failure planes could be seen. With the plane
strain tests, failure always occurred on distinct planes
running from corner to corner of the sample, in the same way
as in earlier tests shown in Figs. 5.3.24 to 5.3.27.
Fig. 5.8.10 shows the values of intermediate principal
stress during the plane strain tests. These show considerable
scatter but are of the same order as the values obtained in the
earlier undrained tests shown in Fig. 5.3.29. Fig. 5.8.11
shows the values of the pore pressure parameter A in vertical
and horizontal triaxial compression tests. The value is about
0.4 in vertical samples and 0.65 in horizontal samples.
These tests show that for vertical and horizontal
samples, the differences in strength are due to different pore
pressure changes and that there does not appear to be any
anisotropy of the c' and (1)1 values. This is in agreement with
Simons (1963) who found with Norwegian clays that for vertical
samples at least the values of c' and c' were not significantly
affected by stress path. If there is any variation in c' and
(1) 1 with orientation then the limiting values are likely to be
associated with either the vertical plane or the horizontal
plane, and would therefore be unlikely to show up in tests on
vertical and horizontal samples where failure takes place
across these planes rather than along them. Tests were there-
fore carried out on samples inclined at angles of 15, 30, 45,
o
60 and 75 to the vertical. Lower shear strength on either
the vertical or horizontal plane would then be most likely to
290

show up in tests on samples inclined at 30° and 60° respectively.


The results of these tests are plotted in a polar
diagram in Fig. 5.8.12. The compression tests show the same
steady decrease in strength with increasing angle to the vertical
as in earlier tests. The extension tests, however, show the
opposite trend, the minimum strength being obtained from
vertical samples. However the strength difference as i in-
o o
creases from 0 to 90 is relatively small. Fig. 5.8.13 shows
the failure values from all the tests, and it is apparent that
the point are all equally scattered about the same straight
line. This line is the same failure line as that given in
Fig. 5.8.7. There is thus no evidence from these tests of any
significant anisotropy of the effective stress failure envelope.
There is no evidence at all of any bedding plane effect such as
that reported by Lo and Milligan (1967) for example, from tests
on Canadian clays. Fig. 5.8.14 shows the averaged stress paths
in compression and extension tests at i = 0, 30, 60 and 90°.
This again demonstrates that the undrained strength variation
results from a different pore pressure versus deviator stress
relationship rather th-an from variation in the effective stress
failure envelope.
These last tests were carried out on specimens trimmed
from Cylinder T27, whereas the vertical and horizontal tests
were on material from Cylinder T28. However, the material
from the two cylinders, and from the other cylinders at the same
depth can be considered to be virtually identical, at least as
far as strength tests are concerned. This is confirmed by a
comparison of the average strength from triaxial compression
tests on vertical, isotropically consolidated samples. Such
tests have been carried out at the same strain rate on material
291

from Cylinder T27 (section 5.7), Cylinder T28 (above) and


Cylinder T29 (section 5.6). The average strength was as follows:

Cylinder Compressive Strength (kN/m2)


T27 31.4 (Table 5.7.1)
T28 32.8 (Table 5.8.3)
T29 31.8 (Table 5.6.1)

There is thus only a very small variation in strength and


the results given in Figs. 5.8.12 and 5.8.14 are unlikely to
have been affected by the fact that they include results from
different cylinders.
While no definite statements can be made regarding the
relative slope of the stress strain curves because of the scatter,
the following general comments apply. The difference in slope
between plane strain and triaxial loading appears very small and
probably less than the 4/3 ratio suggested by elastic theory.
The change in initial slope of the stress strain curve with
sample inclination also appears to be small and the tests con-
firm the evidence from earlier undrained tests (Figs. 5.3.16
to 5.3.18) that the initial modulus in horizontal compression
is only marginally less than that in vertical compression.
The stress strain curves diverge as loading proceeds, the point
of divergance occurring between about 1/3 and 1/2 of the devia-
tor stress in the vertical stress.
Table 5.8.1 Cylinder T 28 Consolidated Undrained Tricixigi Tests
( all samples consolidated to 20.7 k N / m2

Volume Failure Values


Test Type Sample Water Density Initial Change
No Content gm Suction during CE 0- St rain du
IC m3 A_ H
% 7 Cons. % k N / m2
kN/m- yo

a 54.1 1,58 17.8 -0,4 36.3 1,8 13.0


c 58.6 1.65 24.4 - 0.0 33.7 1.8 11.7
Vertical Compression
n 63 7 1.63 5,9 - 2,6 29.1 4.0 12.8
o 63.4 1,63 6.6 -2.2 32.2 3.9 12.6
b 52.6 1.66 24,2 -0.6 26.8 5.3 15.2
d 60.6 1.67 5.9 - 2.8 22.7 2.3 17.1
Horizontal Compression .
i 64.1 1.63 8.3 -2.3 20.8 5.4 15.5
i 60.2 1.66 9.7 -1.5 23.4 4,7 16,8
e 62.2 1.64 9.9 - 1.6 -17.6 -3.2 3.0
Vertical Extension f 60.9 1.64 11.4 -1.6 -18.8 -4.0 0.7
k 64.6 1.63 8.1 -1.8 -17.2 -5.1 2.5
P 64.2 1.61 9.7 -1.5 -20.3 -6.2 0.4
g 63.5 1.64 4.9 - 2.5 -22.4 -2.0 - 4.6
h 60.7 1.66 3.7 - 3.3 -22.8 -2.1 -4.1
Horizontal Extension 61.8 1.64 9.0 - 1.4 - 23.4 -2.6 -5.9
1
m 66 8 1.61 8.8 - 1.4 - 24.3 -2.6 - 5.5
Table 5.8.2 Cylinder T 28 Consolidated Undrained Plane Strain Tests
( all samples consolidated to 20.7 kN / m2 )

i Volume Failure Values


Test Type Sample Water Density Initial Change
No Content gm Suction during 0-- 0- Strain .Au 0--2 -0-
1 3
- /crn kN/m2 Cons.
kN/m2 kN/m2

d 62.8 1.61 8.6 -0.6 35,4 2.1 13.5 6.1


Vertical Compression a 58.3 1 63 17.4 -0.04 36.8 1.6 14.5 9.2
g 60.1 1.61 10.9 -1.0 35.5 1.9 14.3 51
h 64.3 1.61 11.4 -0.6 35.3 2.1 14.6 6.5

c 53.6 1,67 16.3 -0.3 26.0 2.8 17.2 7.0


Horizontal Compression b 58 3 1.64 18.2 -0.6 26.6 2.5 17.2 6.8
e 59.9 1.61 14.8 -0.6 23.1 2.3 18.1 5.8
f 61.1 1.63 13.3 -1.1 23.4 3..0 18.9 6.9
Table 5.8.3 Cylinder T28 Averaged Values from Cons. Undrained Tr faxiat and
Plane Strain Tests

Strain at Peak Deviator ( (37


1 — Cji
)
Test Type Failure Stress kN/m2
(a-- 0— ,from triaxiai
0- )fro
% 1 3verticca compression tests
( al — Cri )

Triaxial Vertical Compression 2.8, 32.8 1.00

n Horizontal 11 4.4 23.4 0.71

ii Vertical Extension — 4.6 18.5 0.56

i: Horizontal ,, —3.1 23.2 0.71

Plane Strain Vertical Comp. 1.9 35.7 1.09

Horizontal 2.7 24.8 0.76


Table 5.8.4. . Cylinder T 27 Cons. Undrained Triaxial Compression Tests at
Varying Inclinations
( all samples consolidated to 20.7 k N / m2 )

Sample Inclination Water Density Initial Volume Failure Values


No to Content .n3 Suction Change
gm /ci
Vertical % . kN/m2 during
0::_o- Strain Au
Cons. % i-% H `)/0 kid/ m2
7 15 57.0 1.66 20.2 - 0.2 32.8 2.6 14.1
17 IJ
58.9 1.65 9.8 -1.2 35.5 4.3 14.0
3 30 57.4 1.65 17.3 -0.2 31.9 2.3 15.1
11 .., 57.9 1.67 6.2 - 2.3 29.8 7.7 154
19 " 60.0 1.64 8.7 - 1.4 28.4 3.5 14.6
1 45 50.7 1.71 15.1 - 0.5 29.3 3.5 14.3
15 Ji
58.8 1.66 10.3 - 1.4 27.2 4.9 15.6
21 ,, 60.3 1.65 10.0 - 1.5 25.3 3.7 15,5
5 60 60.0 1.64 16.4 - 0.6 26.2 4.3 17.4
1, - 1.8
13 58.7 1.67 8.7 25.2 6.2 17.5
23 ,, 57.7 1.68 12.3 - 1.0 27.2 5.1 15.4
9 75 57.2 1.69 5.4 - 3.6 27.4 7.7 14.4
25 1 45.2 1.76 17.3 - 0.1 25.5 2.4 163
Table 5.8.5 Cylinder T 27 Cons. Undrained Triaxial Extension Tests at
Varying Inclinations
( alt samples consolidated to 20.7 kN / m2 )

Volume Failure Values


Sample Inclination Water ' Density Initial Change
No to Content gm km2 Suction during
Vertical Strain .6,u
°/0 kN/m2 Cons.
kN/m2
8 15 58.0 1.66 8.1 - 2.5 - 23.0 -4.3

CA.)Dl
I I I I I 1 I I I I I 1
18 0 58.0 1.66 8.1 - 1.2 -21.4 - 1.2
4 30 54.7 1.67 17.7 -0.2 -24.7 -4.0
12 58.6 1.67 85 - 1.8 -22.8 -4.4
20 is 59.8 1.64 12.2 - 1.1 - 22.8 -2.7

--1h606 k)L.) tA)


2 45 53.0 1.68 16.5 - 0.2 -26.2 -5.9
16 0 59.3 1.66 11.1 - 1.3 -21.7 -3.6

W
22 11
59.1 1.66 11.8 - 1.0 - 21 .9 - 2.0
6 60 52.3 1.64 15.4 - 0.7 - 23.2 -3.3
14 0 58.3 1.67 9.3 - 2.3 - 22.9 - 4.4
94 ii 53.2 1.67 10.4 - 2.2 -21.6 -3.5

(T1
10 75 59.1 1.66 7.9 - 2.2 -24.2 - 6.1
26 II
46.0 1.75 13.9 - 0.5 -26.2 -8.7
2 98
30

------.■.e
_____ ..a.- - - - -

25 fi

y
,...■■,.......J4■■■■■■•■■.1IN■■■.....■0■.

20
E
N
z 0.

15 •

bs
10

:s<

yx

0 ..0. . . .■ - - -.

15

N.
E 10
N
z

5 i .


1 2 1+
Strain

CYLINDER T 28 CONSOLIDATE D UNDRAINED TRIAXIAL


COMPRESSION TESTS : HORIZONTAL SAMPLES

F I G 5.3.2
299

35

30
d a

E 25
h
z

20
..

----. J.
10
N
E
z
5


0 1 3 4 6
Strain

CYLINDER 7 28 CONSOLIDATED UNDRAINED PLANE


STRAIN TESTS : VERTICAL SAMPLES

F G 5,8.3
300

6
2 3 4 5
Strain

CYLINDER T 23 (.1). LI DA UNHAINED PLANE


S I\ I TE;1.

F I 6 5.3./,
301

"'-
z
E 5_~
o >_. .~~4-"T - --- -.--- -
~
-------'\---- . - - - -
~
\~ \ ..
e
f
k
-51---+-----·-+------1-- P -+-----1------+-----4
:J
<l

~O~____~____~-__--~-__--~-

o~----~--~~----~----~----~----~------

N
E

"z
.:::(.
-10~~--+-----~-----~----~-

bf -15 ~:______+--~---_I_ k -----i----t-------i

o 1 2 3 L, 5 6 7
St r a in ( - ve) %

CYLINDER T 28 CONSOLIDATED UNDRAINED TRIAXIAL


EXTENSION TESTS : VERTICAL SA~/1PLES

FIG 5.8.5
302

h
E

C'

--
- ■------9,
-----------

-5 --A_

-1
4

-10

-5

\)
-10

-20

'------ --'1----
I
-25
r-
n
Strain

CYLINDER T28 CONSOLIDATED UNDRAINED TRIAXIAL


EXTENSION TESTS: HORIZONTAL SAMPLES

F I G 5.8.6
20
A 0 .....- ,---
AM ....----
---- o
C
.....-- .---o o Triaxial Compression Vertical
.--- o
N A Plane Str.
,
..,s,„. ...--
........3x --- Triaxial Horizontal
7)<Q c V Plane Str. is
, -,- a
10 ›.---r Triaxial Extension : Vertical
/'
fl• .,,,- + + 11 Horizontal
/1

c" \/: *1---


Lc, /
Ci 5/ r) 4
----'1J ■ i"
11)
(NI

I
10 20 30

07 -•
2
3 k N / m2

CYLINDER T 28 FAILURE VALUES FROM CONSOLIDATED UNDRAINED TESTS


(,0
cp

o Trioxiat Compression : Vertical ,--


A Plane Strain ,,, • ,, ..--
® Trioxi al n : Horizontal
V Plane Strain ,, • ,.,
+ Triaxial Extension : Vertical ,--
15 .---'
x ,, /J
SO Horizontal
■ //
,--'
..----
...--- /
v"
--
..--- x i /7/0
..-- ' i.,
..---
10 ----
0 ------
(-. ..--- +
10 ri.. .....--
x 0
el) + \\\
---- x
'7, \ ‘57\ \

'-i-\ \
---- w
5 \ 0
-I- X

0— i 0-1
1 + 3 kN / m2
2
0 5 10 15 20 25 30

CYLINDER T 28 STRESS PATHS FROM CONSOLIDATED UNDRAINED TESTS ( AVERAGED CURVES )


305


Viewed from "above" Viewed from "side"

in original orientation in original orientation

Horizontal Sample

Syrnnetrical 'necking"

Vertical Sample

FAILURE MODES IN UNDRAINED TRIAXIAL


EXTENSION TESTS

FIG
306

15
VERTICAL SAMPLES

N
E
10

--.

15
HORIZONTAL SAMPLES

N
E 10
z
,-,...
,.---:--------
Lc' 5

bN

1 2 5
St rain

CYLINDER T 28 INTERMEDIATE PRINCIPAL STRESS


DIFFERENCE IN PLANE STRAIN TESTS

F I G 5.3.10
307

1. 0

0.9

0.8

0.7

0.6

0.5

0.4

0.3

02

® Vertical
0.1
Horizontal

Strain

CYLINDER T 28 CONSOLIDATED UNDRAINED TRIAXIAL


COMPRESSION TESTS VALUE OF A FROM VERTICAL
AND HORIZONTAL SAMPLES ( AVERAGED )

F I G 5.8.11
308

35 = 15°
Compression
Extension

30 0
Each of these =30
is average of
four tests

25
c\I
E = 45°

20r,
Compression

0) = 60°
15
2\
U)
Extension
21)(

5
Average of
four tests

5 10 15 20 25 30
Undrained Strength kN /m2

CYLINDER T 27 CONSOLIDATED UNDRAINED TRIAXIAL


COMPRESSION AND EXTENSION TESTS AT
VARYING INCLINATIONS

F I G 5.8.12
.:309

20

15

0 Av
za
N
..-- 0
0
1J
V--•- 0
10

Corny.

15 V
0 V
45 0
60
75
0
10 15 20
) /
3
kN / m2

CYLINDER T 27 FAILURE VALUES FROM CONSOLIDATED


UNDRAINED TESTS AT VA NYi i''JG INCLINATION S

'5 :; .13
-410
17

15

10

Compression Tests

0
1

=60°
10
i=30
x

x -90
Ns.

1-0

Extension Tests

10 0-3- 15 20 25
2 / m2
STRESS PATHS IN UNIDRAINED TRIAXIAL TESTS AT
VARYING INCLINATIONS
FIG 5 .8 .14
311

5.9 The Pore Water Tension in the Samples

5.9.1 Values Measured in Triaxial Samples


It has been noted earlier in sections 5.5, 5.6 and 5.8
that the pore water tension (suction) values deduced from the
triaxial samples were very variable, and covered a range from
close to zero to up to about 50% greater than the expected value
for a "perfect" sample. The values obtained are summarised in
Fig. 5.9.1 which shows in histogram form all the determinations
actually made. These include those from earlier sections and
those from the drained tests described later. This plot shows
that the values most frequently obtained were in the 8 to 14
kN/m2 range with a fairly regular fall off in frequency on
each side of this range.
Several reasons can be advanced to explain the large
scatter and the fact that the average value is substantially
below the expected value for an ideal sample. Firstly, during
the sampling operation itself there was the possibility that
the soil absorbed some water due to the stress reduction as
the test pit was excavated. This would result in the cylinder
samples having a pore water tension less than the "correct"
value. Secondly there is the possibility of further change
occurring during the operation of trimming and setting up
the 3 in x 11/2 in triaxial samples. It seems that substantial
changes must take place during this operation as a large range
of values was obtained in samples trimmed from the same block.
This range is probably not surprising in view of the fact that
the soil contains many very small vertical holes (probably
root holes) and occassional shells, both of which normally
contained free water. Often when a saw cut was made in the
312

soil, water appeared on the surface, having been squeezed out


of these root holes by the cutting action. Similarly, when
shells were encountered, a relatively large reservoir of water
was usually released on to the soil surface. During storage
in the cylinders this water was probably held in the root
holes and shells under the same negative pressure as the
pore pressure in the block as a whole, but was released during
the trimming process to be absorbed by the soil with a
consequent loss of pore water tension. During the trimming
operation itself there was some loss of water due to surface
evaporation. The significance of this is discussed in
section 5.9.3.
A further factor having some influence is the rubber
membrane. A dry membrane appears to absorb a certain amount
of water from the soil even in a relatively short time. In
the undrained tests described in 5.3 the membranes used were
normally dry and it was found that the sample weight determined
after testing was consistently about 0.5 gms less than the
weight before testing. The difference appears to be mainly
due to the loss of water absorbed by the membrane. In the
following sections some simple experiments are described which
help to clarify some of the factors discussed above.

5.9.2 Pore Water Tension Measured Directly in Sealed Blocks


As the pore water tension in this soil was very low it
was possible to make direct measurements of it by inserting a
probe into the side or top of a block of the soil and reading
the negative pressure with a pressure transducer. This was
done using a probe consisting of a high air entry ceramic tip
0.7 cm in diameter by 1.8 cm long cemented to a brass housing
31.3

to which two small diameter leads were attached. One of these


leads was connected to a pressure transducer and de-aired
water supply so that it was possible to flush water through
the system when necessary. Measurements were made on two
blocks of clay from Cylinders T30 (3.18 - 3.48 m) and T42
(3.48 - 3.78 m). Each block had the same diameter as the
cylinders (10 in) and a height of 4 in. These blocks of clay,
after removal from the cylinders, were stored in the normal
way, that is wrapped in polythene sheeting and sealed with
wax. A small hole was made in the wax and sheeting so that
the probe could be inserted and then sealed over again with
wax. With the block from Cylinder T30 the probe was left in
for a period of three months and the readings taken over this
period are shown in Fig. 5.9.2. During this period several
triaxial samples were trimmed from the portion of the block
remote from the probe. This involved removing the seal over
a small portion of the block and resealing it again each time.
Checks on the zero reading of the transducer were made at
frequent intervals but it was found that very little change
occurred over the three. months period. Fig. 5.9.2 shows that
the pore water tension remained fairly constant although it
showed a slight increase with time. The average value was
about 8 kN/m2. The samples trimmed during this period were
used for drained tests and are listed in Table 5.10.1. The
initial suctions from the triaxial samples given in this table
varied from about 8 to 14 kN/m2 , and thus tend to be somewhat
higher than the value in the block from which the samples were
trimmed. It appears that in this case the evaporation loss
during trimming had a greater effect than any absorption of
free water which may have been available from root holes.
314

With block T42 the probe was kept in the block for several
days only and the pore water tension value measured was
11.2 kN/m2.

5.9.3 Influence of Moisture Evaporation During Sample


Trimming
The rate at which evaporation occurs and its effect
on the pore water tension has been investigated by setting up
two 3 in x 11/2 in cylindrical samples exposed to the atmosphere
alongside one another on a laboratory bench and making measure-
ments throughout the day as evaporation occurred. The pore
pressure probe was inserted into the side of one of the samples
and the other sample was used for weight determinations as
the probe connection did not allow the same sample to be used
both for weight loss and pore pressure measurements. The
experiment was carried out twice, the first time with samples
from Cylinder T30 and the second time with samples from
Cylinder T42. The results are shown in Figs. 5.9.3 and 4.9.4.
The first figure shows the weight loss and the pore water
tension plotted against time. The rate of evaporation is
remarkably constant over the 7 hour period during which readings
were taken and is almost identical for the two samples. The
pore water tension values are less consistent; the sample from
Cylinder T42 shows a more rapid increase in tension than that
from Cylinder T30. The pore pressure transducer ceased to
register further changes after the tension value reached 45
and 49 kN/m2 respectively. The reason for this is not known
but it appears that cavitation was occurring somewhere in the
system.
In both cases the rate of evaporation and rate of
315

change of pore pore water tension is surprisingly large. A


delay of half an hour between trimming a sample and setting it
up in the apparatus would result in a pore water tension
increase of between 5 and 10 kN/m2 , i
in addition to whatever
change may have occurred during the trimming process itself.
In general the time involved in trimming and setting up the
11/2 in by 3 in samples tested in this thesis was fairly short,
resulting from the ease with which samples could be trimmed
with a wire saw. The total time involved in trimming and setting
up one sample was about 30 minutes, so that the actual time
during which evaporation could occur was considerably less than
this. However, it must be remembered that interuptions in the
middle of a sample trimming operation are sometimes unavoidable
so that with some samples significant loss of water through
evaporation undoubtedly occurred. The writer's sample trimming
operations were generally performed between 8:30 am and 10:30 am
when the likelihood of interuptions was at a minumum.
Fig. 5.9.4 shows the moisture loss plotted against the
pore water tension. These curves are in reasonable agreement
with the conpressibility curves given later in section 5.10. \-
(Fig. 5.10.10).

316

22

20 –

18

( 216
1

14

ti 12
a;
0 10
.
6 8-

1,5 6
_a
4-
z

22 24 2 6 28 30
Initial Pore Water Tension ( us ) k N/ m2

VALUES OF INITIAL PORE WATER TENSION ( SUCTION )


DETERMINED . FROM TRIAXIAL SAMPLES

( us = Cell Pressure — Pore Pressure )

Fl (3 5.9,1
12

10
0 0••■• •••••..
■....

0 0 ,
0 0
....., 0 0
0._ 0,
,. , .__x_0--0
,.:...,se

T3 li
s CU a)
-c)
c)
E E E E
E 4- E ._
4:: -4.-0 .4--•

0-) 10) 0
Q} 2 a 0-
C)

CL 12.
L.
0 C E
0 0
V) V)
0
V) V)


18 26 4 12 20 28 4 12 20 28 4 12 20 28 4
Feb March April May June
CYLINDER T 30 PORE WATER TENSION IN SEALED BLOCK MEASURED
OVER THREE MONTH PERIOD
318
9 -
8 - - - - - - - --- ..... --

V" /
(V)

E 6
(j
/v
/ --_._- -- --

~
Vl .....
~
Vl
0 4 -;-~
-1

CJ
~
/
'-
:J ~~
-t-J
U)

0
2
Gf*
y ~

2:

Gf'
/
0

60

50 -;1""-

~/
....... - ,//
,/
""tJ

/'
E
~40
/' . ..J.. . .fJ/

..x V ....... 0
.../
/'

c
.Q30 / ,
L~
/"
/"

~
(/)
c
I
) /' "
0
....... /

lJ ........
,/
:/
/'
....... e

j /
./
/'

o from Cyl.
0>10
1- ~/
.....
0


~
T 42
L-
a
D-

. bEl ~
AI
,/ ~
II
" T 30

~
"

o 1 2 3 4 5 6 7 8
Time ( hours

EVAPORATION LOSS OF iv10lSTURE AND CHANGE IN


PORE WATER TENSION IN 3inx1~- in SAt"1PLES
EXPOSED TO LABORATORY ATtvl0SPHERE

F[G 5.9.3
319

Pore Water Tension ( kN/m2 )


0 10 20 30 40 50 60 70
4.
.. c't,
2
.
.‘
eN
4
e .
E
U

6 o
N

U) .
o
0
_J
--\
8
a)
J 0 from Cy( T42
U)
.6 10
. - T 30

12

EVAPORATION LOSS OF MOISTURE VERSUS PORE


WATER TENSION IN 3 in x 1-,k in SAMPLES
EXPOSED TO LABORATORY ATMOSPHERE

F IG 5.9.4
320

5.10 Drained Tests

In this section the results of a number of drained

tests are presented. These include tests to determine Ko,

tests to measure the relative stiffness of the soil skeleton


in the two directions, and tests following various stress paths

after consolidation to the in situ stresses. Some considera-

tion is given to the determination of the preconsolidation

pressure and to the relationship between drained and un-


drained behaviour.

5.10.1 Estimation of Ko
As mentioned earlier, no direct measurement of Ko in

the field has been possible. A laboratory estimate of the

"virgin" value of Ko has been made by consolidating an un-

disturbed sample under Ko conditions in the triaxial cell. The


sample was set up in the same way as in earlier tests and the
pore pressure measured after the application of a known cell
pressure. The starting point of the test was therefore the
value of the initial pore water tension. The vertical stress was

then increased under drained conditions and the lateral stress

adjusted to maintain the condition of zero lateral strain. To

ensure adequate dissipation of pore pressure the test duration


2
was 23 days, ie an increase in vertical stress of about 6 kN/m
0h2
per day. The formula t 277-5 cv (Bishop and Henkel 1962)
= 6
gives the time for 95% dissipation of pore pressure as 14 hours
and the readings were taken at 24 hr intervals.

The lateral strain was calculated from the volumetric

strain and the axial strain. To ensure no lateral strain these

had to be kept equal. This meant making adjustments to the Kopp


321

variator B as the test progressed. Actually, during the first

few days of the test, when the required increase in lateral

stress (cell pressure) was very small, the increase was made

by raising the mercury pot manually, but after this initial

phase the two pots were raised together and Kopp B adjusted

daily as necessary.

Details of the sample are given in Table 5.10.1, and

the results of the tests in Figs. 5.10.1 and 5.10.2. The first

figure shows a plot of 6'3 against a' and also the magnitude

of the lateral strains which occurred as the test progressed.

It is seen that the Ko value at the start of the test is very

low, and then increases gradually to reach a constant value of


about 0.45. It is believed that this value corresponds to the
true "virgin" bahaviour of the soil. As mentioned earlier

(5.6.2) the formula Ko = 1 - sinq)' gives a Ko value of 0.47.

The Ko value at the start of the test is only about 0.15.

Fig. 5.10.2 shows a plot of the vertical compression


versus vertical effective stress from this K test alongside
o
the same plot from oedometer tests. These oedometer tests

are described in section 5.10.2. It is evident that the curves


are similar in shape and show a very definite "preconsolidation"

pressure. The value of this preconsolidation pressure is some-

what higher in the curve from the triaxial test than the
oedometer test.

As a check on the Ko value obtained in the above test a

second sample was set up and consolidated under the constant

stress ratio of 0.55, which was used in earlier tests. Exactly


the same procedure was used as in consolidating the samples back

to the assumed in situ Ko stresses, except that the process was


continued to much higher stress levels. The sample details are
322

given in Table 5.10.1 and the test result is given in Fig. 5.10.3.

The upper part shows the stress path followed and the lower
part shows a plot of axial strain against volumetric strain

as the test proceeded. The volumetric strain is seen to be

consistently greater than the axial strain, indicating that the


sample diameter is decreasing as the stresses are applied.
The ratio 0.55 is thus greater than that applicable to true

Ko consolidation. While this test does not give the precise


value of K
o it indicates that it is somewhat less than 0.55
and agrees with the results of the Ko test described above.

The strains occurring in the initial stages of this test'


correspond closely with those found in the earlier tests and
summarised in Fig. 5.6.3.

5.10.2 Measurement of Soil Skeleton Stiffness in the Vertical


and Horizontal Direction

It has been shown earlier that in undrained compression

tests the variation in strength with inclination of the sample


axis is the result of variation in the pore pressure response
rather than variation in the values of c' and (p'. To account

for this variation in pore pressure response it is necessary

to examine the relative stiffness of the soil skeleton in the


vertical and horizontal direction. This has been done in

several ways, as set out below.

(a) Evidence from Volume Changes during "Ko" and Isotropic

Consolidation.
The strains occurring. during consolidation to the in

situ K0 stresses (discussed in 5.6.3) suggested that the skele-


ton is less stiff in the horizontal direction than the vertical

direction. This point is brought out in Fig. 5.10.4, which


323

shows a plot of the volume change occurring during isotropic

consolidation to the mean of the vertical and horizontal stresses

in situ, and during "Ko" state. The volume changes are plotted
against the initial pore water tension. It is clear from this

figure that the volume changes during isotropic consolidation

are consistently greater than those during "Ko" consolidation.


On average the volume change during isotropic consolidation is

some 50% to 100% greater than during "Ko" consolidation.

If for example the initial pore water tension is

10 kN/m2 then the volume change during isotropic consolidation

is about 1.5%, but during 'Ko" consolidation is only 0.8%.


The explanation for this behaviour appears to lie in the fact

that during isotropic consolidation the vertical stress is


less and the horizontal stress is greater than the corresponding

stresses during "Ko" consolidation. The reduction in compression


in the vertical direction during isotropic consolidation is

more than counterbalanced by the increase in compression in


the horizontal direction which can only result if the skeleton

is less stiff in the horizontal direction. This argument is

qualified by the fact that the true mean stress level during
isotropic consolidation is actually slightly greater than that

during "Ko" consolidation. With Ko taken as 0.55 the value of

1/3(ai + (1) is 0.775 aj after isotropic consolidation

but only 0.70 al after "Ko" consolidation. It does not seem

possible that this small difference in itself can account for


the large difference in volume change during consolidation.

(b) °odometer Tests.


The simplest method of measuring the relative stiffness

in the two directions is by performing oedometer tests on


324

vertical and horizontal samples, as has been done by Zeevaert


(1953), for example. Two vertical and two horizontal tests
of this type were carried out on samples trimmed from Cylinder
T30. Loads were added at 24 hour intervals with constant
stress increments of 11 kN/m2 over the first stages of the tests.
In the latter stages of the tests this increment was increased.
The results are plotted using a linear scale in Fig. 5.10.5.
It is clear that the vertical samples show a very marked
"preconsolidation" pressure and that below this pressure the
soil is considerably stiffer in the vertical than in the
horizontal direction. The stiffness ratio is nearly two, up
to a stress of about 35 kN/m2. The preconsolidation pressure
is lower and less pronounced in the horizontal direction.
The behaviour of the soil is thus very similar to that re-
ported by Zeevaert. In Fig. 5.10.6 the curves are averaged
and plotted in the conventional way as void ratio against pressure
using a log scale for the latter. The anisotropic effect dis-
appears at higher stress levels and the two curves become
identical. This applies also to the rate at which consolida-
tion occurs. Fig. 5.10.7 shows curves of degree of consolida-
tion plotted against root time for two loading increments.
At the lower stress level, consolidation occurs more rapidly
in the vertical direction than the horizontal direction, but at
the higher stress level the two rates are almost identical.

(c) Triaxial Compression Tests after Isotropic Consolidation.


In section 5.8, consolidated undrained tests were de-
scribed which were carried out on samples at varying inclina-
tions, all isotropically consolidated to an effective stress of
20.7 kN/m2. Drained tests were therefore carried out using the

same procedure, to see whether the differing pore pressure-


325

response in undrained loading could be related to differences

in the deformation modulus (or stiffness) during drained loading.

These tests were restricted to horizontal and vertical samples

only. The tests were carried out at a strain rate of about 1%


per day and four vertical and four horizontal samples were

tested. The primary object of these tests was to investigate

the deformation characteristics at small strains and for this

reason the tests were not taken to failure. They were stopped
after a strain of 10% had been reached. The results are given

in Figs. 5.10.8 and 5.10.9. The first figure shows the results

of each set of four tests and the second figure shows the re-
sults after averaging each set of values. The soil behaviour in
these tests is clearly similar to that in the oedometer tests

shown in Fig. 5.10.5. Up to a deviator stress of about 15 kN/m2

the strain in the horizontal samples is almost twice that in


the vertical samples. The initial drained modulus of the

vertical samples is thus nearly double that of the horizontal

samples. Both stress strain curves show a fairly marked change


in slope after an initial relatively steep portion. The be-

haviour is this simildr" to that of the Norwegian clays reported

by Berre and Bjerrum (1973), who refer to the stress at which


the slope changes as the "critical shear stress" and consider
it equivalent to the preconsolidation pressure revealed in an

oedometer test.

The volume change versus strain curves are, sur-


prisingly, almost identical in the two sets of tests. At the

start of the tests the volume change corresponds to a Poisson's

ratio of about--0.10, as indicated in Fig. 5.10.9. This low value


of Poisson's Ratio is to be expected in view of the similarity

in behaviour between an oedometer test and a drained triaxial


326

test, and also in view of the low Ko value measured in the early

stages of the Ko test described above. A low Poisson's ratio


suggests that the lateral stress required to maintain the Ka

condition is likely to be very low at the start of the test,

and thus will not greatly affect the vertical stress versus
strain relationship at the low stress level. The relationship

between this drained behaviour and the undrained behaviour is

discussed in detail in section 5.10.4.

(d) Strains during Isotropic Consolidation.


A further measure of the relative stiffness of the soil
skeleton in the two directions can be obtained by comparing the

axial and lateral strains which occur during isotropic consolida-

tion. Fig. 5.10.10 shows the result of such a test carried out

on a 3 in x 11/2 in triaxial sample from Cylinder T30, the

starting point of the test being an isotropic stress state


equal to the initial pore water tension in the sample. After

setting up the sample and leaving it overnight to ensure equi-

librium, the consolidation pressure was increased at a constant


rate by raising the mercury pot with the motor and gearbox

system. The rate of pressure increase was about 8 kN/m2 per

day. The volume change was measured in the usual way and the

vertical strain determined by raising the pedestal on which

the sample was mounted until the top cap just came into contact
with the load cell. Dial guage readings taken in this position

gave a direct measure of the vertical deflection. The nomen-

clature used in plotting the results is as follows:

eV = volumetric strain (positive for volume decreace)


eA = axial strain

eH = lateral strain
327

The results show that at the start of consolidation

the axial strain is less than half the lateral strain but as

the pressure increases the strain increments become more equal.

From a pressure of about 60 kN/m2 upward the strain changes in

the two directions have become almost identical, and the soil
is thus behaving isotropically. The test thus confirms the

earlier indications that at the in situ stress level the skeleton

is about twice as stiff in the vertical direction as in the


horizontal direction.

It appears that it is this difference in skeleton

stiffness which leads to differences in pore pressure response


and thus undrained strength in undrained compression tests at
varying inclinations.

5.10.3 Triaxial Tests following various Stress Paths after


"K " Consolidation
0
Some further information on the behaviour of the soil

during drained loading has been obtained by carrying out a

series of tests on samples initially consolidated to the assumed

in situ stresses. The 'samples were consolidated to these


stresses using the same procedure as that used for the samples

described in section 5.6. After consolidation the samples were


left for three days under the applied stresses before the next
stage of the test was started. Three pairs of tests were
carried out as follows:
(a)Tests in which the horizontal stress was held

constant and the vertical stress increased or de-

creased until failure occurred (samples u and s).


(b)Tests in which the vertical stress was held con-

stant and the horizontal stress decreased or


328

increased until failure occurred (samples t and r).


(c) Tests in which the deviator stress was held constant
and an equal all round stress increase or decrease
applied (samples v and y).

The stress paths representing these tests are shown in


Fig. 5.10.11. With the tests in which the horizontal stress
was held constant the intended stress path was automatically
followed, but in the other tests there was a small amount of
"wander" from the intended paths. The tests in which the
deviator stress was held constant were carried out by varying
the back pressure rather than by varying the cell pressure and
ram pressure together. This made the procedure very simple as
it involved the raising or lowering of one mercury pot only.
All the tests were carried out at constant rates of stress
change, the duration of the tests varying between one week
(sample r) and three weeks (sample t).
In presenting the results the following nomenclature
has been used:
Stresses: . A
= axial stress
o H = horizontal stress
Strains: eA = axial strain
e = horizontal (or lateral) strain
H
eV = volumetric strain.

The results are presented in Figs. 5.10.12 to 5.10.19


and details of the samples are given in Table 5.10.1. The
first four figures show the results of the first two pairs of
tests presented in the conventional form, that is as curves of
deviator stress and volume change versus axial strain.
329

The stress strain curve in Fig. 5.10.12 shows a fairly

marked change in slope at a deviator stress of about 30 kN/m2

which appears to correspond to the critical (or preconsolida-

tion) pressure noted earlier in the oedometer tests and in the


drained triaxial tests starting from the isotropic state

(Fig. 5.10.9). It is surprising that in neither case does the

change in slope of the stress strain curve lead to a similar

change in the rate at which volume change is taking place. The


volume change curve in Fig. 5.10.12 again indicates a very

low value of Poisson's Ratio at the Start of the test. The

other compression curve in Fig. 5.10.14 does not show a marked


shange in slope at any particular stress level, although the

volume change curve suggests a marked increase in the rate of

volume change as failure is approached. The curves from the

extension tests (Figs. 5.10.13 and 5.10.15) show that the


axial strains are very small over that part of the test in which

the deviator stress is reduced from the K


o value to the iso-
tropic state. Once the deviator stress has started to increase
again in the opposite direction the strains increase signifi-
cantly.

The failure values from these tests all agree fairly

well with the failure envelope from undrained tests as is


indicated in Fig. 5.10.11. However, sample u failed before

the envelope was reached, and sample t had crossed over the

envelope even though maximum deviator stress had not quite been

reached when the test was terminated at a strain of 14%.

Fig. 5.10.14 gives the results of the test in which

the deviator stress was hold constant and the vertical and
horizontal stresses increased by the same amount. This shows

the rather interesting fact that under these conditions the soil
330

behaves isotropically, ie the vertical and horizontal strains

are the same. This result is not unexpected as the greater

vertical stress at the start of the test will tend to reduce

the vertical stiffness of the skeleton in relation to the

horizontal stiffness, thus counterballancing the anisotropy

apparent when starting load applications from the isotropic

stress condition.
Fig. 5.10.17 shows the initial part of the curves in
Fig. 5.10.16 along with the results of the corresponding test

in which the stresses were reduced by the same all round amount.
The isotropic behaviour is no longer evident as in this case

the sample length hardly changes and the volume change is


almost entirely the result of horizontal strains.
The results of the first two pairs of tests are pre-

sented in a different form in Figs. 5.10.18 and 5.10.19. These

show the vertical and horizontal strains plotted against the


relevant stress change in each case. The effects of changing

the direction in which the stress is applied are more readily

seen in these figures.

5.10.4 Relationship between Drained and Undrained Behaviour


So far no attempt has been made to interpret the be-

haviour of the soil in terms of elastic theory or any other


mathematical "model", and questions of the "fundamental" be-

haviour of the soil in terms of such models are largely outside


the scope of this thesis. A brief examination will however be
made to see whether elastic theory provides a means of inter-'.

preting behaviour in drained tests and of relating such behaviour


to undrained tests. This examination is concerned only with the

initial portions of the stress strain curves and no attempt is


331

made to define a point at which behaviour could be considered

to depart from elastic. Anisotropic elasticity will be assumed

initially and the relevant parameters defined in accordance with

the nomenclature used in Chapter 4 (page 103). The modulus E


will be defined nominally as the secant modulus at an axial

strain of 0.25%.
Tests involving isotropic consolidation and "Ko" con-
solidation as the starting point will be considered separately.

(a) Isotropically Consolidated Samples.


The data given in Figs. 5.10.6, 5.10.9, and 5.10.10

suggests a vertical stiffness about twice the horizontal stiff-

ness so we can assume

n = 2
The value of E' and 13 can be obtained directly from the
vertical test using the relationships.:
I
deA =ET du'
x
and
de
1(1 - V)
p' =
3 deA

These give
E' = 3200 kN/m2

and
p3 = 0..10

The value of pi can be obtained from the horizontal

tests using the relationship


' dew
P3
1
pl deA

This gives pi = 0.15.


Assuming the above values of n, and the value of

E' can also be calculated from the isotropic consolidation test,


332

using the relationship


Ex =de
fl
l
+ 2n(1 - p') - 4 pil
deV 1 3

This gives EX = 2400 which agrees tolerably well with the value
above obtained from the vertical triaxial tests. This last
value is from only a single test so that the earlier value is
probably more reliable.
We can now use the above values and anisotropic elastic
theory to predict the behaviour of the soil in undrained tests
starting from the same isotropic consolidation stress of
20.7 kN/m2. This is done in Table 5.10.2 which shows values of
initial modulus in terms of the vertical drained modulus E',
and values of the pore pressure parameter A. The theory thus
predicts an initial modulus about 20% greater for vertical un-
drained tests and some 26% less for horizontal undrained tests.
Fig. 5.10.20 shows the undrained stress strain curve (from
Fig. 5.6.17, B series) plotted beside the drained curve (from
Fig. 5.10.9). These undrained tests were carried out at a
strain rate approximately 8 times faster than the drained
tests; to allow for this the undrained deviator stress values
need to be reduced by about 12%. Taking this into account
it appears that the relationship of the undrained curve to the
drained curve in vertical tests is in good agreement with that
expected from elastic theory. Apart from this one aspect
however the agreement with elastic theory predictions is not
good. The predicted difference in undrained modulus between
vertical and horizontal tests is not found in the experimental
results. Figs. 5.3.16 to 5.3.18 in particular show the initial
modulus to be almost the same for vertical and horizontal
samples. It is of interest to note that it was in this same
333

aspect that Atkinson's (1973) tests on London clay showed


greatest departure from the theory. The large difference in
vertical and horizontal modulus predicted by theory was not
found in the experimental results.
Predicted differences in behaviour between triaxial and
plane strain tests have also not been found in practice. The
difference in modulus appears to be much less than that predicted
both with vertical and with horizontal tests. The pore pressure
parameter A was also found not to vary greatly-between triaxial
and plane strain tests although the theory predicts values some
50% higher in plane strain tests. The A values for the tri-
axial tests, given in Fig. 5.8.11 are about 0.4 and 0.6 for
vertical and horizontal samples respectively and thus sub-.
stantially higher than the values of 0.2 and 0.4 predicted by
theory. This divergence from predicted pore pressure values is
illustrated in Fig. 5.10.21. This shows the actual stress
paths alongside the predicted ones. The agreement is somewhat
better in the extension tests than the compression tests. In
both cases the theory predicts the relationship between stress
paths from vertical and horizontal tests but does not correctly
predict their actual position in stress space. The stress
paths in the compression and extension tests do not have the
same gradient at the starting point, suggesting that the
0d elastic" parameters governing behaviour in extension tests

are different from those in compression tests.


(b) Samples Consolidated to the Ko State.
It is more difficult to evaluate these tests in terms
of elastic theory as the restriction to the Ko stress state as
starting point does not allow all the parameters to be deter-
mined, assuming of course that anisotropic elasticity can in

fact be applied. Howver, there is some evidence that starting


334

from this Ko state, and defining the elastic parameters in terms


of stress and strain changes from the Ko state, the soil be-
haviour may be close to isotropic. The result in Fig. 5.10.16
suggests this, at least with respect to increasing stresses.
We will therefore make the assumption that the soil is iso-
tropic, in which case E' and p' can be calculated in a number
of ways.
Firstly, from the tests in which aA is constant (Fig.
5.10.18) we obtain:
For increasing a' E' = 3300 kN/m2, p' = 0.10

For decreasing aA E' = 5100 kN/m2, p' = 0.35.


It is immediately apparent that the parameters applicable in
compression loading are not the same as those in extension
loading.
Secondly, from the test involving an equal all round
stress increase (Fig. 5.10.16)
3dul
E = (1 7 2 p)
V

Assuming p = 0.10, then E = 2700 kN/m2.

Thirdly from the tests in which q is constant (Fig.


5.10.19) we have the relationships
daH
E' = (1 -u 'de
)de
H
and
deA / deA
= d
eH "-de 2)
H

which give:
For increasing a, E' = 3050 kN/m2 , p' = 0.07

For decreasing a;/ E' = 1630 kN/m2, = 0.33


335

This last value of E.' is exceptionally low partly because of the

very marked curvature of the stress strain curve in this case.

We thus obtain a rather mixed set of values although

the values of E' and p' from all the tests involving an increase
2
in stress level are fairly consistent at about 3000 kN/m

and 0.10 respectively. The tests involving a reduction in


stress level show different values. In the test with a'

constant and a' decreasing the value of E' is higher but in

the test with a' constant and a' decreasing the value of E' is
lower. In both cases the value of p' is substantially higher.

The lack of agreement between the parameters from the different

types of test cannot be overcome by interpreting the behaviour

using anisotropic parameters rather than isotropic ones, as the

different values arise when the sign of the stress change is

reversed, rather than when the direction is changed.

If we now use isotropic elasticity to predict behaviour

in undrained tests then the modulus will be given by:

3 E'
E = + p') - 1.36 E'
u 2 (l
where
E' = drained modulus

Eu = undrained modulus
The actual undrained curves (from Fig. 5.6.17, D series,

and Fig. 5.6.18, F series) are shown in Fig. 5.10.22. In the

compression test the undrained modulus is substantially higher


than the drained modulus but in the extension test there is

very little difference. The stress paths from the undrained

tests are shown in Fig. 5.10.23, which also indicates the stress

path corresponding to isotropic elastic behaviour. It is evident

that in the compression test the path followed is almost iden-


336

tical to this. However, in the extension test the path is


different, and again suggests that the "elastic" parameters
applicable to compression loading are not applicable to exten-
sion loading. The stress path in extension tests is close to
that applicable to anisotropic behaviour with the vertical
stiffness twice the lateral stiffness, ie the predicted stress
path for vertical samples given in Fig. 5.10.21. This is per-
haps not surprising as the drained tests showed a much higher
E' value in the extension test involving a reduction in a A.
It is of interest to note that the pore pressure change
which occurs in a "simulated" sampling operation due to the
release of the deviator stress (Table 5.6.3 and Fig. 5.6.7)
is in fact comformable with anisotropic elastic behaviour.
The A parameter in Table 5.6.3 is close to 0.20 and thus almost
identical with that given in Table 5.10.2 for anisotropic
elasticity.
This attempt to interpret behaviour in terms of
anisotropic or isotropic elastic theory has not been very
successful. While the theory correctly predicts certain
aspects of behaviour it cannot be claimed that the soil be-
haviour in general approximates to elastic behaviour.

5.10.5 The Preconsolidation Pressure p'


To close this section, this brief comment on the pre-
consolidation pressure (ip) is included. The value of
can be estimated from a number of the tests already discussed
in this section. These tests and the corresponding pc values
are summarised in Table 5.10.3. The standard Casagrande con-
struction has been used in determining these values. The
lateral stress (a ) and the deviator stress at the same stage
A
337

of the test are also listed. The ocdometer tests and the K

triaxial test give the sharpest change in.gradient in the


stress deformation curve and the values from these tests are

considered the most reliable.

The sharp change in slope and thus definition of the

value is obscured to some extent in the other tests, all

of which involve some strain changes in the lateral direction.

It appears from the table that a lateral stress higher than


that corresponding to Ko loading tends to reduce pc and make

its value less apparent. There is however reasonable agree-

ment between the values and they confirm an apparent


overconsolidation ratio of 1.5 to 2.0.

Table 5.10.1 Cylinder Sample T 30 ( 3.18 - 3 48 m ) Details of Samples used for Drained
Triaxial Tests

Consolidation Stage Failure Values


Test Type Sample Water Density Initial
No Content 3 Suction , „
gm/cm AV AL OA: - OTri AV
kN/m2 '3 'A - 'Fl Strain
% % yo
k N /m2 k N /m2 k N/m2

K o measurement q 59.8 1.6 6 14.1 - - - - - - -

Constant stress
ratio consolidation o 61.0 1.63 12.9 - - - - - - -

Isotropic - 1.66 -
x 12.2 - - - - -
consolidation

r 58.9 1.66 8.8 14.7 12.3 1.37 1.1 25.0 2.7 - 2.3
Stress nar tal tests C
- 1 s 59.5 1.54 9.2 14.5 12.0 0.78 1.0 • -.1 6.6 7.0 -1.2
a f ter consolidation ) t 60.1 1.66 7.9 14.9 12.1 - 1.45 1.1 -78.9 13.7 15.0
to in situ K 0 'I u 59.8 - 1.67 9.9 14.5 12.3 1.73 0.7 40.0 12.5 5 .8
stresses. / v 58,8 1.67 13.7 n 14.5 12.4 0.52 0.8
59.8 1.5 7. 12.0 14.6 12.1 0 94 1.3 - - -
'''
w
CJ
CD
339

Table 5,10.2 Initial Deformation Modulus and Pore Pressure


Parameter in Undrained Compression Tests Predicted by
Anisotropic Elastic Theory
( All samples consolidated isotropically to = 20.7 kN/m2 )

Initial Deformation Modulus


Type of
A
loading
Drained Undrained

/
Triaxial Vertical E 1.190 Ex 0.20

Triaxial Horizontal 0.5 EX 0.735Eix 0.4 0

Plane Str. Vertical 1.01E- x/ 1.409 E„ 0.3 2

/
Plane Str. Vertical 0.51 E< 1 409 Ex 0.67

Parameters assumed:
n = 2
/1A3 = 0.10

/.4 = 0.15
Table 5 .10.3 Values of the Preconsolidatlon ( or Critical )
Pressure ( pc )

Figure , Deviator
Type of Test Pc/ H
OT Stress
Reference
kN/e. kN /m2 ( P/ — G-1-I )

Oedcmeter Test 510 5 41 ?

K 0 Test 5 .10 .2 51 21 30

Drained Triaxial Test 5.10.9 47 20.7 2 5.3


after isotropic Consolidation

Drained Triaxial Test 5.10.12 46 14.5 31.5


after "Kofi Consolidation

Isotropic Consolidation 5.10.10 38 38 0


Test
341

+02

+0.1 .
C
- 0
o •

-t-n e
— -01 e

0-0.2 1

100

80

0
0
cl 60
E

0
40

LC) Initial 41-0


tension
20

0 2n P1 170 14 160
kN / rn2

CYLINDER T30 TRUE K o TEST STARTING FROM


ISOTROPIC STRESS EQUAL TO THE PORE
WATER TENSION

FIG 5.10.1
342


Pressure
kN/ m2
0 20 40 60 80 100 120 140 160

0
4
0 Triaxiat K o Test

8
0
c
0
OedometerVY °
1;CL
2
1 12 Test
E
0 0

20

COMPRESSION VERSUS PRESSURE CURVES FROM


K0 TR IAXIAL TEST AND OEDOMETER TEST

FIG 5.10.2

3 41

60

...
(N- 40
Pore c.) 1
z Water O•CDS
Tension -,7•° '''
20 6•\

0` a
. . Assumed
1
in situ stresses

0 20 40 60 80 100

CT kN / m2

10 /
e
/
/ e
8 /

&I/

4," .
6 I/
c c° / a
Es '157

/ 6
4

ci
3<
L
e„,--/ In situ stress level
ryi,,..
O.' _
1; ' '-"--
0 8 10 12 14
ev OA

CYLINDER T30 DEFORMATION DURING CONTINUED


CONSOLIDATION AT ASSUMED K STRESS RATIO

F I G 5.10.3
3.
0\
Isotropic Consolidation to
Mean In situ Effective Stress o
0 ‘' 1‹: Consolidation to assumed
3
,
in situ Effective Stresses r
o

1/4,,+
i \


Isotropic
2 „------ o

\o o
1
-.r. c, o
CO N. 0
,----"-v-
Ka
No N

E
J + + "■,
0
0
---............s. ....,,,,
`'' .....,, --....


0 12 15 20
Initial Pore Water Tension ( Suction ) kN/rn2

COMPARISON OF VOLUME CHANGES DURING ‘‘Kc; CONSOLIDATION AND


ISOTROPIC CONSOLIDATION TO SAME MEAN STRESS LEVEL
10 20 30 40 50 60 70 80 90 100
,,,,,...„,,,
r----z.. Pressure kN/rn2
-----'',;-"■
-- . ---,_z__--___
,........
...---..7:----:,'-
-... -1.-

.., .,.
". \
2
\ \
\
\IL \
. ‘o \
\ \
4 \ \
\ N

\ e\
N4 Q
5 \ \\
\
_____ \ \
\ \
e \
8
0 Vertical N.
\

u-, 10 Horizontal
.0 N
N
.
F \ N
0 N a N
12
.
N N
N ... •.o
14 0
CYLINDER T30 : OEDOME-TER TESTS ON SAMPLES WITH COMPRESSION AXIS
VERTICAL AND HORIZONTAL ( Linear scale )
I
e Vertical
1.6
.0 0 Horizontal
,..,
-11)......... CO

1.4

1.2 Irk

1
O
1)

to
-0 '
0

0.8

re
Pressure kN/m2
0.6
10 50 100 500 1000

CYLINDER T30: OEDOMETER TESTS ON SAMPLES. WITH COMPRESSION


AXIS VERTICAL AND HORIZONTAL ( Log Scale )

347

® Vertical
10
Horizontal
\\ '
\\ Loading increment :
20 \\
o
-21.6 to 32.4 l<N/m2
0 \\\ ----- 550 1, 982
0 \\
30 0

...
75 \
0 0\
\\
° 40 \\
\\
\\
4-
\),..\ ,.
0
50 ■
-s--..
-...--.-.....-..
-,..,..,.
60 -.,---.-
.... e
)...z.......z... — ..
.. ....
70
0 1 2 3 4 5 6 7
j
Time ( min

CYLINDER T30 OEDOMETER TESTS : INFLUENCE


OF ORIENTATION AND STRESS LEVEL ON
RATE OF CONSOLIDATION

FIG 5.10.7
348

50

40

C■I 30
E

Vertical . Horizontal
10
Samples Samples

4 2
Strain Vo
0 4 6 8 4

a)

DRAINED TRIAXIAL COMPRESSION TESTS ON


VERTICAL AND HORIZONTAL SAMPLES
AU Samples Consolidated Isotropically to 20.7 kN/m2

FIG 5.10.8
349
50

40

E 30 a

16,0 20
Vertical
° Ho 'zonal •
a
10


0 10
Strain

0 6 10

4
:s,-.....
`-'," N. N....,...
/s.s. .N. ........,,

6 (6-Ns
..... -,,,

N
8

DRAINED • TRIAXIAL TESTS ON VERTICAL AND


HORIZONTAL SAMPLES (AVERAGED VALUES )

0 5 .10. 9

Pressure kN/m2
10 20 3U L,L) bU 50 70 50 Y U 1U U
'-.- -------- —
E'S -1- 1-
z

,6. eV A
0 e t.,1
6
eA

ti.

VOLUMETRIC , AXIAL , AND


HORIZONTAL STRAINS DURING
ISOTROPIC CONSOLIDATION OF six.

A TRIAXIAL SAMPLE
351

0 10 20 30
A
2

CYL T 30 DRAINED TRIA)(IAL TESTS STARTING FROM K o


STRESSES STRESS PATHS FOLLOWED

F IG 5.10.11

352
0

6 I
Axial Strain °/0
4 5 6 11 12 13 14

40

30

E
z

- 20

10

CYL T 30 DRAINED TRIAXIAL TEST STARTING FROM K0


STRESSES:(u) 0- CONSTANT • GT-
A- INCREASING

F I G 5 i0;12
353

e -.0
,

,..-0.
..-----------
Axial Strain (-ye)
2 3 4 5 6
15

10

E
N
z

-10

-15

-20

CYL T30 DRAINED TRIAXIAL TEST STARTING FROM K0


STRESSES (s) 0- CONSTANT 0-;- DECREASING

FIG 5.10.13
354
---r---
4
",

",
.; e
all
."
3 ,<I!:r-.

;' 0
".
~'"
~ /'
o 2
v --

v
/
a 1/ Strain ( °10 )
o 2 3 4 5 6

25 ~.

~
~--(9--_
-0_
--- 0--

20 f - -
.
/ --- .

N
.
E
~ 15
.;:;{.

tjI 10
I
b<t

5 .~

o "-

CYL T 30 DRAINED TRIAXIAL TEST STAf<T1NG FHOivl Ko


STRESSES: (r) 0; CONSTANT DH DECREASING
355

-12

-16 L


Strain % ( -ye)

0 3 4 5 9 10 11 12 13 14
20

"E
z
-20

LT-40

-60

-80

CYL T 30 DRAINED TRIAXIAL TEST STARTING FROM Ko


STRESSES: ( t 0-- CONSTANT (T1-
1 INCREASING

F I G 5.10.15
356

kN 1m 2
50 60

c 6
o
'-

CYL T 30. DRAINED TRIAXIAL TESTS STARTI[\JG FROM Ko


STRESSES~( v) EQUAL ALL ROUND STRESS INCREASE
AT COi\l.STANT VALUE OF 0; - OH

FIG 5 ;lO . 'l.6


357


.n0-1 kN/ m2

15 10 5 0 5 10 15
—1.5

0
—1
0

'11

- - -
0 ----'■-..„...
.---' 43

..:-

4
0.5
c

1.5
e A
V
2 eA o
eH 0.

2.5 ,..............

CYL. T 30 DRAINED TRIAXIAL TESTS STARTING FROM Ko


STRESSES EQUAL ALL ROUND STRESS CHANGES
AT CONSTANT VALUE OF 0– —

I 6 5.10.17
358

50
'\ --------
,

45


4 *

Starting Stress
( K 0 State )
c4E: 2
z

2J

Isotropic Stress
1
State ( = Cc; )

1
_
DRAINED TRIAXIAL
. TESTS: Ci:i CONSTANT
c.,

e o
H
eA .

:,..------ 0
-3 -2 —1
Axial and Lateral Strain .
( Cl

F I 0 5.10.18

359

f•

50

45

40

35

3
O

° Isotropic Stress
0
State (--: Cr—' 1
V
2

N
z
2

1 Starting Stress
( K o State )

0 -
1 DRAINED TRIAXIAL
0
TESTS ; OT CONSTANT

0 s I
.
eA
e o
H

. ,.
2 3 4 5
—2 —1 0
Axial and Lateral Strain ( 0 /0 )

F I G 5.10.19
360

35

...........il_____..0
1 ..D---....--1--,----__.1.
30

20

/.

L.0 10
Vertical Samples

cu
Drained 0
Undrained 0

3
. Strain. i % )

TRIAXIAL COMPRESSION TESTS : DRAINED AND


UNDRAINED STRESS STRAIN CURVES AFTER
ISOTROPIC CONSOLIDATION

FIG 5.10.20
361
Vertical tests
Horizontal
15

10
Compression
tests
E
z

Extension Stress paths


tests predicted from
drained tests

—10

7C.


10 / / 15 20 25

2 kN/m2

ACTUAL AND PREDICTED STRESS PAT HS IN UNDRAINED


TRIAXIAL TESTS STARTING FROM ISOTROPIC STATE

F I G 5.10.21
7.--- 362
35

30
a

tl
e
o

20 a

t
z Compression
K o Stresses

10
Extension

6<
0

—10

Vertical Samples
, Drained a
Undrained

-20
—3 2

Strain i% )

TRIAXIAL TESTS : DRAINED AND UNDRAINED STRESS


STRAIN CURVES AFTER K oll CONSOLIDATION

FIG 5.10022

363

25
/
I .
20 /
------

o
15
N
E
z Compression
_,-----
----
10

Stresses

5 /
N

Extension /

5
. /
-Isotropic elastic stress
' path predicted from
-----_,..„„
drained tests
/
—10 ---....:.,,,..,

■ ,t.,.

-15
0 5 10 15 20 25 30
I;N / m2
2

ACTUAL AND PREDICTED STRESS PATHS IN UNDRAINED


TR IAX IAL TESTS STARTING FROM K o STRESSES

FIG 5 .10.23
364

5.11 The Undrained Residual Strength

A number of "undrained" ring shear tests have been

carried out in an attempt to obtain information on the rate of

fall off in strength with displacement on a shear plane and

to measure the undrained residual strength. These tests were

carried out on samples cut from cylinder sample T26 from a


depth of 3.18 to 3.48 m.

The high permeability and coefficient of consolidation

of the soil suggested that truly undrained tests would be


difficult to carry out and this indeed proved to be the case.

The porous stones of the ring shear apparatus could easily be

made impermeable by sealing them over with waterproof paper

but drainage through the gap between the porous stones and the

confining rings could not be effectively prevented. Also

there was no way of preventing loss of water from the gap


between the upper and lower confining rings. The two factors

which needed some preliminary investigation in order to ob-

tain conditions as close to undrained as possible were the

normal load on the sample and the rate of travel (ie the rate
of rotation).

5.11.1 Tests on Undisturbed Samples

The first test was carried out with a normal stress

equal to the total overburden stress in the field (= 50 kN/m2


= 7.2 psi) and a rate of rotation of 10/min. (a displacement

of 0.11 cm/min.). With this procedure it was found that there

was only a slight fall off in shearing resistance after the peak

stress was reached. By rotating the apparatus by hand through


180° the value was decreased substantially (to about half the
365

peak value) but as soon as the motor drive was resumed the

strength rose rapidly again. It seemed clear from this be-

haviour that consolidation took place very rapidly and that with

this procedure there was little possibility of obtaining a

reliable undrained result.


A second test (b) was therefore carried out with a

greatly reduced normal stress and a higher speed of rotation.


2
A normal stress of 9 kN/m (1.3 psi) was used with a speed of

rotation of 5°/min. (0.55 cm/min.). The result of this test

was again not satisfactory as the peak load was very low and

only dropped very gradually to a constant value. When the

sample was dismantled it was found that the shearing had not

taken place at the centre of the sample but between the

sample and the upper confining stone. This danger is always


present when low normal stresses are used. It appears that
there is insufficient lateral stress to develop adequate shear

resistance between the sample and the confining rings, so

that it is relatively easy for the sample to slide around


between the rings and shear along the top (or bottom) confining

stone.
To try and overcome this difficulty the next test (c)

was carried out using a slightly higher normal stress at the

start of the test and reducing this stress once shearincr.had

started. The initial stress applied was 18 kN/m2 (2.6 psi)

and this was reduced to 9 kN/m2 (1.3 psi) after 3 minutes.


The speed of rotation was 5°/min. (0.53 cm/min.) at the start

of the test but this was increased by a factor of 5 to 2.58 cm/min.

after the test had been running for 4 minutes. The speed was
again increased by a factor of 5 after a further 4 minutes and

maintained at this speed for 8 more minutes after which the


366

test was stopped. The total test time was 16 minutes. The

stress displacement curve printed out by the chart recorder

suggested that this test had been relatively successful. The

curve is shown in Fig. 5.11.1. It is seen that there is a

fairly sharp peak after which the stress falls rapidly and

eventually levels off to a reasonably constant value. When

the sample was taken out and examined it was found that a very

distinct shear plane had formed at the centre of the sample and

there was no evidence of movement between the sample and the


top and bottom plattens. The surfaces on which shear occurred
appeared smooth and flat and were covered by a thin "paste"

of remoulded soil. They were not "slickensided" as is normally


the case with stiff overconsolidated clays.

Water content determinations were made on the intact

soil before and after the test and also on the paste of re-

moulded soil. The results were as follows:

W %
Intact Soil Before Test 53.8
After Test 53.4
"Paste" from slip surface 48.8

It is apparent that even with the low normal loads 'used

in this test some loss of water has still occurred from the -

remoulded soil and the residual strength indicated by this test


is likely to be a little above the true value. A further

somewhat unsatisfactory aspect of the above test was that the

measured peak strength was considerably below the peak value

of 18 kN/m2 measured in undrained triaxial tests on the same


soil.
367

An attempt was therefore made to repeat the above test

using the same procedure except that the reduction in normal

stress was made after 2 minutes, not 4 as previously. However


the result of this test was unsatisfactory as shearing clearly

took place only between the sample and the upper confining

stone. A second attempt was then made to repeat the test but
this time shearing occurred between the sample and the bottom

porous stone. It appeared that the successful completion of

test (c) above was due more to good fortune than good manage-

ment.

5.11.2 Test on Undisturbed Sample with Pre-cut Plane

Further attempts to test intact undisturbed samples

were abondoned and instead a sample was prepared with a pre-cut


shear plane through the centre of it. It was a relatively

simple procedure to do this using a wire saw. Before setting

up the sample in the apparatus the upper part was rotated

through 360° (relative to the lower part) by hand. The test


2
was then carried out with a normal stress of only 4.5 kN/m

(0.6 psi) and and initial displacement rate of 5°/min.

(0.53 cm/min.). After 10 minutes the rate was increased to


25o/min. and after a further 6 minutes increased again to
_
12bo/minute. The test was run at this speed for 4 minutes and

then the speed reduced to the original value of 5°/minute.


The result of this test is shown in Fig. 5.11.2 as a graph of

shear stress against displacement. It is seen that the shear

stress stays reasonably constant although it tends to wander up


and down somewhat when the fastest rate of rotation is used.

It is of interest to note that there was almost no peak in the

curve at the start of the test and the measured shear strength
368

after more than two revolutions of the apparatus was almost the

same as at the start. The influence of speed of rotation is

very small. At the completion of the test water contents were

again measured with the following results:

W %

Intact Soil Before Test 61.5


n After Test 62.2

"Paste" from slip plane 56.4

There is again a drop of about 5% from the undisturbed

to the remoulded soil suggesting that the measured residual

may still be on the high side. However the value is in good


agreement with that determined from the earlier test.

5.11.3 Tests on Remoulded Samples


In view of the difficulties described above in
attempting to measure the undrained residual strength from un-

disturbed samples it was decided that a more promising approach

might be to carry out a series of tests on remoulded samples


prepared at different water contents. The undrained residual at

the natural water content could then be obtained from a graph of

residual strength versus water content. For this purpose six


samples were prepared at water content intervals of about 3%;
the highest water content being just above the liquid limit.

The portion of the cylinder sample used for this series of tests

had the following properties:


Natural water content 49%

Liquid limit -51

Plastic limit 25

It should be mentioned here that there was considerable

variation in natural water content throughout this cylinder, and


369

presumably also in Atterberg limits. However the variation should

not greatly influence the validity of the results as tests on

another cylinder suggested that the liquidity index and sen-

sitivity were probably reasonably constant.


The undrained ring shear tests on these six samples

were carried out using a rate of rotation of 5°/min. (0.53 cm/min.)


2
and a normal stres of 1.3 kN/m (0.2 psi). In addition to the
ring shear tests, vane tests were also carried out on the same

samples. These were done using the laboratory vane which had

a diameter of 1.26 cm and a height of 1.90 cm. The rate of


o
rotation was 6 /min. which gave a time to failure almost the
same as the ring shear tests. These vane tests were carried

out as a matter of interest as no attempt appeared to have


been made prior to this to see whether the two types of test

gave compatible results. The ring shear tests were continued

for a period of 25 minutes but the vane tests were stopped

after 10 minutes. Two vane tests were.carried out at each


water content and the average taken 'though in every case the

values were almost identical.


It was found that in the ring shear test the shear

stress rose fairly rapidly to a peak which remained fairly


constant for several minutes and then decreased very slightly

to a "residual" value. When the test was continued beyond

about 15 or 20 minutes there was a tendency for this residual


value to rise again. The explanation for this seems to be that

consolidation was taking place very slowly during the test.

With the vane tests the stress rose rapidly to a peak and then
almost immediately decreased again at a more rapid rate than in

the ring shear test. Continuing the vane test did not result

in the stress levelling off to a constant value although there


370

was a drop in the rate at which it decreased.


In Fig. 5.11.3 graphs are shown of shear stress against

displacement for both ring shear and vane tests at two of the

water contents. These were typical of the series. For cal-

culating the displacement in the vane test it has been assumed

that shear occurred on the cylinder defined by the diameter of

the vane. There is a small error here as the displacement on

the two ends of this cylinder will be rather less. The curves

show very clearly that peak strength in the vane test is reached

at a much smaller displacement than in the ring shear test

and that the post peak fall off is much more rapid. The smaller
displacement in the vane test is presumably due to the fact
that the vane forces shearing to take place on a specific

surface almost as soon as the test is started. In the ring

shear test the soil is not constrained to shear on a specific


plane, although it is likely that as the test proceeds shearing

is eventually confined to a specific plane or at least to a

narrow zone. In the test with the lowest water content it was

possible after dismantling the apparatus to detect a specific


shear plane at the centre of the sample. The reason for the

continuous fall off in strength after the peak in the vane

test is not known definitely but is possibly due to a tendency

for a slight gap to open between the cylinder of soil being

rotated and the surrounding stationary soil, It is thus not

possible to obtain a "residual" value from the vane tests.

The results of these ring shear and vane tests are


summarised in Table 5.11.1. The residual values from the

ring shear test has been taken simply as the lowest value to

which the strength fell off after passing thepeak. The


371

difference between peak and residual was in all cases very


small. The results are plotted in Fig. 5.11.4. It is seen
immediately that there is remarkably good agreement between
the peak values from the ring shear and vane tests. The
residual value also lies close to the peak except for the sample
with the lowest water content where it was substantially lower.
Also shown in this figure are the results of unconfined compres-
sion tests on remoulded samples of this soil. These have been
restricted to the dryer end of the range of water contents
used for the ring shear and vane tests. Even with these samples
it was difficult to obtain a definite strength value as the load
was still rising slightly at axial strains of 20%. The values
plotted in Fig. 5.11.4 have been taken at this 20% strain.
It is seen that the unconfined strength values are consistently
lower than the values obtained from the ring shear and vane
tests. The unconfined test is clearly not a very suitable
means of determining the strength of very soft remoulded clays.

5.11.4 SumMary
To summarise this series of tests Fig. 5.11.5 has
been prepared. This shows the result of ring shear test (c)
on the undisturbed soil replotted as a graph of strength
against displacement. On the same figure the results of measure-
ments made by other methods are indicated. The peak shear
strength from the ring shear and vane tests plotted in the
figure has been taken from Fig. 5.11.4 corresponding to the
natural water content of 49%. There is a slight anomaly here
as the water content of ring shear sample (c) was substantially
higher but the comparison is probably still valid as the
difference in water content is presumably accompanied by a
372

change in composition. Also shown in the figure are the

strengths measured on the pre cut plane sample and determined

from undrained triaxial tests on remoulded samples. The peak

strength from undrained triaxial tests on undisturbed samples is

also indicated. As mentioned previously the peak value of

14.2 kN/m2 measured in the ring shear test is substantially

lower than the value of 18.4 obtained from the undrained tri-

axial tests. The reason for this is• not known.

Taken overall, the tests present a reasonably consis-

tent picture, and show that the undrained brittleness index

obtained in a ring shear test is in agreement with the value

one would expect from measurements of the sensitivity of the

soil. In this case the sensitivity is slightly greater than 6,


1
so that the brittleness index IB (= 1 - —) should be about 83%.
The value obtained from the ring shear test is approximately

80%, the exact value depending on whether the residual strength


is taken from test (c) on the intact soil or from the series

of tests on the remoulded samples.


Perhaps the most useful aspect of these tests was the

information they provided on the compatibility of the ring

shear apparatus and the vane test when used to measure the

strength of soft remoulded soil.


373

Table 5.11.1 Measurement of Undrained Strength of


Remoulded Samples in Ring Shear Apparatus and
Laboratory Vane ( Sample T 26 )

Water Ring Shear kN/rn2 Vaneeak


Sample Content
Peak Residual
kN / m2
1 37, 2 9.69 7. 36 9.79
2 39 -3 6.37 6.16 6.25
3 42.6 4 .65 4.43 4,60
4 45.2 3 ,42 3.23 3.52
5 48.7 2 .71 2 -47 2.78

6 51 -3 1.95 1.57 1.70


Travel Distance ( cm.
)
1.0S 2.11 7.26 12.42 39 66 92 119
20

0.53 cm / min --, . 2.56 cm /min > 13.3cm/rain 1.

16
0.,
a
a
. . C.
'r)
. .
C
(NI 12 u C
E rcf C.)
0_
. 2 a
z 0

. c..,
,
. 1"
0
co
2

•.
rO
C)

)4
.
c

. ., .. •. • • ' • • • • • • • •


8 10 12 14 16
Time ( minutes )
1
CYLINDER SAMPLE 26 UNDRAINED RING SHEAR TEST ( c ) RESULTS AS PLOTTED
BY CHART RECORDER
1
i i
Z 0.53cm /m. 4..‹.----- 2.58cm / min ', 13.3cm /min

l i

44 -41
d
EN--.......-....."-r.."3"-u--Nr•r
1

1
I

1
1

1

0 5 10 15 20 25 30 35 40 45

Displacement ( cm )

0.53 cm / min

In
U]
1)

4-.
111

_c

45 50 55 60 65 70 75 80 85 90
Displacement ( cm )

CYLINDER SAMPLE T 26 UNDRAINED RING SHEAR TEST ( f ) SAMPLE WITH PRE— CUT
SHEAR PLANE
376

10

8
C

E
O
6
z

O
LI)

O Ring shear
ro
(I)
_c • Vane
v.)

w = 37.2 °/6

...,..I
3 6
Displacement ( cm )

E
" O

z I ,
\\
r
2

0 Ring shear

I-
• Van e

u)

W :: 48.2 Cl.


0 2 3 5 6
Displacement ( cm )

CYLINDER SAMPLE T 26 STRESS DISPLACEMENT CURVES FROM


RING SHEAR AND VANE TESTS ON REMOULDED SAMPLES

F 1 G 5.11.3
377
11

A Natural Water Content = 49

10 Wt_ = 51

W = 25
\ P

\
9

\IN

111\
'4E

z \
5 \

\ ,
\ c.
Na

N..
`s..

3
ro
(/)

2 0 Ring Shear Peak.


-....„.
• 11 11 Residual. :t.c

X Vane Test Peak.

. LS Unconfined.

0

34 36 38 40 42 44 4G 48 50 52
0
Water Content O

CYLINDER SAMPLE T 26 RING SHEAR TESTS AND VANE


TESTS ON REMOULDED SOIL AT DIFFERENT WATER
CONTENTS

F I G 5.11.4

16
Peak shear strength from undrained
triaxial test = 18.4 kN/m2

12 I \
Residual value from ring shear
test with pre-cut plane.

Peak shear strength from


undrained triaxial test
cN, on remoulded soil.
8
z

Peak shear strength from


tf)
ring shear and vane tests
on remoulded soil.

1
.I.. 2.53cm /min ____>)
, 13.3 cm/min 7,-_,,_
JI
0.5 8 cm/rn

1 1 , ! I I I. I I I 1 I I
2 4 6 8 10 12 14 16 13 20 27 24 26 23 30 68 70 72 74 114 116 115 120
Displacement ( cm. )

CYLINDER SAMPLE T26 UNDRAINED RING SHEAR TEST (c) SHOWING RESIDUAL STRENGTH COMPARED WITH
REMOULDED STRENGTH
379

5.12 Long Term Consolidation Tests

Some limited information on the secondary consolidation

or creep behaviour of the soil has been obtained by carrying

out long term consolidation tests in standard oedometers.

These were carried out on samples from depths of 1.5 m and


3.3 m approximately. Two samples were prepared from cylinders

from these depths; the location of the samples within each

cylinder being as close to one another as possible. All four

samples were then loaded up to their effective overburden

pressure, and one of each pair of samples then had an additional


stress increment of 40 kN/m2 added to it. The samples were

then left to creep under these loads for about 15 months.


Details of the samples and the loading sequence are given in

Table 5.12.1. Where step loading was used the loads were left

on for only 40 minutes before the next movement was applied.


The compression values given in Table 5.12.1 represent the

total compression at the end of this 40 min. period. This

40 min. duration was a rather arbitrary figure, being somewhat

in excess of the time necessary for full pore pressure dissipa-


tion.

The results of these tests are given in Figs. 5.12.1

and 5.12.2. The first figure shows a direct plot of compression


against time, and it is clear from this that the behaviour of

the two pairs of samples is fairly similar. The samples loaded

only to overburden pressure undergo a comparatively small

amount of both primary and secondary compression, whereas the


samples loaded to 40 kN/m2 above the overburden pressure under-

go a much larger amount of both primary and secondary com-

pression. However, towards the end of the period over which


380

readings were taken, the rate at which compression was

occurring appears to be very similar for all four samples.

This is made clearer in the second figure, which shows the


creep rate plotted against time. The readings over the first

40 days show systematic trends with steadily decreasing creep

rates except for the shallower sample at its overburden stress,

in which case the rate remained fairly constant. After 40

days the curves all show periods of "instability" during which

the creep rate accelerated considerably. However, by the time


300 days had been reached the creep rates were settling down

again to steady values of very similar magnitude for all four


samples. Thus the ultimate creep rate appears not to be re-

lated to the stress on the sample.


The time interval until the creep rate accelerates

varies from sample to sample. The shallow sample with the

40 kN/m2 excess pressure on it showed a sharp increase in


creep rate after 45 days but the steeper sample with the 40

kN/m2 excess pressure did not show this sharp increase after

about 130 days.


The behaviour Of these samples is not unlike that

reported by Bishop and Lovenburg (1971) for the normally con-

solidated Pancone clay from Pisa.


Table 5.12.1 Long Term Consolidation Tests : Details of Samples
and Primary Compression Values

Cylinder Mean Water Density Applied Total I


No Depth Sample Content gm km2 Loads 'Primary „
(m) 1-(N / m2 ComRcession
A.,
a 89.6 1.51 17.0 0.68
T18 1.53
b 91.9 1.50 17.0 0.75
56.2 5.13

c 58.7 1.67 11.0 0.34


30.6 1.16
T 38 3.93
d 59.2 1.56 11.0 0.38
30.6 1.20
69.8 5.08
382

c
o

0 Overburden Pressure
15- 3.9rn { • II 1/

1.5 m {B!l:::. Overburden Pressure


+ 40 kN/m2
1 7.5 -
1/
" -

2
I I I· I II d
10
I I I Lilli
100
--'--
600
Ti me ( Days)

LONG TERM OEDOMETER TESTS: COMPRESSION


VERSUS TIM E CURVES

FIG 5:12.1
383

1-
G Overburden pressure
3.9 m { 4 " ... L.. O k N/m2.
"
1.5 m {G Overburden pressure
A " 11 + 40 kN/m2.


" "-
"•
~
(1)
0...
",
A
"
"
Ojr---~--~--------------------I---------------­

0.
(1)
(l)
'-
U
0. 011--- .-+---

--- --- G
. E1---& .... ...:
".,
\
.\ I

8 \ ,"
....... B/EJ
\.
'"
/

O.OO'II--'-~JJ , I EJ, ! '1 'L. .!'~_ _---:""_-J-~!--t..!--:'.___


' LI
L. 10 100 1000
( days)

LONG TERt\~ OEOO~/1ETER TESTS: CREEP F<ATE


V E F-< SUS T I f'.1 E
384

CHAPTER 6

DISCUSSION AND CONCLUSION

In this chapter the main findings of the experimental

work are summarised and discussed and some comments made about

the use of laboratory values in field analysis and design.

The laboratory values are also compared with those obtained

from in situ vane tests and those indicated by the behaviour of

the trial bank itself.

6.1 The Undrained Strength

The first question looked into with regard to the un-

drained strength was the influence which reconsolidating the

samples to the in situ stresses had on the undrained strength,


and it was found that the influence was in fact very small.

The reason for this is apparent when the stress paths followed

in undrained tests are studied in relation to the initial pore

water tension in the samples. The data on this aspect is


summarised in Fig. 6.1.1 which is a slightly idealised.diagram

based on the test results given in Chapter 5. The wide range of

initial pore water tension values and thus effective stress at

the start of conventional undrained tests does not lead to a


corresponding range of strength values as the stress paths

tend to converge towards a common failure zone.


Thus for this particular clay conventional undrained

tests give strength values only marginally less than those

from samples reconsolidated to the in situ stresses. This is

in direct contrast to the behaviour of Norwegian clays reported


385

by Bjerrum (1973) and does not support Bjerrum's general

assertion that samples should be reconsolidated to the in situ

stresses before testing in order to obtain a correct measure of

the undrained strength. This is not to say however that

Bjerrum's approach may not be necessary in particular cases.

More information is needed on the influence of soil type and

depth before any general conclusions can be drawn. It appears

likely that with samples taken from a considerable depth the

effect of loss of pore water tension is much greater than

with samples from a shallow depth and reconsolidation to in


situ stresses may provide a more reliable measure of undrained

strength. It is possible also that in soils of medium to high


plasticity the cohesion component of strength plays a more

important role than in soils of low plasticity and the effect


of reconsolidation is more important with the latter. These two

factors may explain why reconsolidation to the in situ stresses

appears to have a dramatic effect with Norwegian clays but has

little effect with the soil investigated here.


The influence of sample quality is a further factor to

which consideration must be given. The samples tested in this

thesis were large. "block" samples taken in cylinders from

shallow depths in a test pit, whereas the samples from which

Bjerrum's data was obtained were tube samples taken from bore-

holes. It is highly probable that the degree of disturbance


in the large cylinder samples is considerably less than that
normally encountered in tube samples, although reliable informa-

tion on this aspect is almost impossible to obtain as no


definitive means exists •for measuring sample disturbance. Ladd

and Lambe (1963) have used the value of pore water tension in

relation to the value applicable to an "ideal" sample as a


386

measure of sample disturbance. There may be some virtue in

this approach although it does not follow that a sample in

which the pore water tension is the "correct" value is necessarily

a perfect sample.
The pore water tension values from two blocks of the
2
Mucking clay were 8 kN/m2 and 11 kN/m (section 5.9.2) which
represent respectively 47% and 65% of the value for a perfect

sample of 17 kN/m2. The loss of pore water tension in this


case is therefore not as great as that frequently encountered

in tube samples, where a loss of around 80% is considered

typical by Ladd and Lambe (1963). From this aspect there is


thus some evidence to support the view that the large cylinder

samples are of superior quality to typical tube samples.

Examination of the pore water tension in the Mucking clay

emphasises the difficulties involved in obtaining reliable


values of pore water tension. Substantial changes can occur in

trimming small samples, so that values obtained by comparing

pore pressure and cell pressure in triaxial samples may be


significantly different from the value in the block from which
the samples were trimmed.

The second question investigated was the anisotropy of

the undrained strength. The behaviour of the Mucking clay


tends to support Parry's (1971) assertion that anisotropy is

not a predominant factor contributing to descrepancies between

actual field behaviour and behaviour predicted from laboratory


measurements. It is of considerable significance but is only

one of a number of factors which need to be taken into account.

The principal fact to emerge from the anisotropy study is that.

variations in undrained strength arise almost entirely from

variations in the pore pressure response which occur with


387

variations in the method of load application. The tests show

clearly that the effective stress failure envelope is not in-


fluenced by the inclination of the failure plane or the way in

which the sample is brought to failure. These findings help to

explain the descrepancies which appear to arise when different

methods are used for measuring anisotropy of undrained strength.


In particular, compression tests with the axis at varying in-

clinations give results which do not appear compatable with


vane tests, as has been discussed in section 2.2. The descre-

pancy arises when attempts are made to assign a particular value

of undrainedstrengthto a particular plane, ie a plane having


a certain inclination to the vertical. The writer's tests

suggest that the undrained strength on a particular plane will

vary according to the way in which the stresses are applied as

this will determine the resulting changes in pore pressure.


It should be emphasised that differences in undrained strength
are not necessarily the•.result of anisotropy of the soil.

They may equally result simply from differences in the method

of measurement. For example, undrained extension and com-


pression tests will result in different values even with an

isotropic soil so that Bjerrum's use of these tests as a

measure of anisotropy does not appear to be a valid procedure.


Similarly, in a vane test•on an isotropic material it is

possible that the strength on the horizontal end surfaces may

not be the same as on the vertical cylindrical surface. The

mode of applying the shear stress in each case is somewhat


different and may result in different pore pressures and thus

undrained strength. The above facts should be kept in mind


when evaluating methods used for measuring soil anisotropy.

Ladd and Foott (1974) for example present a set of values


388

(Table 1, page 765) of undrained strength for Boston Blue Clay

which are listed as indications of the anisotropy of the soil.


The tests comprise compression and extension triaxial and plane

strain tests and a simple shear test. It should be remembered

that strength differences similar to those in the table would be

obtained with a material which obeyed the laws of isotropic


elasticity and a Mohr-Coulomb failure criteria.

The most valid method of measuring undrained strength

anisotropy in relation to stability problems would be to carry

out plane strain compression and extension tests on samples


initially consolidated to the in situ stresses. Tests of this

kind automatically take account both of the anisotropy of the


soil and the different methods of stress application leading
to failure. Because of apparatus limitations these tests

were not carried out on Mucking clay. Results from other

clays were given in Section 2.3. In the absence of equipment

suitable-for such plane strain tests, the best measure of un-


drained strength anisotropy, in the writer's view, is provided

by undrained compression tests carried out at different inclina-

tions, as done by Lo (1965). Information so far available in-


dicates that tests of this type show the same order of undrained

strength variation as that obtained in compression and extension

plane strain tests on samples consolidated to the K


o state.
The variation in pore pressure response and thus un-

drained strength in undrained tests at varying inclinations

(with the Mucking clay) is shown to result from variations in


the stiffness of the soil skeleton in the vertical and hori-
zontal direction. The lower stiffness in the horizontal

direction results in higher pore pressure during compression

loading in the horizontal direction than during compression


389

loading in the vertical direction and thus leads to lower

strength in the horizontal loading direction. This provides a


more rational explanation of undrained strength anisotropy than

Bjerrum's (1973) explanation in terms of the greater ability


of the soil structure to resist shear in one direction than the

other.

The third aspect of undrained strength behaviour in-

vestigated in some detail was plane strain behaviour in relation


to triaxial behaviour. This was probably the least satisfactory

part of the test programme. Firstly, because of apparatus

limitations the tests were restricted to compression only.

Secondly the scatter in the test results was such that no very
definite conclusions could be drawn. The results suggest that

the peak strength is marginally higher in plane strain com-

pression than triaxial compression, and that the stress strain

curve is slightly steeper with a smaller strain to failure.

However, the difference is very small and could not be defined

quantitatively without carrying out a very large number of tests.


:3 9 0

( a ) Stress paths

90 010 of measured strength

values would lie in

this range

11
Correct undrained

strength

0)
E

/
/
/ / Stress paths in

/ (
undrained tests

/ / / i
In situ stresses

/ / / 1

/ / I

1..1
L I
10 20 30 40
0-1- '4- GT' kN/m2
2
90 0 /0 of initial pore

-4-- water tension values were

between these limits

b) Pore water tension


values

0 _10 20 30 40
Initial Pore Water Tension ( us ) kN/rn2

RELATIONSHIP BETWEEN INITIAL PORE WATER TENSION


AND MEASURED UNDRAINED STRENGTH

FIG 6.1.1
391

6.2 Use of Undrained Strength Values from Laboratory Tests


for Analysis and Design

6.2.1 Relationship between Principal Stresses and Strength

on Failure Planes

Undrained compression tests determine the strength of


soil in terms of the maximum deviatbr stress the sample can

withstand, ie in terms of the principal stresses at failure.

However, in the analysis of field situations, particularly those


involving the stability of slopes, slip circle methods are

normally used, which require that specific values of shear


strength be assigned to each part of the slip surface. It is

necessary therefore to make some assumptions about the way


in which the strength on a shear surface is related to the

strength measured in terms of principal stresses. Normal

practice is to take the undrained shear strength Cu as half the

deviator stress at'failure, and to assume this value will


operate around a slip surface. In compression tests in the

laboratory failure occurs on planes inclined at angles of


between 55° and 65° to the plane pe.tpendicular to the sample
axis, so that the shear resistance mobilised on this plane at

failure is clearly not half the deviator stress. It is given by


a1 a3
S = ( )sin 2a as shown in Fig. 6.2.1(a). The justifica-
2
tion for taking the strength on a shear surface as half the
failure deviator stress is not clear to the writer although

the difference between this value and the value given by


G
1 3
( sin 2a is relatively small. In the tests on Mucking
)
2
clay the angle a averaged about 60° and the strength on the
al c3
failure plane was therefore 0.86 ( or nearly 15% less
2
than the value of Cu where C (1 1 - (5 3. Skempton (1948)
u 2
392

does not discuss this point in detail although he states that the

conventional (I) 7-
. 7 0 analysis will correctly predict failure
stresses even though it will not give the correct position of

the failure surface.


In addition to the above uncertainty there is the added

complication of anisotropy. It has been emphasised in the

above discussion that the tests on Mucking clay (and tests on

other soft clays such as those of Le Lory and Lai (1971))

suggest that it is not possible to assign unique strength

values to particular planes. Despite this fact, when laboratory


strengths are measured in terms of deviator stress and analyses

are carried out using slip circle methods some assumption must

be made about the strength on planes in relation to the strength


in terms of deviator stresses. Lo (1965) has made the assump-
tion that the principal stresses at failure are inclined at a

fixed angle to the slip surface in the way shown in Fig.

6.2.1(b). That is
i = f 0 -

-where i = inclination of major principal stress to the verti-


cal ( = inclination of sample axis in laboratory

test)
f = angle between failure plane and plane normal to

sample axis
0 = inclination of the slip surface to the vertical,

measured as shown in the figure.


It should be noted that this does not mean that planes

inclined at the same angle are assigned the same strength. At

points B and C for example the inclination a is the same but


the angle 0 is different, so that the strength is different

due to the assumed relationship between 0 and the principal


393

stress orientation. At each point around the surface there are

in fact two possible choices for the orientation of the principal

stresses (assuming a fixed angle between slip surface and major

principal stress direction) and Lo does not discuss the justi-

fication for the orientation he has chosen. However, it

certainly appears the more likely of the two possibilities.

Along the slip surface from A to B for example the major prin-

cipal stress is likely to be close to vertical while from C

to D it is likely to be close to horizontal. The laboratory


tests therefore suggest maximum strength in the AB region and

minimum strength in the CD region.

Lo's assumption starting from the position of the slip

surface and reasoning backwards to obtain the position of the


principal stresses and the strength thus leads to approximately

the same distribution of strength as does direct consideration

of the principal stress directions and their relationship with


strength revealed by laboratory tests. The most logical pro-

cedure would be to use a method such as finite element analysis

to determine the principal stress directions and then to assign

strength values to each-point throughout the soil mass making

use of the laboratory test data. This would lead to some


differences with Lo's assumed distribution. For example Lo's

assumption assigns the maximum strength on the slip surface to

some point between A and B where the angle a is 60° (a1 in


vertical direction) and the minimum strength to a point near C

where a is 30° (a1 in horizontal direction). These values are


for the case where f is assumed to be 60°. Finite element

analysis would suggest that al is probably closest to the

vertical at point A and closest to the horizontal at point D

so that the strength should maximum and minimum at these points

respectively.
394

The above discussion is somewhat academic as the in-

fluence of anisotropy on the factor of safety is not very great.

Whether the anisotropy is determined from plane strain tests

such as those in Figs. 2.3.2 and 2.3.3, or from compression

tests at varying inclinations, the minimum strength is un-

likely to be less than about 60% of the maximum strength.

In the analysis of a typical slope this leads to a reduction

in the safety factor of about 22% (Lo, 1965). The same answer

would be obtained if the average of the maximum and minimum

strengths were used in a conventional analysis. The above

figure of 22% is based on circular arc analysis. In practice


the effect of anisotropy may be to make the slip surface some-
what non circular with consequently a greater influence on the

safety factor than that suggested by circular arc analysis.

6.2.2 The "Normalised Behaviour" Approach of Ladd and Foott

(1974)
In their 1963 paper Ladd and Lambe emphasised the diffi-

culties in obtaining accurate measurements of the undrained


strength of clay,due particularly to disturbance and the loss

of pore water tension which occurs during sampling operations.

However they did not favour reconsolidating the samples to the

in situ stresses before undrained testing as they believed

this would lead to an overestimate of the undrained strength

because of the reduction in void ratio occurring during re-


consolidation. They suggested a rather complicated procedure

for removing this effect. Ladd and Foott (1974) have pre-

sented a refinement of this approach which they call the

"normalised soil parameter" method. The steps involved in ob-

taining the undrained strength at a particular depth, .


395

illustrated in Fig. 6.2.2(a), are as follows:

1. The undrained strength is measured in laboratory


tests over a range of consolidation pressures up to a value in
excess of that at which "virgin" or "normalised" behaviour
is clearly established. A curve such as ABC is obtained, point B
marking the stress level at which normalised behaviour begins.
Above this stress the soil will exhibit a constant value of
S u/aVC'
where Su = undrained strength
aVC = vertical effective consolidation stress in the
laboratory test.
2. Using the value of a VC at point B, a further series of
tests are carried out in which the samples are overconsoli-
dated to OCR (overconsolidation ratio) values of say 2, 4 and 6.
A relationship between S u /aVC and OCR is thus obtained as
shown in the lower part of Fig. 6.2.2(a).
3. The OCR in the field is obtained from the value

(= o in Ladd and Foott's paper) determined in a conventional


vm
laboratory oedometer test and a knowledge of the in situ
stresses.
4. The undrained strength is then read off the relation-
ship in 2. at the OCR concerned.
This approach has the virtue of providing a good pic-
ture of the way in which the soil behaviour varies with stress
level and the way in which it is related to "virgin" or "nor-
malised" behaviour. However the method assumes that at the in
situ stress level any strength the soil has in excess of the
"virgin" strength is purely the result of stress history,
that is of overconsolidation. The tests are carried out at
stress levels higher than the in situ stresses and the behaviour
396

at the in situ stress level is obtained by extrapolating back-

wards from these higher stresses. The method thus obscures

the influence of structural or "cemented" bords which is nor-


mally only apparent when tests are carried out at or below the

in situ stress level. This appears to be a major drawback in

the method as interparticle bonds may make up a major part of

the cohesive component of strength at low stress levels. Fig.


6.2.2(b) illustrates in an idealised manner the way in which

Ladd and Foott's method may lead to a lower strength than the

true value because it obscures the contribution of structural


or cemented bords.
The method may be satisfactory for certain types of

clays, and particularly for samples taken from considerable

depth where the influence of stress level and stress history


may be much more important than any cohesive component of

strength. In general, however, and especially for samples from


shallow depths, it appears essential that high quality samples

be obtained and that tests be carried out at in situ stress

levels. This fact is emphasised by Raymond (1973) when discussing

the properties of a Canadian varved clay. Raymond points out

that the effective stress failure envelope of the soil can be


represented by a straight line through the origin (c' = 0) at

high stress levels, but that at low stresses the value of c'

is clearly greater than zero. Raymond states that at the time


the tests he is discussing were performed little was known

about the strength properties of cemented soils and adds that

"much more emphasis should have been put on obtaining good un-
disturbed samples and testing them at low confining pressures".
Raymond uses the term "unstructured value of 4)1 " to refer to
397
Plan e on which
failure actually
takes place

t
(E.: LT )slnaP, 03
2 2

Plane on which failure


is implied to occur

( a ) STRENGTH ON ACTUAL FAILURE PLANE AND THAT


NORMALLY ASSUMED FORUR 95 P0 LANANALYSIS

BET WEEN INCLINATION OF SL I P


b) RELATIONSHIP
SURFACE AND ORIENTATION OF PRINCIPAL
STRESSES ASSUMED BY LO (1965 )

F I G 6.2.1
393

Series of undrained tests at


various overconsolidation ratios

Series of undrained tests to


determine point B at which
normalised behaviour begins

Value of SuA--
- read from graph
-vc
at in situ value of OCR

Overconsoliclatton ratio ( OCR )

(a ) LA DD AND FOOT'S METHOD

Possible additional component


of strength from bonding not
taken into account by Ladd
and Foot's method.

Component of strength due


--------- to overconsolidation.
Normalised or "virgin" strength

Crvc
( b) POSSIBLE "COMPONENTS" OF SHEAR STRENGTH AT
FIELD STRESS LEVEL IN A SOFT CLAY

FIG 6.2.2
399

the value applicable at higher stresses where the effect of

•the structural component of strength is no longer apparent.

6.3 Laboratory Undrained Strength in Relation to Field


Vane Strength and Strength Indicated by Trial Embankment

6.3.1 Field Vane Strength


A considerable number of field vane tests have been

carried out at the site of the trial embankment, using the

standard "Geonor" vane apparatus. The results of these, for

which the writer is indebted to Binnie and Partners and to


Mr. R.S. Pugh, are presented in Fig. 6.3.1. The tests were

carried out at a speed of rotation which resulted in a time to

failure of about 5 minutes. This compares with an average time


to failure of about 12 minutes in the undrained triaxial tests

presented earlier in section 5.3. There is thus not a large

difference in strain rate in the two types of test, although

according to the trend shown in Fig. 5.7.9 it could account


for a strength difference of about 10%.
The vane test results do not vary greatly over the site,

and do not show a significant or consistent variation with


depth. Only after a depth of 5 m is there an indication of

the strength increasing with depth. The laboratory tests on

the other hand show a marked increase in strength with depth


below 2 metres. The .reason for this difference in behaviour is
not known although it may be partly attributed to the presence

of fissures in the upper 2.3 m of the soil at the test pit site.

These fissures (or shear surfaces) undoubtedly led to lower


strengths in the laboratory tests where the soil is free to fail
400,

on the weakest plane. However, the fissures may not have had

a significant effect on the vane test results where the failure

surface is predetermined by the action of the vane and the

soil cannot choose to fail on planes of weakness.

The average values from all the vane tests are shown
in the lower right of the figure, together with the average.

strengths from the undrained laboratory tests on vertical

samples. At shallow depths the vane strengths are substantially


higher than the laboratory strengths but deeper down the values

tend to converge. It may have been more relevant to draw the

laboratory curve through the average strength from the vertical


and horizontal samples instead of just the vertical samples.
This would not alter the upper part of the curve but would

reduce the values over the lower part of the curve by about
15%. Taking into consideration the influence of the fissures

it appears that the laboratory tests give strength values which

are generally only slightly lower than the field vane values.

The results tend to agree with those of Skempton and Henkel


(1953) obtained from tests at Tilbury and Shellhaven, respec-

tively on the upstreani and downstream sides of the estuary from

the Mucking site. Skempton and Henkel found no significant

difference between laboratory compression tests and vane


tests throughout the depth of about 15 m over which tests were
conducted.

6.3.2 Field Behaviour.


It is not the purpose of this section to give a de-

tailed account of the behaviour of the soil as revealed by the


trial embankment. This will be given by Pugh (1976). However,

the laboratory test data presented in this thesis is of


401

little practical significance unless it can be related to

field behaviour in some way. The writer is grateful to

Mr. Pugh for some preliminary results of his observations and

analysis of the trial embankment which make possible a brief

comparison between field and laboratory behaviour.


The embankment at Mucking was built of clean sand placed

in layers about 0.5 m thick. Deformation and piezometer


observations were made at the completion of each layer. One
side of the bank was made steeper than the other to ensure

failure on one side first. The bank was long enough to ensure

that behaviour was plane strain on the section over which

observations were made, and the construction time short enough

for the loading to be essentially undrained. Several aspects

of behaviour will be considered in turn as follows.

(a) The undrained strength.


Failure of the first slope occurred when the bank

reached a height of 4 m. A conventional undrained slip circle

analysis using the strength data given in Table 5.2.2 and


Fig. 5.2.2 gave a minimum factor of safety of 1.21. The

strength values used were the average of the vertical and


horizontal samples. With the field vane strengths the factor
of safety would have been somewhat higher. This value of

safety factor substantially greater than unity appears to be


in keeping with the findings of Bjerrum (1973), Pilot (1973)

and Parry (1971) who each analysed a number of case records

of embankment failures on soft clays. The data in section 5.7

suggests that the large difference in the rate of loading

between the laboratory tests and the trial embankment may be

the main factor accounting for the overestimate of safety


402

factor in the present case. In the laboratory tests the time

to failure was about 10 minutes while the field loading took


2 to 3 months. According to Fig. 5.7.10 this would account

for a reduction in strength of at least 30% and thus result

in a calculated safety factor of only about 0.9. Taking into


account the strain rate effect therefore suggests that the

laboratory tests underestimate the field strength.

(b) The deformation characteristics.


Predictions of field deformations based on the labor •

ratory stress strain data have been calculated using a finite

element computer programme written by Mr. M. Hama, fellow


research student at Imperial College. The programme assumes

non linear "elastic" behaviour and calculates the deformations

in a series of steps corresponding to the load increments in


this case. From the'stress strain curves for vertical samples
given in Figs. 5.3.1 to,5.3.18 a set of averaged curves have

been obtained and used for the analysis. The stress state in
the ground at the start has been calculated from an assumed
Ko value of 0.55. The most reliable field deformation measure-

ments were those obtained from the inclinometers and these

indicate actual deformations ranging between about 1/2 and 2/3


of the predicted deformations in the early stages of the
embankment construction, ie at low stress levels. At higher

stress levels the field deformations increase more rapidly

than the predicted values and as failure is approached the

field deformations are higher than the predicted values.

Taken overall, the measured deformations are thus in fair


agreement with the predicted ones and the initial deformation
modulus measured in the laboratory is not less than half the
403

value indicated by field behaviour. The lack of agreement in


the early stages of loading may be due more to the way in which
the computer programme handles the stress strain properties
of the soil than to the inaccuracy of the laboratory curves
themselves. An examination of Fig. 5.6.19 shows that the
initial modulus of the curves starting from the Ko condi-
tion is significantly higher than that obtained from the com-
pression curve starting from the isotropic state at a deviator
stress corresponding to a Ko value of 0.55. This latter value
is the value which the computer programme would make use of as
the modulus for each loading stage is obtained from conven-
tional stress strain curves (ie starting from an isotropic
stress state) at the appropriate value of deviator stress.
As the stress level increases the curve from the K starting
o
position flattens out more repidly than the curve from the
isotropic state. The use of the curve from the Ko state should
thus predict smaller deformations at the low stress level and
larger deformations at high stress levels and lead to better
agreement with the actual deformations.

(c) The undrained stress-displacement curve. and the


residual strength.
One aspect in which the laboratory data does not appear
to provide a good indication of field behaviour is in the
strength versus displacement relationship given in Fig. 5.11.5.
The rapid decrease in shear resistance with displacement
revealed in the ring shear test appears to suggest that once
failure started in the field the rate of movement would be
rapid and that considerable movement would be required before
equilibrium would be restored. Failure in the field actually
404

took place relatively slowly over a period of several hours,

and the position in which the sliding mass came to rest suggested

a reduction in strength to about 55% of the original value,

not to 10% or 20% as the residual strength and sensitivity

values would suggest. There are several possible explanations

for this. It is highly unlikely that failure in the field

occurs instantaneously on a continuous failure surface. It is

more likely to start in a zone beneath the sand fill and take

the form of a series of failure planes which eventually merge to


form a continuous slip zone which propagates outwards towards
the toe area of the fill. Even on the slip surfaces which do

form there may not be the same fall off in strength with dis-
placement as in the laboratory test due to some drainage occur-
ring into the adjacent soil.

Apart from the last aspect just discussed, the field

behaviour suggests that the laboratory tests, particularly the


conventional undrained tests in section 5.3, give values of un-

drained strength and deformation modulus which are in good


agreement with field values and sufficiently accurate for engi-

neering purposes. This conclusion is perhaps somewhat contrary


to current trends in the soil mechanics field. There appears

to be a growing belief that laboratory tests do not give results


which can be applied directly to field situations, particularly
as far as the deformation characteristics are concerned.
Marsland (1973)and Atkinson (1974) for example present data

for London clay which suggest that the undrained modulus in the
field is some two to five times higher than the value given by

laboratory triaxial tests. Butler (1974) has pointed out that


for London clay the value of Eu/Cu from laboratory triaxial
405

tests is in the vicinity of 150 but from field plate bearing

tests is from 600 to 900. Butler believes that a value some-


where between these extremes is applicable to London clay in

the mass.
It is important when considering this question to

distinguish between overconsolidated fissured clays and soft

normally consolidated clays such as the Mucking soil. Soil

mechanics literature contains numerous examples where measured

settlements on soft clays are in good agreement with values

predicted from laboratory e-log p curves. If it is possible


to obtain reliable e-log p curves from laboratory tests on
soft clays it should be equally possible to obtain reliable

drained or undrained stress strain curves from triaxial tests.


The apparent agreement between laboratory measurements of
deformation modulus and the field value in the present case is

given some support by an Eu/Su ratio obtained by Bjerrum (1964),


and reported by Simons ,(1974) from the behaviour of a tank at

Shellhaven which is in close proximity to the Mucking site.

The E /S u value was 220. With the Mucking soil the initial
u
deformation modulus from the laboratory tests is between 3000
2
and 4000 kN/m2 , and the shear strength about 15 to 20 kN/m .

This gives an. Eu/Su ratio of about 200 which agrees well with

Bjerrum's field value.

Much more information will be needed before general

conclusions can be drawn regarding the compatability of labora-


tory and field deformation behaviour for soft clays. Attention
needs to be concentrated on obtaining high quality samples for

laboratory tests (possibly requiring trial pits as in the present

case) and on reliable field measurements of deformation

behaviour.
1,,06
Shear Strens:.1th ( I< f\J / m 2
A 10 20 0 10 20 0 10 7'0
-~I-~-"";'--T---""'--S---'

Results from several loeCltions at trial embankment

Shear Strength ( kN/m2) Shear Strength ( l<N 1m2 )


o 10
---r
20 30 o 10 20 30
I /
Test Pit Labora\ory
Limits of scatter Curve //
1 - - - - i - - - ~----t \ ~
/ ---
\
Average
I tTeS\S
I' Vane

-- 2 t - - - - t - - .'r--I-----t
E ,..
-
I. J-
E
\
~ 3
0..
...c:
"0.3
a"
\-
--
<IJ
o

/~I-----f-
o
\r
-J~\
4

5~-~~~-------- 5 ~
Test P'lt I~esults Comparison average of
VGlU9S frOfn fie l d vane
and laboratory tests

FIELD VANE TESTS


407

REFERENCES

Aas, G. (1965) A study of the effect of vane shape and rate


of strain on the measured values of in situ shear
strength of clays. Proc. 6th Int. Conf. on Soil Mech.
Toronto. Vol 1, 141-145

Aas, G. (1969) Vane tests for investigation of anisotropy of


undrained shear strength of clays. Proc. Geotech.
Conf. Oslo. Vol 1, 3-8

Adams, J.I, and Radhakrishna, H.S. (1970) Loss of strength due


to sampling in a glacial lake deposit. Proc. ASTM
Symposium on Sampling of Soil and Rock. Toronto.
109-120

Akeroyd, A.V. (1972) Archaelogical and historical evidence for


subsidence in Southern England. Phil. Transc. Royal
Soc. London, A272, 151-169

Atkinson, J.H. (1973) The deformation of undisturbed London


clay. PhD Thesis, University of London

Atkinson, J.H. (1975) Anisotropic elastic deformations in


laboratory tests on undisturbed London clay. Geotech-
nique, Vol 25 No 2, 357-374

Berre, T. and Bjerrum, L. (1973) Shear Strength of Normally


Consolidated Clays. Proc. 8th Int. Conf. Soil Mech.
Moscow. 1.1, 39-49

Bishop, A.W. and Henkel, D.J. (1953) Pore pressure changes


during shear in two undisturbed clays. Proc. 3rd
Int. Conf. on Soil Mech. Zurich, Vol 1, 302-308
408

Bishop, A.W. and Eldin, G. (1953) The Effect of stress history


on the relation between q) and porosity in sand. Proc.
3rd Int. Conf. on Soil Mech. Zurich,
Vol 1, 100-105

Bishop, A.W. and Henkel, D.J. (1957 and 1962) The measurement
of soil properties in the triaxial test. Edward Arnold,
London

Bishop, A.W. (1958) Test requirements for measuring the


coeficient of earth pressure at rest. Proc. Brussels
Conf. on Earth Pressure Problems, Vol 1, 2-14

Bishop, A.W.,Webb, D.L. and Lewin, P.I. (1965) Undisturbed


samples of London clay from the Ashford Common shaft:
strength-effective stress relationships. Geotechnique
15, 1-31

Bishop, A.W. (1966) The strength of soils as engineering


materials. Sixth Rankine Lecture. Geotechnique 10:2,
89-129

Bishop, A.W., Green, G.E. and Skinner, A.E. (1973) Strength and
deformation measurements on soils. 8th Int. Conf.
Soil Mech. Moscow. Vol 1.1, 57-64

Bishop, A.W. and Wesley, L.D. (1975) A Hydraulic Triaxail


Apparatus for Controlled Stress Path Testing. Geotech-
nique 25 (to be published)

Bjerrum, L. and Kenney, Tx_ (1967) Effect of structure on


the shear behaviour of normally consolidated quick
clays. Proc. Geotech. Conf. Oslo. Vol 2, 19-27

Bjerrum, L. (1967) Engineering geology of Norwegian normally


consolidated' marine clays as related to settlement of

buildings. 7th Rankine Lecture. Geotechnique 17, 81-113


409

Bjerrum, L. (1972) Embankments on Soft Ground. Proc. A.S.C.E.


Specialty Conf. on Performance of Earth and Earth-
Supported Structures, Purdue University, Vol 2, 1-54.
Vol 3, 111-159

Bjerrum, L. (1973) Problems of soil mechanics and construction


on soft clays. Proc. 8th Int. Conf. on Soil Mech.
Moscow.

Bjerrum, L. and Anderson, K.H. (1972) In situ measurement of


lateral pressures in clay. Proc. 5th European Conf.
Soil Mech. Madrid. Vol 1, 11-20

Boehler, J.P. and Giroud, J.P. (1971) Measurements of Soil


Anisotropy. Proc. Ninth Annual Symposium on Engineering
Geology and Soil Engineering. Idaho. 175-187

Brand, E.W., Chai Muktabhant, and Somchai Akrapongpisai (1972)


Comparative study of undrained strength measurements of
a soft clay. Proc. 3rd Southeast Asian Conf. on Soil
Engineering, Hongkong. 199-204

Butler, F.G. (1974) Heavily over-consolidated clays. Proc.


Conf. on Settlement of Structures, Cambridge, Pentech
Press, 531-578

Conlon, R.J. and Isaacs, R.M.F. (1970) Effect of sampling and


testing techniques on the shear strength of a glacial-
lacustrine clay from Welland, Ontario. Proc. ASTM
Symposium on Sampling of Soil and Rock. Toronto.
109-120

Davies, P. (1975) Creep Characteristics of Three Undisturbed


Clays. PhD Thesis, University of London
410

De Lory, F.A. and Lai, H.W. (1971) Variation in undrained


shearing strength by semiconfined tests. Canadian
Geotech. Journal 8, 538-545

Duncan, J.M. and Seed, H.B. (1966) Strength variation along


failure surfaces in clay. ASCE Journal Soil Mech.
and Found. Eng. 92 SM6, 81-104

Evans, G. (1965) Intertidal flat sediments and their environ-


ments of deposition in the Wash. Quart. Journal of the
Geological Soc. London. Vol 121, 209-245

Greensmith, J.T. and Tucker, E.V. (1973) Holocene Transgression


and Regressions on the Essex Coast, Outer Thames
Estuary. Geologie en Mijnbouw, Vol 52 (4), 193-202

Henkel, D.J. and Wade, N.H. (1966) Plane strain tests on a


saturated remoulded clay. ASCE Journal Soil Mech.
and Found. Eng. Vol 92, SM6, 67-80

Ladd, C.C. and LaMbe, T.W. (1963) The strength of undisturbed


clay determined from undrained tests. ASTM-NRC
Symposium on Laboratory Shear Testing of Soils, Ottawa,
Canada

Ladd, C.C. (1965) Stress-strain behaviour of anisotropicaily


consolidated clays during undrained shear. Proc. 6th
Int. Conf. Soil. Mech. Vol 1, 282-286

Ladd, C.C. and Varallyay, J. (1965) The influence of stress


system on the behaviour of saturated clays during un-
drained shear. M.I.T. Research Report R65-11, Soils
Publication No 177

Ladd, C.C. and Foott, R. (1974) New design procedure for


stability of soft clays. ASCE Journal of Geotech. Div.

Vol 100, GT7, 763-786


411

Lambe, T.W. (1964) Methods of estimating settlement. ASCE


Journal Soil Mech. and Found. Div. Vol 90, SM5, 43-67

Lambe, T.W. (1967) Stress path method. ASCE Journal Soil •


Mech. and Found. Div. Vol 93, SM6, 309-331

Lo, K.Y. (1965) Stability of slopes in anisotropic soils.


ASCE Journal Soil Mech. and Found. Eng. Vol 91, SM4,
85-106

Lo, K.Y. (1966) Discussion on above paper. ASCE Journal Soil


Mech. and Found. Eng. Vol 92, SM4, 77-82

Lo, K.y..and Milligan, V. (1967) Shear strength properties of


two stratified clays. ASCE Journal Soil Mech. and
Found. Eng. Vol 93, SM1, 1-15

Lo, K.Y. and Morin, J.P. (1972) Strength anisotropy and time
effects on two sensitive clays. Canadian Geotech.
Journal 9, No 3, 261-277

Lo, K.Y. and Holt, R.T. (1974) Directional variation in un-


drained shear strength and fabric of Winnepeg upper
brown clay. Canadian Geotech. Journal, 11.3, 430-437

Marsland, A. (1973) Laboratory and in-situ measurements of


the deformation moduli of London clay. Building
Research Establishment Current Paper 24, 1973

Massarsch, K.R., Holtz, R.D., Holm, B.G. and Frederikson, A.


(1975) Measurement of horizontal in situ stress. Proc.
ASCE/GED Specialty Conf. on in situ measurement of
soil properties, Raleigh North Carolina, June 1975

Parry, R.H.G. (1960) Triaxial compression and extension


tests on remoulded saturated clay. Geotechnique,
Vol 10, No 4, 166-180
412

Parry, R.H.G. (1968) Field and laboratory behaviour of a


lightly overconsolidated clay. Geotechnique, Vol 18,
No 4, 151-171

Parry, R.H.G. (1970) Overconsolidation in soft clay deposits.


Geotechnique 22, 442-445

Parry, R.H.G. (1971) Undrained shear strength in clays. 1st


Aust.-New Zealand Conf. on Geomechanics. Vol 1, 11-15

Parry, R.H.G. and Nadarajah, V. (1974) Anisotropy in a natural


soft clayey silt. Engineering Geology Vol 8, No 3,
287-309

Parry, R.H.G. (1971) Stability analysis for low embankments


on soft clays. Proc. Roscoe Memorial Symposium,
643-668

Pilot, G. (1972) Study of five embankments on soft soils.


Proc. ASCE Specialty Conf. on Performance of Earth
and Earth Supported Structures, Purdue University,
Vol 1, 81-100

Pugh, R.S. (1976) Strength and deformation characteristics


of a soft alluvial clay under full scale loading con-
ditions. PhD Thesis, University of London (in
preparation)

Raymond, G.P. (1973) Foundation failure of New Liskeard em-


bankment. Highway Research Record No 453, 1-17

Raymond, G.P., Townsend, D.L. and Lojkasek, M.J. (1971)


The effect of sampling on the undrained soil properLies
of a Leda soil. Canadian Geotech. Journal. Vol 8, 546-557
413

Schjetne, K. (1971) The measurement of pore pressure during


sampling. N.G.I. Publication No 94

Simons, N.E. (1958) Laboratory determinations of the coeficient


of earth pressure at rest. Proc. Brussels Conf. on
Earth Pressure Problems, Vol 1, 2-14

Simons, N.E. (1963) The influence of stress path on triaxial


test results. ASTM-NRC Symposium on Laboratory Shear
Testing of Soils, Ottawa, Canada. 270-278

Simons, N.E. (1974) Normally consolidated and lightly over-


consolidated cohesive materials. Proc. Conf. on Settle-
ment of Structures, Cambridge, Pentech Press, 500-531

Skempton, A.W. (1948) The cp. = 0 analysis of stability and its


theoretical basis. Proc. 2nd Int. Conf. Soil Mech.
Vol 1, 72-78

Skempton, A.W. and Henkel, D.J. (1953) The post glacial clays
of the Thames estuary at Tilbury and Shellhaven. Proc.
3rd Int. Conf. Soil Mech. Vol 1, 302-308

Skempton, A.W. and Sowa, V.A. (1963) The behaviour of saturated


clays during sampling and testing. Geotechnique Vol 13,
No 4, 269-290

Taylor, D.W. (1948) Fundamentals of Soil Mechanics. John Wiley


and Sons. p. 397

Vaid, Y.P. and Campanella, R.G. (1974) Triaxial and plane


strain behaviour of natural clay. ASCE Journal Geotech.
Div. Vol 100, GT3, 207-224

Wallbancke, H.J. (1968) Holocene sea levels in southern


Britain. MSc (Eng) Report, University of London
(Imperial College of Science and Technology) July, 1968
414

Wiesel, C.E. (1973) Some factors influencing in situ vane test


results. Proc. 8th Int. Conf. Soil Mech. Vol 1, 475-479

Wroth, C.P. and Hughes, J.M.O. (1973) An instrument for the in


situ measurement of the properties of soft clays.
Proc. 8th Int. Conf. Soil Mech. Vol 1, 487-494

Zeevaert, L. (1953) Discussion on Session 1: Theories and


Hypotheses of a General Character, Soil Properties,
Soil Classification, Engineering Geology. Proc. 3rd
Int. Conf. Soil. Mech. Vol 3,. 113-114

You might also like