THE EFFECT OF STRESS PATH ON LONDON CLAY Volume I A …

443
THE EFFECT OF STRESS PATH ON THE DEFORMATION AND CONSOLIDATION OF LONDON CLAY Volume I A Thesis Submitted to the University of London (Imperial College of Science and Technology) fbr the degree of Doctor of Philosophy in the Faculty of Engineering by Nitindra=Nath Som, B.C.E., D.I.C. Imperial College September 1968

Transcript of THE EFFECT OF STRESS PATH ON LONDON CLAY Volume I A …

THE EFFECT OF STRESS PATH ON

THE DEFORMATION AND CONSOLIDATION OF

LONDON CLAY

Volume I

A Thesis Submitted to the

University of London

(Imperial College of Science and Technology)

fbr the degree of

Doctor of Philosophy in the Faculty of Engineering

by

Nitindra=Nath Som, B.C.E., D.I.C.

Imperial College September 1968

u nothing takes place without sufficient reason, that is to say, nothing happens without its being possible for one who knows things sufficiently, to give a reason which is sufficient to determine why things are so and not otherwise"

Gottfried Wilhelm Leibniz (1646-1716)

1.

ABSTRACT

Theoretical and experimental investigations have been carried

out to study the influence of stress path on the deformation and con-

solidation of over-consolidated clays - with particular emphasis on

London clay - in relation to settlement of structures. It is shown

that for a proper understanding of the deformation of a soil beneath

a foundation the soil should be tested in the laboratory under the

same set of effective stresses that it will undergo in the field and

that the influence of lateral stresses carrot be ignored.

The stresses and displacements in non-homogeneous soil media,

beneath circular and strip footings, have been calculated from

Gibson's analytical solutions. A numerical method is suggested for

determining the "immediate" (elastic) settlement of structures founded

on a medium whose modulus of elasticity varies with depth.

The influence of lateral stressesonthe deformation charact-

eristics of undisturbed London clay has been studied from both oedo-

meter and triaxial tests while the stress path for one-dimensional

compression is determined from specially designed oedometers. The

effect of small pressure increments on the compressibility of London

clay is also studied in the oedometer.

The experimental results are examined in the light of their

influence on the settlement of structures and a method of settle-

ment analysis is proposed that takes into account the stress path

2.

of the elements of soil beneath a foundation: comparisons are made

with the existing methods of analysis.

The pre-consolidation pressures of London clay are determined

from the stress deformation characteristics of samples loaded to

high effective stresses.

3.

ACKNOWLEDGEMENTS

The work described in this thesis was carried out in the

Soil Mechanics section of the Department of Civil Engineering,

Imperial College of Science and Technology, University of London,

under the general supervision of Professor A. W. Bishop, whose in-

terest and help throughout the work is gratefully acknowledged.

The author expresses his deep gratitude to Dr. N. E. Simons

who directed the research programme and gave active help and en-

couragement not only in all aspects of the work but also in overcoming

many problems the author encountered throughout the course of the

work.

Sincere thanks are extended to Professor R. E. Gibson of

King's College, London, for his invaluable assistance in connection

with the theoretical work on stress distribution described in Chapter

6. The author is grateful to Mr. A. E. Skinner and Mr. G. E.

Green for their help in surmounting numerous practical difficulties

which inevitably arose during the experimental work and to Dr. N.

R. Morgenstern for taking part in many instructive discussions.

The author wishes to thank his colleagues, in particular

Mr. H. T. Lovenbury, Dr. S. K. Sarma, Dr, P. L. T. Phukan and Mr.

I. C. Pyrah who assisted in various ways during the course of the

work and with whom many fruitful discussions were held. Sincere

4.

appreciation is also expressed to Messrs. D. T. Evans, L. F. Spall,

E. W. Harris and F. D. Evans for their active help with the laboratory

work and for their sympathy and understanding in general.

The author is further thankful to Dr. and Mrs. N. E. Simons

for their affectionate hospitality on the occasions he had the

opportunity to visit them at their home.

The arduous task of typing the often unreadable manuscript

was undertaken by Miss E. Hamilton and the author can only express

his admiration for the patience she has shown throughout.

The work was made possible by a generous grant from the

Construction Industries Research and Information Association to whom

the author is grateful.

TABLE OF CONTENTS Page

ABSTRACT 1

ACKNOWLEDGEMENTS 3

TABLE OF CONTENTS 5

CHAPTER 1 - INTRODUCTION 9

CHAPTER 2 - BRIEF REVIEW OF PAST WORK 17

2.1 Influence of Stress Path on the Deformation of Saturated Clays 17

2.2 Methods of Settlement Analysis 25

CHAPTER 3 - A STUDY OF CASE RECORDS OF SETTLEMENT OF STRUCTURES ON CLAY 35

3.1 General Definitions 35

3.2 Structures on Over-Consolidated Clays 37

3.3 Structures on Normally-Consolidated Clays 42

3.4 Summary and Discussion 46

CHAPTER 4 - LONDON CLAY - GEOLOGY AND STRESS HISTORY 54

4.1 Geology of the London Basin 54

4.2 Previous Work on London Clay 57

4.3 Index Properties 59

4.4 States of Stress In-situ 60

CHAPTER 5 - STRESS PATH 66

5.1 The Concept of Stress Path 66

5.2 Definitions 67

5.3 Stress Paths in Laboratory Triaxial Tests 67

5.4 Stress Path in the Field due to Foundation Loading 69

5.5 Stages of an "Ideal" Experimental Programme for Settlement Analysis 73

5.6 Methods of Settlement Analysis 74

6. Page

CHAPTER 6 - STRESS DISTRIBUTION IN SOIL MEDIA 78

Soil as a Homogeneous, Isotropic, Elastic Medium 79

6.3 Non-Homogeneity in Soils 81

6.4 Two-Layer Systems 82

6.5 Three-Layer Systems 86

6.6 Multi-Layer Systems 88

6.7 Non-Homogeneous Medium 89

6.8 Non-Linear Soil Medium 107

6.9 Summary 109

CHAPTER 7 - STRESSES DURING CONSOLIDATION IN THE FIELD

117

-7.1 Development of Pore Pressures in Saturated Clay

117

7.2 Pore Pressures Beneath a Circular Foundation 120 Stress Changes During Consolidation

122

CHAPTER 8 - SAMPLING, PRELIMINARY MEASUREMENTS AND EXPERIMENTAL PROGRAMME 127

8.1 Location of Sites 127 8.2 Description of Sites, Sampling and Storage 127 8.3 Index Properties 129 8.4 Moisture Content 130 8.5 Stresses in the Ground and After Sampling 131 8.6 Experimental Programme 138 8.6.1 Oedometer Tests 139 8.6.2 Triaxial Tests 142

CHAPTER 9 - EQUIPMENT AND PROCEDURES OF TESTING 151

9.1 Oedometer Tests 151 9.1.1 Standard Oedometers 151 9.1.2 High Pressure (Hydraulic) Oodometer 154 9.1.3 Controlled Rate of Strain Oedometer 161 9.1.4 Oedometers Fitted with Strain Gauges 164 9,2 Triaxial Tests 168

6.1 Introduction , 6.2

78

7. Page

CHAPTER 10

10.1 10.1.1 10.1.2 10.1.3 10.1.4 10.1.3 10.1.6 10.2

10.2.1 10.2.2

10.2.3 10.3 10.3.1 10.3.2 10.3.3 10.3.4 10.4 10.4.1 10.4.2

10.4.3 10.4.4

- RESULTS OF OEDOMETER TESTS 181

Tests in the Standard Oedometer 181 Determination of Swelling Pressures 181 Time-Settlement Relationships 184 Pressure - Void Ratio Relationships 185 Compressibility Characteristics 188 Coefficient of Consolidation 190 Discussion of Results 191 Tests in the High Pressure (Hydraulic)

Oedometer 198 Presentation of Results 198 Discussion of Results 202 (a) Initial Response of Pore Pressures :(b) Measurement of Strains (o) Consolidation Characteristics (d) Side Friction Advantages of the Hydraulic Oedometer 212 Controlled Rate of Strain Tests 215 General 215 Pressure Void. ratio Relationships a17 Compressibility Characteristics 220 Determination of Pre-Consolidation Pressures 224 Tests in Oedometers fitted with Strain Gauges 234 Introduction 234 Variation of Lateral Stresses during Con- solidation 237

Stress Paths 239 Discussion of Results 242

CHAPTER 11 - RESULTS OF TRIAXIAL TESTS 259

Presentation of Data 259 Shear Strength Parameters 265 Deformation under Undrained Conditions 267 Pore Pressure Parameters A and B 276 Volume Change Characteristics 285 Volumetric Strains 285 Axial Strains 296 Elastic Parameters of London Clay 305 Rate of Consolidation 309

CHAPTER 12 - A COMPARATIVE STUDY OF THE OEDOMETER AND TRIAXIAL TEST DATA 323

12.1 Volumetric Compressibility 323

8. Page

12.2 Axial Strains 326 12.3 Rate of Consolidation 332

CHAIT ER 13 - THE STRESS PATH METHOD OF SETTLEMENT ANALYSIS 333

13.1 Introduction 333 13.2 Formulation of the Problem 333 13.3 Distribution of Stresses 334 13.4 "Immediate" Settlement 336 13.5 Consolidation Settlement 338 13.6 Comparison of the Different Methods of

Settlement Analysis 344 13.7 Rate of Settlement 345

CHAPTER 14 - CONCLUSIONS AND SUGGESTIONS FOR FURTHER RESEARCH 360

14.1 Conclusions 360 14.2 Suggestions for Further Research 365

APPENDIX A 367

APPENDIX B 370

APPENDIX C 375

APPENDIX D 378

APPENDIX E 383

APPENDIX F 384

REFERENCFq 387

ILLUSTRATIONS Volume II

9•

CHAPTER 1

INTRODUCTION

Settlement analysis of engineering structures has become

an integral part of Civil Engineering practice for many years but

the method of systematic analysis has undergone but little change

since its inception over forty years ago following Terzaghi's

classic work on the theory of consolidation. The soundness of the

basic principle (i.e. the principle of effective stress) cannot, of

course, be doubted and in many cases, the method has proved adequate

in predicting, at least for practical purposes, the settlement of

structures on a wide variety of clays. Nevertheless, conditions in

the field often differ from the simplified assumptions that have to

be made for the analysis (e.g. one-dimensional strain) which some-

times result in the over-estimation of the magnitude of settlement

and under-estimation of the rate of settlement. This is particularly

true of structures on over-consolidated clays, a phenomenon which was

recognised by Terzaghi as far back as 1936 when he stated, in opening

the discussion on settlement of structures at the First International

Conference on Soil Mechanics, "it appears that the important differ-

ences between theory and reality are limited to those cases in which

the loaded material was intensely pre-compressed at some stage of

its geological history" (Terzaghi 1936a). It was not until Skempton

and Bjerrum's important contribution in 1957 that some improvement

10.

was achieved (Skempton and Bjerrum 1957).

While the Skempton and Bjerrum approach to settlement

analysis is certainly a step in the right direction - it recognises

that the foundation soil undergoes lateral deformation during load

application and that the subsequent consolidation is a function of

the excess pore pressures set up in the clay - it still assumes that

during consolidation the clay is laterally restrained. Although

this condition is approximately true in certain problems, such as

that of a thin layer of clay lying between beds of sand, it is more

often than not that lateral deformations can occur in the field -

consider, for example, the case of a structure founded on a thick

bed of clay. Whether such lateral strains will influence the settle-

ment to any great extent depends on the stress changes during con-

solidation and their influence on the deformation properties of the

soil. Therefore, to make successful use of laboratory data to pre-

dict the deformation of a soil under a given set of stresses it is

necessary to test the soil, applying, as closely as possible, the

same stress changes as those to which the material will be subjected

in nature. Moreover, the behaviour of soils being essentially non-

linear the deformation properties, such as the Young's modulus,

Poisson's ratio and compressibility, will vary with stress level, and,

for this reason, it is also desirable that the soil after sampling

be first brought back to the stresses prevailing in the ground before

subjecting it to the stress changes it is likely to undergo on load-

11.

ing (see Moretto 1965). Thus the concept of stress path logically

comes into the picture of soil deformation in the field and it is

this aspect of the behaviour of over-consolidated clays that is

studied in this thesis. The experimental work has been carried out

on undisturbed London clay from two sites - Oxford Circus in Central

London and Ongar in Essex.

The thesis begins, after a brief review of past work, with

a study of case records of settlement of structures on both normally

consolidated and over-consolidated clays, from which certain differ-

ences in the respective behaviour of the two types of clay become

immediately apparent. The remainder of the thesis is devoted to

examining in detail the deformation characteristics of over-

consolidated clays in the light of the influence of stress paths,

with particular emphasis given to the behaviour of London clay.

It is recognised that a knowledge of the geology and stress

history of a soil is essential in the understanding of its properties

and Chapter 4, therefore, gives a brief summary of the previous

history of London clay and its relevance to foundation problems.

Chapter 5 seeks to explain the stress path of an element of soil

beneath a foundation both during load application and subsequent

consolidation. When this is compared with the stress paths implied

in the existing methods of settlement analysis some of the causes of

the discrepancy between calculated and observed settlements become

apparent.

12.

A major requirement of settlement analysis is the calcula-

tion of stresses in the soil medium due to the applied foundation

pressure. This is required not only to calculate the "immediate"

(i.e. the end of construction) settlement but also to determine the

distribution of initial excess pore pressures which subsequently

dissipate to cause the time - dependent consolidation settlement.

It is customary to use the Boussinesq analysis for an isotropic,

homogeneous, elastic medium for the purpose, but a real soil is

neither elastic nor homogeneous. Even considering small strains

within which linearity of stress - strain relationships can be

assumed to be valid, the modulus of elasticity in general varies

with depth either as a consequence of increasing effective stresses

with depth or due to the presence of different geological formations

possessing unlike engineering properties. In Chapter 6, a survey

is made of the available data of the distribution of stresses in

layered elastic media, while Gibson's analytical solution (Gibson

1967) for non-homogeneous elastic medium has been used to calculate

numerical results for circular and strip footings. A short section

is also devoted to study the effect of non-linearity of stress -

strain relationships on the distribution of stresses. Chapter 7

is concerned with the development of pore pressure in a soil medium

and the changes of effective stresses during consolidation which

have an important bearing on the settlement.

When an undisturbed sample is removed from the ground for

testing in the laboratory there exists in the sample a state of

stress which is completely different from that to which it was

subjected before sampling. An analysis of this has been made in

Chapter 8, following Skempton and Sowa (1963). The importance is

pointed out of recognising this state of stress when determining

the stress path that should be followed to reproduce, as closely

as possible, the stress path the element is likely to undergo in

the field. Chapter 8 also contains a full description of the

experimental programme.

A number of new instruments were built to study the re-

levant properties of London clay. These and all the other equip-

ment that have been used in connection with the experimental work

are described in Chapter 9.

The results of the different series of oedometer tests

are presented in Chapter 10, the main body of which can be divided

into four parts:

(a) The effect of rest period and the magnitude of load in-

crements on compressibility was studied in the standard Bishop type

oedometer. It is easily understood that the stress increases

caused by a building load decreases with increasing depth beneath

the base of the foundation. Yet it is common practice to test

samples in the laboratory using the pressure increment ratio of 1,

the results of which may not be directly applicable for small load

increments in the field. Moreover, the inevitable disturbance and

13.

ik.

the release of pressures caused by sampling may result in a break-

down of the original structure of the clay which may not be regained

when large pressure increments are applied at short intervals. In

order to study these effects specimens were allowed to rest at the

in-situ overburden pressure for up to 90 days and then subjected to

load increments of different magnitudes.

(b) The influence of controlled rate of strain as opposed to

the conventional step loading, on the compressibility of London clay

was studied in a specially designed oedometer. The opportunity was

also taken to study the deformation of London clay at effective

stresses of up to 7,000 lbs/in2 from which it has been possible to

estimate the pre-consolidation pressure of London clay at Ongar,

Essex, and Wraysbury, Middlesex.

(c) A newly built high pressure oedometer was used to study

some consolidation characteristics of undisturbed London clay. In

this apparatus, the specimens were loaded by hydraulic pressure and

settlement and pore pressure were measured over a small part of the

surface thus keeping the influence of side friction on this measure-

ment to a minimum.

(d) The stress path during one-dimensional consolidation of

undisturbed London clay was determined in three specially built

oedometers which were fitted with strain gauges on the outside to

measure the lateral stress.

The programme of triaxial testing was devised firstly, to

15.

determine the stress - strain modulus and the pore pressure para-

meters for the range of stresses normally encountered in practice.

For this, specimens were first brought back to the estimated in-

situ state of stress and then subjected to stress increments under

undrained conditions. The influence of stress path on the axial

and volumetric strains during consolidation have been studied from

triaxial consolidation tests. Although drainage was permitted only

in one direction the specimens were allowed to deform laterally as

well as vertically thus removing the lateral restraint which is

characteristic of oedometer tests. It has thus been possible to

investigate the possible effects of the lateral stresses on the

deformation of the clay during consolidation. The data are given

Chapter 11.

A comparative study of the oedometer and triaxial test

data reveal some important deficiencies of the existing methods of

settlement analysis in which the influence of the lateral stresses

is completely ignored. It is seen that the consolidation settle-

ment is largely dependent on the stress path which, in the field,

may differ considerably from that for laboratory oedometer tests.

Direct application of the oedometer test results may, therefore,

give an incorrect estimate of the settlement of a structure in the

field.

In Chapter 13 a method of settlement analysis, which takes

account of the influence of stress path, is developed on the basis

16.

of the experimental evidence obtained. This method which may be

called the "Stress Path Method" (Lambe 1964, 1967) is found to give

estimates of settlement which are very different from those given

by existing methods of analysis. On the subject of the rate of

settlement, one-dimensional theory of consolidation, as is well

known, is again found to be grossly inadequate. Although

mathematical treatment of three-dimensional consolidation did not

form a part of this thesis some available results, obtained from

approximate numerical analyses, have been used to calculate the

rate of settlement under three-dimensional conditions.

17. CHAPTER 2

BRIEF REVIEW OF PAST '1ORK

The stress - deformation characteristics of soils have

been the subject of many investigations, but in this brief review

it will be possible to consider only those which are of direct

relevance to the work described later in the thesis. Attention

will, therefore, be directed, firstly to the contributions that

relate deformation and pore pressures in saturated clay to stress

changes under drained and undrained conditions and, secondly, to

the methods of settlement analysis that are in use at the present

time.

2.1 Influence of stress ath on the deformation of saturated

clays.

Most of the work reported in the literature on this subject

are parts of wider studies on the shear strength of soils. Our

consideration will primarily be restricted, however, to the stress

- deformation behaviour of saturated clays when subjected to ex-

ternal pressures under drained and undrained conditions.

Modern studies on the deformation of soils began more than

forty years ago with Terzaghi's class work on the consolidation of

clays (Terzaghi 1923) and are based on the by now well established -

18.

principle of effective stress! (For a summary of Terzaghits early

work between 1921 and 1925, see Skempton 1960b). Rendulic (1936,

1937) considered the general problem of the void ratio - effective

stress relationships of "isotropic" clay and came to the conclusion

that, for a given clay with a given initial condition, void ratio

was an exclusive function of the three principal effective stresses.

He carried out a series of drained and consolidated undrained tri-

axial tests on remoulded Wiener Tegel and showed that there existed

a unique relationship between water content and effective stresses

from which it was possible to predict the effective stress and hence

the pore pressure at any point in an undrained test. Later work by

Henkel (1958, 1960, 1960a) has demonstrated the soundness of Ren-

dulicfs conclusion. From an extensive study of drained and un-

drained tests on remoulded Weald clay Henkel was able to show that

the behaviour of clays during shear could be represented by a series

of stress paths, each associated with a particular water content, on

the principal stress space.

Rutledge, in the Triaxial Shear Report (1947), presented

what has since been known as the American Hypothesis. From results

-of triaxial and oedometer tests on undisturbed samples of clay it

was concluded that the water content of saturated clay was primarily

a function of the major principal stress and essentially independent

Theoretical considerations of the principle of effective stress are discussed by Bishop (1959) and Skempton (1960a).

19.

of the other stresses. Although later works of Bjerrum (1954),

De Wet (1962), Broth.. and Ratnam (1963), Raymond (1965), Lee and

Farhoomand (1967) have shown that the hypothesis describes the

behaviour of many soils reasonably accurately its validity has not

been found to be general (see below). Moreover, as shown by

Henkel (1958), the hypothesis implies a certain shape of the water

content contours for undrained tests in the principal stress space

which are not supported by experimental evidence. We shall come

back to the American Hypothesis later in the thesis.

In order to predict the behaviour of clay under undrained

conditions it is necessary to have some knowledge of the pore

pressure set up by the application of external load. Skempton

(1948) introduced the 7. theory which was the first of a series

of attempts to relate the pore pressure and the applied stresses

in an undrained test. In terms of the compressibility Cc and

the expansibility Cs, the development of excess pore pressure

was found to be govenned by the equation

Au. C + 1 ( cr — i (173) r.; - + 2

(2.1.1)

where X = Cs/Cc and AT 1 and .d g.3 are the increases of

major and principal stresses. Apart from the assumption of the

validity of elastic theory Skempton also assumed that the Cc and

Cs determined from all round consolidation and swelling tests were

20.

applicable to cases where shear stresses were present. Using the

experimental data, obtained by Hvorslev (1937) and Taylor (1948) on

Weiner Tegel and Boston Blue clay Skempton showed that equation

(2.1) predicted the pore pressure in undrained tests reasonably

accurately. Odenstad (1949) extended Skempton's X theory to

take account of the effect of dilatancy associated with the applicau

tion of shear stresses but made no quantitative check on his theory.

Bjerrum (1954) suggested that the pore pressure in undrained tests

could be related to volume changes in drained tests, on the basis

that the application of axial stress in the undrained test was

identical, with respect to effective stresses, to a drained test

where the axial stress increase was accompanied by a decrease of

cell pressure such that the volume of the sample remained constant.

Thus Bjerrum got the following expression for the excess pore

pressure in terms of the applied stresses,

Au = L1 0-3 + 1

C ( cr - XT s2

ccl

(2.1.2)

where Cc1

and Cs2

are the compressibility corresponding to

axial compression and the expansibility corresponding to decrease

of radial stresses respectively. Although equation (2.2) is

apparently similar to equation (2.1) Bjerrum's approach had the

advantage over the A theory that Gel and Cs2

were the para-

meters determined from relevant drained tests and the need for the

21.

application of Cc and Cs data obtained from all round tests to

the individual stress changes in the shear stage was no longer there.

The next major step forward in this field was the intro-

duction by Skempton (1954) of the pore pressure parameters A and

B, the application of which to stability problems was immediately

demonstrated by Bishop (1954). Although the basic analysis was

done within the framework of the theory of elasticity the final

equation for expess pore pressure,

u= f 6'3 + A( 1 - d 03) J

(2.1.3)

was expressed in terms of parameters which could be readily deter-

mined from triaxial tests. Equation C2;1:3)is extensively used to-

day in the effective stress analysis of Soil Mechanics problems.

Its application in settlement analysis has been demonstrated by

Skempton and Bjerrum (1957). A more general expression for the

excess pore pressure, in terms of the octahedral normal stress and

the octahedral shear stress, has since been proposed (Skempton

1960, Henkel 1960) to take account of the intermediate principal

stress, but its validity has still to be established.

The experimental work to examine the influence of stress

path on the volume change soils has, in general formed part of

shear strength studies, mostly of sands and remoulded clays. The

data on the behaviour of undisturbed clays for small stress changes

22.

are limited. Henkel (1958) observed from tests on remoulded clays

that unique relationships existed between the average principal

effective stress and water content at failure and that the latter

was always lower than the water content at the corresponding

effective stress for all round consolidation. Roscoe, Schoffield

and Wroth (1958) used the concept of critical void ratio to describe

the behaviour of soils at failure, according to which the water

contents at the "end" points of all triaxial tests, when plotted

against the average principal effective stress, lie on a unique

straight line, at which further increase of axial strain causes no

more volume change in the drained test or pore pressure change in

the undrained test (i.e. the soil then continues to deform at con-

stant stress and constant void ratio). From a series of triaxial

and plane strain tests in which specimens were consolidated iso-

tropically as well as under Ko

conditions, Sowa (1963) and Wade

(1963) found that the water content during consolidation of re-

moulded Weald clay was controlled by the average principal effective

stress (see Henkel and Sowa 1963).

On the other hand many research workers have found that

the volume change of soils is governed primarily by the major

principal effective stress, the other stresses exerting little or

no influence; (Triaxial Shear Report 4.947, Bjerrum 1954, De Wet

This point will be described in greater detail in Chapter 11.

23.

1962, Broms and Ratiam 1963, Raymond 1965). Yet again, Akai and

Adachi (1965) found that the volume change of a remoulded Japanese

clay in one-dimensional consolidation was larger than that caused

during isotropic consolidation for the same average effective stress

- a phenomenon which the authors attributed to an extra volume

change caused by the shear stress.

Different hypotheses are, therefore, available to des-

cribe the same phenomenon of soil deformation. All that can be

said, at this point, is that there is no generally applicable

hypothesis and the one that should apply to a particular soil has

to be determined experimentally. An important point to note here

is that, absent from most of the above investigations (a notable

exception is the Triaxial Shear Report) are tests on undisturbed

clays. The applicability of any hypothesis to practical problems,

therefore, still remains to be studied.

The works referred to in the preceding paragraphs des-

cribe the behaviour of soils under axi-symmetric stress conditions

(i.e. two of the principal stresses are equal). In recent years

many investigations have been carried out to study the influence of

the intermediate principal stress on the shear strength of soils.

Herd again not much data are available of the deformation of clays

under small stress changes.

The early works in this field consisted of the plane

strain tests in which the strain along the longitudinal axes of

24.

prismatic samples was kept zero, (Wood 1958, Wade 1963, Cornforth

1964, Henkel and Wade 1967). Other investigators have used hollow

cylindrical specimens where the intermediate principal stress could

be varied by applying different pressures at the inside and outside

of the specimens (Kirkpatrick 1957, Haythornthwaite 1960, Broms and

Ratnam 1963, Wu, Loh and Malvern 1963). Attempts have also been

made to use cubical test specimens and control the three principal

stresses independently (Kjellman 1936, Jakohson 1957, Shibata and

K:rube1965, Ko and Scott 1967).

Only a few of the works mentioned above, however, refer

to the behaviour of clays, the experimental work being conducted

mostly on sands. Moreover, as has been said earlier, it was the

shear strength of soils that was the primary concern of the in-

vestigators. (For a detailed discussion of the failure criteria

of soils and the influence of the intermediate principal stress on

the shear strength of soils, see Bishop 1966). Very little informa-

tion can, therefore, be obtained of the influence of the intermediate

principal stress on the deformation of clays. From undrained tests

on remoulded clays, both Shibata and KarUbe(1965) and Henkel and

Wade (1966) have shown that when the intermediate principal stress

lies between the major and minor principal stresses common stress

paths are obtained if the data are plotted on the octahedral normal

stress - octahedral shear stress space. Their test data, as well

as those of Wu, Loh and Malvern (1963), indicate that the pore

25.

pressure set up under undrained conditions can also be expressed in

terms of the octahedral normal stress and the octahedral shear

stress. The data on the consolidation of clays under independently

controlled stresses are even more scarce. Broms and Ratnamts

study with hollow cylindrical specimens shows that the water content

of remoulded Kaoline is a function of the major principal stress

and independent of the minor and intermediate principal stresses

(Broms and Ratnam 1963).

2.2 Methods of settlement analysis

(A) Determination of the magnitude of settlement

The principl& of effective stress and Terzaghi's theory

of one-dimensional consolidation have been the essential basis for

all settlement analyses of structures founded on clay. The

earliest method of analysis, expressed by the equation

Pc my .

(3"Z . d z

(2.2.1)

where mv is the compressibility determined from oedometer test

64zis the increase of vertical stress at a depth z and

z is the total thickness of the clay stratum,

was propsed by Terzaghi (1929) for calculating the consolidation

settlement of a layer of clay subjected to lateral confinement,

i.e. where all settlement was due to one-dimensional compression

26.

of the clay stratum. Although this method, which has been called

the "conventional" method by Skempton, Peck and McDonald (1955), is

valid only in cases where the condition of no lateral strain is at

least approximately true (see Skempton and Bjerrum 1957) it has

been extended to cases in the field where the foundation rests on

a deep bed of clay (Taylor 1948). In such cases there are lateral

deformations during load application which give rise to what is

known as the "immediate" settlement. Therefore, although the

"conventional" method has in many instances, given reasonably good

predictions of the total settlement - particularly for structures

on normally consolidated clays (see Terzaghi 1936a) - it is

physically inadequate to describe completely the behaviour of a

clay that undergoes important lateral deformations.

It has been common practice for many years to calculate

the "immediate" settlement, i.e. the settlement that takes place

under the condition of no volume change and is a result of the

shear deformation of the clay, from the standard equations of the

theory of elasticity (Terzaghi 1943, see also Scott 1963 and Harr

1966) which has the following general expression,

qn . B I Q 2 I

E (2.2.2)

where qn is the net foundation pressure

B is a suitable dimension of the foundation

27.

E is the Young's modulus of the clay

") is the Poisson's ratio and

Ip is the influence coefficient whose magnitude depends

on the geometry of the problem.

Whereas equation (2.2.2) still remains the most widely

used expression for calculating the "immediate" settlement, the

method of calculating the consolidation settlement has undergone

some modifications. It has been assumed that equation (2.2.1)

gives the total (immediate + consolidation) settlement of a structure

from which the "immediate" settlement (equation 2.2.2) is sub-

tracted to obtain the consolidation settlement (Skempton, Peck and

McDonald 1955, Skempton and McDonald 1956). Although this approach

is somewhat empirical* it has given good agreement between calculated

and observed settlements of structures on a wide variety of soils

(Skempton and McDonald 1955). This may have been due to opposing

errors in the analysis cancelling each other out,for example, if

equation (2.2.1) over-estimates the consolidation settlements

because the excess pore pressures in the field are less than the

increases in vertical stresses, it does not take account of the

shear deformation.

A major improvement in the analysis of settlement was

For a criticism of this method, see Alderman (1956) and Mayerhof (1956) who suggested that the total settlement should be given by adding the "immediate" and consolidation settlements obtained separately from equations (2.2.2) and (2.2.1) respectively.

28.

achieved by Skempton and Bjerrum (1957) who recognised that there

was lateral deformation in the clay during load application and

that the consolidation settlement was a function of the excess

pore pressures set up by the applied load. Taking account of the

shear stresses, therefore, a new expression for the total settle-

ment was given,

/f= P-

I J oed (2.2.3)

where fk is a factor which depends on the pore pressure parameter

A and the geometry of the foundation and coed is equal to the

consolidation settlement obtained by the straightforward application

of the oedometer test results (equation 2.2.1). A critical study,

in terms of stress paths, of the methods of settlement analysis des-

cribed above will be made in Chapter 5.

Till now the oedometer test has been the cornerstone of

all settlement analyses, although the stress - strain relationships

from undrained triaxial tests are used to determine the Young1s

modulus (E), required for the calculation of the immediate settle-

ment (see Chapter 11 for a discussion of the factors that influence

the value of E). It was realised long ago, however, that certain

factors had to be taken into account if the oedometer test results

could be successfully applied to field problems. Casagrande (1936)

noted the significance of the pre-consolidation pressure and pro-

29.

posed an approximate method of determining this from laboratory

tests. This work was an important step in improving the relation-

ship between laboratory data and field behaviour. Further pro-

gress along this line was achieved by Schmertmann (1953) who

suggested an empirical method of determining the in-situ consolida-

tion behaviour of clays, from laboratory tests on samples with

various degrees of disturbance. Other factors that influence the

consolidation behaviour of clays, such as the rate of loading and

the magnitude of pressure increment have been pointed out by

Langer (1936), Terzaghi (1941), Leonard and Ramiah (1959), Crawford

(1964), Leonard and Altschaeffl (1964) and others - for a more

detailed account see Chapter 10 - but these factors have not

generally been taken into consideration in settlement analysis.

It has been mentioned already that the oedometer test

results are almost universally used today for all settlement

analyses; little use has so far been made of the triaxial test.

Of course, in situations where volume change of the clay during

consolidation results in one-dimensional strain direct use of the

oedometer test data will give satisfactory. But stress changes

during consolidation in the field are often such that the volume

change of the clay is accompanied by significant lateral strain,

in which case the assumption of one-dimensional strain would be

erroneous: To take this into consideration T. W. Lambe, in a

An early laboratory study of the influence of lateral stresses on the consolidation settlement was made by Hruban (1948).

30.

series of important contributions (Lambe 1964, 1965, 1965a, 1967)

has put forward the "Stress Path Method" of settlement analysis.

According to this, an "average element" beneath a foundation is

first located from a preliminary investigation. This "average

element" is then sampled and a number of laboratory triaxial tests

are performed, duplicating the effective stress path the element

is likely to undergo in the field during loading and consolidation.

The measured axial strain from the laboratory tests, multiplied by

the thickness of the clay layer, then gives the total settlement

which can, of course, be separated, if desired, into the "immediate"

and consolidation settlements. In this way the actual deformations

caused by the appropriate stresses beneath a foundation can be

evaluated. Although the assumption is made that the stresses re-

main unchanged during consolidation* (see Chapter 5) and there may

be practical difficulties in selecting the "average element", it is

the present author's view that Lambe's approach is a significant

advance in Civil Engineering analysis - it certainly gives a better

insight into the mechanics of soil deformation in the field.

Davis and Poulos (1963, 1966, 1968), in a series of

papers, have proposed a similar method of settlement analysis, based

on the triaxial test and elastic stress distribution. Essentially,

it consists of determining the Poisson's ratio (required for stress

This, of course, is to simplify the theoretical and experimental work involved, and not an inherent fault in the method.

31.

analysis) of the clay from triaxial tests - for the appropriate

stress range - and then subjecting representative samples from

various depths to stress changes, in the triaxial apparatus, that

they are likely to undergo in the field. The measured axial

strain multiplied by the thickness of the corresponding layer gives

the settlement and the sum of the settlements for each layer, then

gives the total settlement. A very similar procedure has also

been proposed by Kerisel and Quatre (1968)*. These methods are

more rigorous than Lambe's method in the sense that the stress path

of a number of elements beneath a foundation can be considered.

The difficulty of selecting the "average element" is thus avoided.

Although neither Davis and Poulos nor Kerisel and Quatre take into

account the stress changes that occur in the field during consolida-

tion they have claarly demonstrated the advantages of using the

triaxial test in settlement analyses, both from theoretical and

practical standpoints.

(B) Determination of the rate of settlement

The prediction of the rate of settlement has always been

the most uncertain part of a settlement analysis. Although settle-

ment in the field almost always occurs under three-dimensional con-

ditions, Terzaghi's theory of one-dimensional consolidation (Terzacshi.

Kerisel and Quatre (1968) also give charts for the cal-culation of vertical and horizontal stresses for different shapes (circular, square or rectangular) of rigid•and flexible footings, for all values of Poisson's ratio.

32.

1929) still remains the basis for the analysis of the rate of settle-

ment, and this often leads to considerable error. Theoretical

study of the rate of settlement has not been undertaken in the work

presented here. Only a brief review will, therefore, be presented

of the major advances that have been achieved in recent years in

studies on consolidation. The theories concerning secondary con-

solidation and creep will be excluded from this review.

The solution of Terzaghi's theory of one-dimensional con-

solidation for a wide range of the initial distribution of pore

pressures and boundary conditions have been given by Terzaghi and

Frftlich (1936) (also Terzaghi 1943). Gray (1945) solved the problem

of consolidation of contiguous clay layers having unlike compressi-

bilities and Gibson (1958) analysed the case. of clay layers varying

in tbickness with time. The variation of permeability and time -

dependent loading were considered by Schiffman (1958) while Abbott

(1960) studied the one-dimensional consolidation of multi-layered

soils. Schiffman and Gibson (1964) presented general solutions for

the one-dimensional consolidation of non-homogeneous clay layers

(i.e. compressibility and permeability varying with depth) and

showed that for structures founded on the upper part of London clay

non-homogeneity alone would cause the rate of settlement to be faster

than that predicted by the Terzaghi theory, even if consolidation

was one-dimensional. Davis and Raymond (1965) modified the theory

of consolidation for non-linear pressure void ratio relationships

33.

while Barden (1965) formulated a new theory to take the variatidos

of both compressibility and permeability during consolidation into

account. (see also Chapter 10). Gibson, England and Hussey

(1967) derived the general theory of one-dimensional consolidation

for a saturated clay layer undergoing large strains, using a form

of Darcy's law in which the relative velocity of the pore water and

the soil skeleton is assumed to be proportional to the excess

hydraulic gradient.

Although the above works constitute major improvements

on the original Terzaghi theory of one-dimensional consolidation,

in practical problems with deep beds of stiff clays where strains

as well as pressure increment ratios are small, they do not improve

the prediction of the field rate of settlement to any great extent,

(except the non-homogeneous solution of Gibson and Schiffman 1964).

For this we have to turn to the theories of three-dimensional con-

solidation.

The general theory of three-dimensional consolidation of

an isotropic, elastic medium was formulated by Biot (1941))who

showed that the problem of consolidation in the field was :intimately

linked with the problem of stress distribution. In a series of

papers Biot also analysed the particular case of a soil having a

Poisson's ratio = 0 and loaded uniformly over an infinite strip

(Biot 1941a, Biot and Clingan 1941). Later he extended the theory

to cover the more general case of the porous anisotropic medium

34.

(Biot 1955) and gave physical interpretations to the elastic co-

efficients involved (Biot 1957). A number of solutions to pro-

blems of three-dimensional consolidation have been developed in

the context of the radial flow of water to sand drains (Rendulic

1935, Barron 1948, Richart 1957, Hansbo 1960) but the solution of

foundation problems was nor forthcoming until Gibson and Lumb

(1953) obtained numerical solutions to a simplified form of the

consolidation equation (see Chapter 13) for a few specific cases

which, at once, showed the inadequacy of the Terzaghi theory in

predicting the rate of settlement in the field. In a series of

subsequent papers Gibson and McNamee (1957, 1960, 1963) derived

rigorous analytical solutions to the consolidation of various

foundation problems - the case of the circular fdoting was also

considered by Josselyn de Jong (1957) - but numerical evaluations

for a wide range of soil properties are not yet available. Rowe

(1964) developed a theory of consolidation for the flow of water

to a lateral boundary in a stratified medium. Recently Davis

and Poulos (1966) have suggested an approximate method for solving

the three-dimensional consolidation equations and produced a number

of charts which can be used to predict the rate of settlement of

circular and strip footings founded on layers of finite depth.

This will be considered in greater detail in Chapter 13.

35.

CHAPTER 3

A STUDY OF CASE RECORDS OF

SETTUMENT OF STRUCTURES ON CLAY

General definitions

An idealised time-settlement curve for a structure founded

on saturated clay is shown in Fig. 3.1. Prior to the application of

the structural load a depth of soil is excavated to foundation level

during which the soil heaves upwards as a result of the release of over-

burden pressure. During subsequent construction as the pressure on

the foundation increases the soil begins to settle until the net

pressure on the foundation is zero and the settlement is approximately

equal to the heave that occurred during excavation. As construction

proceeds the net pressure on the soil increases and the structure

continues to settle. By the time construction is complete and the

full load is applied the structure has undergone what is commonly

known as the "immediate" settlement (p.). If construction is I

sufficiently rapid this settlement takes place essentially under

condition of no volume change and is primarily due to shear deforma-

tion of the clay.

The application of the structural load, however, causes

excess pressures to develop in the ground water which then begin to

dissipate. The process of consolidation is associated with a change

36.

of volume of the subsoil and the structure undergoes further settle-

ment. This "consolidation" settlement (pc) increases with time

at a rate depending on the coefficient of consolidation of the clay

as well as drainage conditions. The total amount by which the

structure has settled when the excess pore-pressures have fully

dissipated is called the total primary settlement (Pp) and is

comprised of the "immediate" and the "consolidation" settlements.

For some clays, however, the settlement does not cease with

the end of primary consolidation. The long term settlement which

occurs due to creep at esentially constant effective stress and is

known as the "secondary" settlement may continue for many years,

even decades.

In the following sections of this chapter a number of case

records are studied of settlement of structures on both over-

consolidated and normally consolidated clays. It will be noticed

that for most structures measurements of heave that occurs during

excavation are not available mainly because the first observation

of settlement is not made until construction has progressed to a

certain extent. It is, therefore, assumed that the heave and the

subsequent settlement that takes place on restoration of the ex-

cavation load are small compared to the total settlement. This

assumption is justified at least in cases where the net foundation

pressures are relatively large and where the excavations are not

left open for too long a period. Therefore, for the structures

studied, where both the above conditions are satisfied, only the

immediate and consolidation settlements under the net foundation

load are considered.

3.2 Structures on over-consolidated clays

(i) Fire Testing Station, Elstree, London (Skempton, Peck and McDonald 1955)

The Fire Testing Station near Elstree, North London, is a

a reinforced concrete structure 138 ft. x 36 ft. in plan and was

built between April and August 1935. The building is supported on

5 ft. x 10 ft. x 6 ft. deep mass concrete footings situated at a

depth of 7 ft. into the Brown London Clay overlying the stiffer

blue London Clay.

Settlements were observed for 4 years during and after

construction by the Building Research Station. The average time/

settlement curve for seven footings shows that the net settlement

at the end of construction was 0.32" and the structure was still

settling after 4 years.

(ii) Chelsea Bridge, London (Buckton and Fereday 1938) (Skempton, Peck and McDonald 1955)

The new Chelsea Bridge over the River Thames which re-

placed the old bridge in 1937 is a self-anchored suspension type

with two mass concrete supporting piers 28 ft. x 106 ft. in area

founded at a depth of 31 ft. below the river bed in the blue London

37.

Clay.

The settlements of the two piers have been observed for

18 years after construction by which time settlement virtually

stopped (Fig. 3.2).

(iii) Waterloo Bridge, London (Buckton and Currel 1942) (Cooling and Gibson 1955)

The new Waterloo Bridge built between 1938 and 1941 iu a

reinforced concrete structure supported on four river piers 27 ft.

x 117 ft. in area. The time/settlement curves for the four piers,

which are founded at a depth of about 22 ft. below the river bed

into blue London Clay, have been published by Cooling and Gibson

and the one for pier 3 is reproduced in Fig. 3.3.

The average net settlement of the four piers at the end

of construction - taken as the settlement when the full load was

applied together with the net final settlements - is given in

Table 1. The piers were constructed in open cofferdams built with

steel sheet piling driven to 10 ft. below foundation level. The

driving of the piles caused fissures to open up into which water

penetrated allowing considerable swelling to take place. The

foundation did not settle to its original level until after the

piers were constructed and load from the superstructure was trans-

ferred onto the piers by jacking.t. The settlements quoted are net

of this initial heave.

38.

(iv) High Chimney, Bulgaria (Stefanoff et al 1965)

A 600 ft. high chimney was erected in 1962 on a thick bed

of Pliocene lacustrine deposits. The foundation consists of a

100 ft. diameter raft resting on a thin bed of sand cushion over-

lying a bed of highly plastic clay with natural water content near

the plastic limit.

The average time/settlement diagram of four points under

the edge (Fig. 3.4) indicates that the chimney virtually ceased to

settle after only 3 years. Nearly 70% of the total settlement took

place during construction even though the structure took only six

months to erect.

(v) 13 Storey Building, Santos, Brazil (Teixeira 1960)

A 13 storey building in Santos, Brazil, was founded on a

soil, quite untypical of the area, consisting of a thick layer of

highly preeconsolidated silty clay underlying a stratum of dense

fine sand. The structure took just over a year to build during

which 38% of the total settlement occurred.

(vi) Apartment Building, Oslo (Simons 1963)

A nine-storey apartment building was built in 1956 of

reinforced concrete and was founded on the "unusual sequenbe of

16 m of overconsolidated clay overlying normally consolidated clay".

39.

40.

Settlement observations (Fig. 3.5) show that maximum settlement was

reached 33 years after construction and since then no further settle-

ment has occurred.

(vii) High Block, Oslo (Bjerrum 1964)

A 12-storey building 12.5 m x 71.4 m in plan with continuous

strip footings was supported on a layer of very stiff weathered clay

overlying stiff silty clay followed by a soft layered clay resting

on morraine. The thickness of the different layers varied along

the length of the building which has almost ceased to settle after

only two years. Nearly 60% of the total settlement occurred during

6 months of construction.

(viii and Dam D7, and Powerhouse, U.S.S.R. ix) (Nitchiporovich 1957)

Settlement observations on a large number of hydraulic

structures in the U.S.S.R. indicate that most of these structures

have ceased to settle after 10-20 years and that "principal settle-

ment (65 to 85 percent of the total) takes place at the end of con-

struction". All the structures are built of mass ..oz reinforced

concrete with total height ranging from 20-30 m and base width of

14-20 m. The foundation consists of alluvial and glacial clayey

soils and Jurassic, Permotriassic and Devonian clays with moisture

content somewhat lower than plastic limit. Settlements of two

such structures, Dam D7 and Powerhouse PH3, built on 100 m of

Permotriassic clays are given in Table 1.

(x) Talab-Hareb Building, Egypt (Banhet 1953)

This building was built on a site consisting of a layer

of very stiff clay underlain by alternate layers of silty clay and

clayey silt overlying fine sand. The building settled 48 mm during

2i years of construction and the total settlement was no more than

56 mm after 8 years.

(xi) Titanium Pigment Plant, Vareness, P.Q., Canada (Casagrande, L., et al 1965)

A large industrial plant in Vareness, P.Q. was built in

1957 on mat foundation on a site consisting of 90 ft. of stratified

clay underlain by 50 ft. or more of organic clay with no visible

stratification. The clay is lightly overconsolidated but the net

pressure increase due to the structure plus the initial effective

overburden pressure is less than the maximum pre-consolidation

pressure. The structure was built in one year during which more

than 65% of the total settlement occurred (Fig. 3.6).

(xii) San Jacinto Monument, Texas (Dawson 1940) (Dawson and Simpson 1948) (Bjerrum 1964b)

The 570 ft. high Monument on the San Jacinto River in

Texas was erected in 1937 to commemorate the victory of General

Sam Houston in the Mexican War of 1836. The structure is founded

41.

42.

on a monolithic concrete base 124 ft. square resting on 120 ft. of

stiff-fissured Beaumont clay overlying a bed of sand. Settlement

records have been kept since 1936 to the present day (Fig. 5.7).

It can be seen that after about 3 years when the settlement tended

to reach a steady value the building has settled another 5 in. and

is still doing so after 20 years. From a study of the irregularities

of the time/settlement curve Dawson (1948) has correlated them with

occurrence of high wind velocities and concluded that the abnormally

high secondary settlement was a result of this intermittantly applied

wind load. The total primary settlement of the structure has been

estimated as 3.5 in.

(xiii) Quddabi Bridge, Egypt) (Hanna 1950)

This is a 3-span cantelever bridge on two abutments and

two concrete piers which rest on a hard plastic brown clay underlain

successively by silty clay and clayey silt. The structure settled

extremely fast and virtually came to equilibrium after only 2 years.

3.3 Structures on normally-consolidated clays

(i) Skabo Building, Oslo (Simons 1957)

This office building approximately 27 m x 15 m in plan was

constructed of reinforced concrete in 1948. The foundation consists

of strip and rectangular footings founded on a weathered crust over-

43e

lying 20 m of scft blue clay followed by a layer of permeable sandy

clay with gravel.Settlement observations were made at six points of

the building. The final primary settlement was obtained by extend-

ing the straight line portion of the settlement vs. log time plot to

25 years. The construction period was only 9 months during which

the points settled 8-10 cms.

(ii) Gymnasium Hall, Drammen (Simons 1957)

This structure, consisting of a heavier and a lighter section,

approximately 44 m x 20 m, is a reinforced concrete framed building

supported on strip and raft foundations. The soil under the heavier

section which is supported on a raft consists of a thin layer of fine

sand underlain successively by soft silty clay, soft clay and silty

varved clay which overlies fine sand at a depth of 27 m. Settle-

ments were observed for nearly 18 years and an average of points 3

and 4 under the heavier wing are shown in Fig. 3.8. Plotting

observed settlement against log time the settlement at the end of

primary consolidation was found to be 53.5 cm while that at the end

of construction was only 9 cm.

(iii) Monadnock Block, Chicago (Peck and Uyanik 1955) (Skempton, Peck and McDonald 1955)

The Monadnock Block is a 16 storey wall-bearing Masonry

building built in 1891/92. The foundation consists of steel

44.

grillage footings resting at a depth of 12.5 ft. below ground surface

on a crust of stiff clay overlying the soft and medium soft Chicago

clay. Settlement records have been kept regularly for almost 60

years from which it is found that the structure settled 4.5 in.

during 2 years of construction while the total settlement after 6o

years was 22in.

(iv) Masonic Temple, Chicago (Peck and Uyanik 1955) (Skempton, Peck and McDonald 1955)

The Masonic Temple is a 20 storey steel-frame structure

113 ft. x 165 ft. in plan and 302 ft. high and was erected between

November 1890 and November 1891. It is supported on spread footings

founded 14 feet below ground surface on a thin crust of stiff clay

underlain by soft glacial clay. The average time/settlement diagram

for the four corners of the building is shown in Fig. 3.9. The

settlement at the end of construction was 2.2 in. and the structure

virtually ceased to settle after 20 years when the settlement amounted

to 9.8 in.

(v) and Locomotive Shed and Turn Table, Kerava, Finland (vi) (Helenelund 1953)

The Locomotive Shed and the Turn Table at Kerava railway

station 29 km. north of Helsinki were built in the Spring of 1933.

The buildings are supported on reinforced concrete rafts founded on

a layer of gravel fill which rests upon a soft clay deposit underlain

45.

by dense moraine and rock. Considerable settlements have occurred

in 16-20 years (37 cm. for the Locomotive Shed and 16.5 cm. for the

Turn Table) though not more than 17% took place during construction.

(vii) Road Embankment, Leigh (Lewis 1963)

As part of an investigation to assess the value of vertical

sand drains in accelerating the settlement of a foundation, an embank-

ment was constructed on a post glacial clay marsh of the Thames

estuary. Settlement gauges were installed in a 200 ft. length of

the embankment, half of which was built with sand drains. Results

of observations in the area without sand drains are shown in Fig.

3.10. The two-stage loading makes it difficult to determine precisely

the immediate settlement, and this has been estimated as 5.5 in.

The total settlement after 7 years has been 22.5 in. A significant

part of the consolidation settlement must have taken place during

construction because of the small thickness of the clay layer.

(viii) Blast Furnace, Shanghai (Yu, Shu and Tong 1965)

Settlement records of a 30 m tall Blast Furnace, supported

on piles penetrating 20 m into the highly compressible alluvial

deposits of the Lower Yangtze Valley (Fig. 3.11) show that during

6 months of construction the structure settled 5 cm. while the net

settlement at the end of primary consolidation was 25 cm.

46.

(ix) Post Office, Bregenz, Austria (Terzaghi 1933) (Skempton and McDonald 1955)

The Post Office, built in 1893, rests on continuous footings

supported on a layer of sand and gravel overlying a stratum of soft

clay 50 ft. thick. Complete settlement records are available of

eight points of the building covering a period of over 30 years.

The average time/settlement curve of four corners is plotted in Fig.

3.12. The building took 11 years to build during which only 16%

of the total primary settlement, complete in 10 years, occurred.

3.4 Summary and discussion

The observed settlement data of all the structures described

above are summarised in Tables 3.1 and 3.2. The immediate settle-

ment ( f)1), the total primary settlement (1°1)) and the secondary

settlement (fs) have been expressed as percentages of the estimated

total settlement after 50 years (P50) - in columns 8, 9 and 10

respectively. In order to obtain f)50 the settlements were plotted

against log time and the final straight line of each curve was ex-

tended to 50 years as shown in Fig. 3.13 in which the methods of

obtaining the total primary settlement and the time to reach the end

of primary consolidation are also demonstrated.

It can be readily seen from Fig. 3.15 that

F 50= Pp ± (0. or

PP +Ps = 1 (3.40)

50 50

47.

It is possible, from an analysis of the data presented in

Tables 3.1 and 3.2, to make some general observations concerning the

respective behaviours of overconsolidated (0.0.) and . .ormally con-

solidated (N.C.) clays.

1. Structures on O.C. clays, in general, settle much less than

those on N.C. clays. This is only to be expected because one of

the main effects of over-consolidation is to reduce the compressibility

of the clay. One further reason may be that the excess pore-pressure

developed in O.C. clays are usually less than those in N.C. clays

for the same total stress changes, resulting in correspondingly

smaller consolidation settlement.

2. The most striking difference in behaviour of the two types

of clays can be found in the proportions of the total settlement

that takes place by the end of construction. On average, for 0.C.

clays as much as 57.5% of the total settlement occurs during con-

struction compaied to only 15.5% for normally consolidated clays.

A direct comparison of this difference is hard to make because the

immediate settlement (i.e. the settlement at the end of construction)

depends on many factors such as the thickness of the clay layer

relative to the size of foundation, the presence of any sand stratum

and length of the construction period. While individual variations

of the first two factors are difficult to take into account, it can

be seen that the average construction periods in both groups are

similar, being 1.2 years for structures on 0.C. clays and 0.85 year

48.

for those on N.C. clays. Also the settlement at the end of con-

struction must include a part of the consolidation settlement which,

however, is believed to be small for the construction periods con-

sidered. Even taking these uncertainties into consideration,

therefore; the big difference in the proportion of the total settle-

ment that occurs during construction on over-consolidated and

normally consolidated clays may be taken as a characteristic of the

difference in behaviour of the two types of clays.

To find a possible reason for this one has to look once

again into the development of excess pore-pressures in N.C. and 0.C.

clays when they are subjected to increase in total stresses. It

is known that the pore-pressure parameter A for N.Ct clays is, in

general, nearer 1 while for 0.C. clays it is considerably less

(Bishop and Henkel 1962). For identical total stress changes

beneath two foundations, therefore, the excess pore-pressures in

the N.C. clay will be higher than in the 0.C. clay. The effect of

this difference on the settlement of structures is two-folds Not

only will the increase in effective stresses during undrained loading

be greater in 0.C. clays producing proportionately more immediate

settlement, but also the consolidation settlement, which is a

function of the effective stress changes caused by the dissipation

of excess pore-pressures, will be correspondingly smaller.

3. The progress of consolidation settlement is comparatively

faster for 0.C. clays than for N.C. clays. In fact, the secondary

49.

phase in O.C. clays is reached between 1.5 and 7.5 years while the

corresponding period in N.C. clays is between 4 and 25 years.

There is, obviously, considerable variation as one would expect.

Nevertheless, on average, the structures on O.C. clay reach their

full primary settlement only 3 years after construction while for

N.C. clays the figure is 9 years. It is possible that the overall

permeability of overconsolidated clays in-situ may be greater

because of the presence of fissures and joints giving rise to a

higher coefficient of consolidation and hence a faster rate of

dissipation. Further, it would, in general, be expected that

O.C. clays should have higher values of cv than similar N.C. clays,

simply because of the very much smaller compressibility of 0.0.

clays.

Another reason which is considered likely is that the

effective depth of clay beneath a foundation (i.e. the depth within

which the ratio of the increase in effective stresses during con-

solidation to the stresses prior to it is significantly large) may

'only be small compared to the thickness of the clay layer. Terzaghi

(1941) found that at some point beneath the Charity Hospital in

New Orleans the settlement was zero in spite of the fact that there

was some increase in vertical stress at the point. It is possible,

therefore, that there exists a threshold value which must be ex-

ceeded before any significant settlement will occur. If the

effective depth within which this threshold is exceeded and which,

50.

of course, will depend on the dimensions of the loaded area, is

small, the drainage path will also be correspondingly small allowing

dissipation to proceed rapidly.

4. The secondary settlement of structures on O.C. clays is

smaller (average 8.8% of the 50 year settlement) than those on N.C.

clays (21.6%). In fact, the average secondary settlement is as

high as 29.2% of the total primary settlement for N.C. clays compared

to only 9.9% for 0.C. clays. It is well recognised that soft

normally consolidated deposits undergo large secondary deformation

when all the excess pore-pressures have dissipated. Indeed in some

clays like Drammen (Bjerrum 1967) and Mexico City Clay (Zeevaert

1958) this long term settlement may constitute the major part of the

total settlement. Over-consolidated clays, however, have been

subjected to higher pressures in their geological history and any

structural loads that are now applied cause them only to recompress.

In such cases secondary deformation is not likely to be significant.

No.

1

Name of Structure

Fire Testing Station, Elstree

2 Chelsea Bridge, London

3 Waterloo Bridge, London (Av. of 4 piers)

4 High Chinmey, Bulgaria

5 13 storey Building, Santos (Av. of 3 points)

6 Apartment Building, Oslo (Avo of points 2 & 3)

7 High Block, Oslo (Av. of points 3 & 6)

8 Dam D7, U.S.S.Ro

9 Power House, U.S.S.R.

10 Talab Hareb, Egypt

11 Titanium Pigment Plant, Q~bec·.:

SETTLEMENT OF STRUCTURE_S _______ ......

1

Period of Observation (years)

4

18

17

6

6

2

13

10

8

7

2

Construction Period (years)

1

1

0.75

2.0

1.75

1.0

3

Settlement at the end of Construction

( p.) ! ~

0.32 in.

1.20 in.

4 .. 4 em.

0.55 cm.

1.75 cmc

22 cm ..

10 emo

4.6 effio

4.4 in.

Total Observed Settlement

) T

7 Time to reach total Primary Settlement (years)

8 9 10 11

Pi -Pp es Ps F5o ('50 e50 ep % % %

5 6 Estimated

Total Settlement Primary after 50 Settlement years

FP 150

0.58 in. 0.56 in.. 0.59 in.

2.0 in. 1.8 in. 2.05 in.

3.7 in. 3.4 in. 4.0 in.

5.o

3.5

7.5

1.5 3.3 cm. 3.2 cm. 3.5 cm.

10.0 cm. 11.6 cm. 4.0 10.6 cm.

1.75 cm. 1.8 cm. 4.5 1.8 cm.

2.8 cm. 2.7 cm. 3.0 cm. 2.0

ON OVERCONSOLIDATED CLAYS

54.o 95.o 5.0 5.3

58.5 87.8 12.2 13.9

37.5 85.o 15.o 17.6

65.7 91.4 8.6 9.4

38.o 86.2 13.8 16.0

31.5 97.2 2.8 3.0

58.5 90.0 10.0 11.0

64.7 91.2 8.8 9.6

60.6 87.9 12.1 13.8

80.7 94.7 5.3 5.6

66.7 91.0 9.0 9.9

-

(Continued)

51.

33 cm. 31 cm. 34 cm. 5.5

17 cm. 14.5 cm. 16.5 cm. 4.0

5.6 cm. 5.4 cm. 5.7 cm. 4.0

6 in. 6.0 cm. 6.6 cm. 5.0

1 2 3

Settlement No. Name of Period of Construction at the

Structure Observation Period end of (years) (years) Construction

( fl)

12 San Jacinto Monu- 20 1.0 2.0in. ment, Texas

13 Quddabi Bridge, 3 0.33 2.5cm. Egypt (West Pier)

Average 1.2

52.

7 8 9 10 11

Total Observed Settlement

pT

5

Total Primary Settlement

131)

6 Estimated Settlement after 50 years

150

Time to reach total Primary Settlement (years)

li Lip P, 1050 p50 p,o PPP % % %

7.7in- 3-51n.

3.5cm. 3.25cm. 3.4 cm. 2.0 73.5 95.6 4.4 4.6

4.0 57.5 91.2 8.8 9.9

No.

TABLE 3.2. SETTLEMENT OF STRUCTURES

Name of Structure

1

Period of Observation (years)

2

Construction Period (years)

3 Settlement at the end of Construction (

1 Skabo Building, Oslo

10 0.75 8 cm.

(Av. of points 5 & 6)

2 Gymnastic Hall, Drammen

21 0.75 9 cm.

(Av. of points 3 & 4)

3 Monadnock Block, Chicago

55 1.5 4.5 in,

4 Masonic Temple, Chicago

22 1.0 2.2 in.

5 Locomotive Shed, Karava, Finland

20 0.5 7.3 cm.

(Av. of points 3, 4, 5)

6 Turn Table, Kerava 16 0.5 1.5 cm. (Av. of points 2, 3, 6, 7)

7 Road Embankment, Leigh

7 0.5 5.5 in.

8 Blast Firnace, Shanghai

6 0.5 5 cm.

9 Post Office, Bregenz

39 1.5 2.5 in.

Average 0.85

* Observed settlement after 55 years

ON NORMALLY CONSOLIDATED CLAYS

53.

4

5 6

Estimated Total

Total

Settlement Observed

Primary after 50 Settlement Settlement years

PP

P50

37 cm. 49 cm. 57.0 crn.

58 cm. 53.5 cm. 66.0 crn.

22 in. 20 in. 22.0* in.

9.8 in. 9.5 in. 10.5 in.

37 cm. 30.3 cm. 42.0 cm.

16.5 cm. 12.6 cm. 19.5 cm.

22.5 in. 21.5 in. 26.0 in.

27 cm. 25 cm. 37.0 cm.

20 in. 15.5 in. 21.0 in.

7 8 9 10 11 Time to reach total p Primary i Pp Ps Settlement 0 5o P.5o P50 PP (years) % % %

14.o 86.0 14.o 16.3

17.o 13.6 81.0 19.0 23.5

25.0 20.5 9100 9.0 10.0

8.0 21.0 87.2 12.8 14.7

5.5 17.4 72.1 27.9 38.7

5.5 7.7 64.6 35.4 54.8

5.5 21.2 82.7 17.3 21.0

4.o 13.5 67.6 32.4 47.9

11.0 11.9 73.8 26.2 35.5

10.2 15.6 78.4 21.6 29.2

CHAPTER 4

LONDON CLAY - GEOLOGY AND STRESS HISTORY

4.1 Geology of the London Basin

The geology of the area consisting of Greater London and

parts of the neighbouring counties of Essex, Hertfordshire, Bucking-

hamshire, Berkshire, Surrey and Kent has been the subject of ex-

tensive study for more than 100 years. Details of these studies

have been published by various authorities, the more comprehencve

works being Woodward (1922), Buchan (1938), Wooldrich and Lynton

(1955), and Sherlock (1962). The following is a short account of

the various geological materials that occur in the area and are of

interest to engineers.

Figure 4.1 shows the solid geology of the sedimentary

formation in the London basin. The Palaeozoic rocks, the nearest

outcrop of which is 100 miles north and west of London, are the

deepest known formation in the area. Under London itself the rocks

which are struck at about 1,000 ft. below sea-level (Clayton 1964)

form a "Platform" over which the late Mesozoic Strata (Jurassic and

Cretaceous) were deposited. The Jurassic rocks, which come to the

surface in Wiltshire, Oxfordshire and a narrow strip of Berkshire,

are, however, absent all over the London area and the succeeding

formation, the Cretaceous, lies directly over the Palaeozoic rocks.

54-

55.

The freshwater deposits of the Lower Cretaceous - the so -cal)ed

Hastings Beds containing beds of sandstone separated by layers of

clay and the later Weald Clay - can be seen in Kent and Sussex.

The Lower Greensand, which is probably of marine origin, occurs in

a narrow band, and was deposited towards the end of this period.

The Gault Clay was the first of the Upper Cretaceous forma-

tions to be deposited on the Palaeozoic rocks across the London

Platform. This clay and its sandy equivalent, the Upper Greensand,

are exposed in the south and the north-west of the area.

The thick bed of Chalk, most of which is very pure lime-

stone; overlies the Gault and underlies all the Tertiary sediments

occupying the London area. It is possible the Chalk attained a

thickness of up to 800 ft., but subsequent erosion appears to have

removed about 200 ft. in places. Between the outcrops in the

north-west and in the south the Chalk forms a shallow trough

commonly known as the London Basin over which the Eocene Beds were

deposited (Fig. 4.2). The Chalk has, for long, been the chief

source of water supply in the London area having acted as a collector

of the rain falling on the Chiltern Hills and North Downs (see fig.

3.2) (Wilson and Grace 1942).

The early Eocene deposits were the Thanet Sands and the

Woolwich and Reading Beds, collectively known as the Lower London

Tertiaries. The former consist of fine grained sands and silts

and exists in thicknesses of up to 70 ft. thinning down to only

56.

15 ft. in the most westerly exposure. The latter are a variable

group of sands, clays and pebble-beds, occasionally cemented, with

average thickness of about 70 ft.

The most important Eocene formation in the area is the

London Clay which appears today as a very stiff dark-grey or bluish-

grey material. Over much of the area where it is exposed the Clay

has turned brown due to the oxidation of its iron salts by weather-

ing.

The London Clay is considered to have been formed by mud

brought down by a large river and deposited under marine and

estuarine conditions. Further middle and late Tertiary deposits

are thought to have been laid down but these were subsequently re-

moved by erosion. The Clay was later covered by the Quaternary

deposits and in certain areas by the successive glaciations, the

pressures from which consolidated it into a very stiff and compact

clay. The overlying sediments and a considerable thickness of

the clay itself have, in many places, subsequently been eroded

leaving the London Clay today under pressures considerably smaller

than they have been subjected to in the past.

In the vicinity of London, the London Clay is sometimes

covered by sandy Claygate Beds, also Eocene, but in other places,

these are missing and the Clay is covered abruptly by another

Eocene formation, the Bagshot Sands. There are several places

today where the full thickness of the London Clay is present, as

57.

indicated by the overlying Claygate Beds, such as Wimbledon (430 ft.),

Hampstead (400 ft.), Ingatestone (532 ft.) and Sheppey (518 ft.).

Over much of the area, however, erosion has removed a great deal of

the Clay leaving, for example, only between 60 ft. and 130 ft. in

Central London.

The post-Tertiary formations of the London Basin consisting

of the Pleistocene and recent deposits covered much of the solid

formations of the area before they were removed by widespread erosion.

Though at the present time a few high hills carry the early deposits

known as the pebble-gravels, the most significant of the Pleistocene

activities was the advance of the successive ice-sheets into the

North of the area which diverted the Thames to its present course

from a position far to the north where it then flowed. The deposits

of gravel and Boulder Clay associated with the ice sheets can be

found over wide areas in the North-west and Essex. The Quaternary

river deposits on the other hand, which consist mostly of gravel,

sand and silts - for example the Boyn Hill, Taplow or Flood Plain

Terraces - predominate the lower Thames valley.

The Drift Geology of the area has been completed by the

recent deposits of alluvium and peat which form the flat meadow and

marsh lands bordering the river and its tributaries.

4.2 Previous. Work on London Clay

London Clay is the most widespread foundation material in

58.

the London Basin and supports a high density of buildings which have

in recent years been built to considerable heights, e.g. the Shell

Tower, 351 ft. (Measor and Williams, 1962), Portland House, 330 ft.

(Frost and Mason, 1963), the Post Office Tower, 600 ft. (Creasey,

Adams and Laurpitt, 1965) and the Commercial Union Building, 387 ft.

(Williams and Rutter, 1967).

As mentioned in the previous section, uplift and erosion

have removed much of the overlying deposits and, in places, a con-

siderable thickness of the Clay itself, which means that London Clay

as it exists today is heavily over-consolidated. Major structural

features in the Clay are the slight laminations (Cooling and Skempton

1942, Bishop 1947), joints, shear zones and a high degree of

fissuring, the importance of which in determining the undrained

strength of the Clay is now well recognised (Bishop 1966, Bishop and

Little 1967, Hooper and Butler 1967).

The engineering properties of London Clay have been the

subject of many investigations. Cooling and Skempton (1942) con-

ducted one of the earliest studies in connection with the foundations

of the Waterloo Bridge. Skempton and Henkel (1957) published results

of tests made on samples taken from deep borings at three other sites

in Central London - Paddington, Victoria and the South Bank. Soon

afterwards a most extensive study was initiated at the Building Research

Station and at Imperial College primarily on the shear strength

properties of the Clay. Samples were tested from five different

59.

depths of the Ashford Common shaft and from a nearby tunnel. The

results are well documented (Ward, Samuels and Butler 1959, Bishop,

Lewin and Webb 1965, Webb 1966, Ward, Marsland and Samuels 1966).

The influence on strength of anisotropy and sample size have been

investigated by Bishop (1966), and Bishop and Little (1967) while

Agarwal (1967) conducted a thorough investigation on the Clay from

Wraysbury. A comprehensive study of the undrained strength of

London Clay has been made by Hooper and Butler (1967). Residual

strength of London Clay was the subject of Professor Skempton's 4th

Rankine Lecture (Skempton 1964). More recently the author's colleague

Mr. H. T. Lovenbury, has been studying the long-term creep properties

of the Clay from Hendon, (Lovenbury 1968).

4.3 Index Properties

From a study of the published work on London Clay it is

possible to list the range and variation of its index properties.

Table 4.1 summarises the results from 11 different sites spread along

the London Basin. The properties of the Clay from Oxford Circus and

Ongar which were used in the present research programme (for details

see Chapter 8) lie within the usual scatter (Fig. 4.3). There

appears to be a general, though not particularly well-defined, trend

towards lower liquid limit and lower plasticity index in moving from

east to west along the Basin, (Bishop et al 1965).

The mineralogical composition of the Clay as determined by

6o.

X-ray diffraction tests is as follows (Brooker and Ireland 1965):

Quartz 15%

Chlorite and Kaolinite 35%

Illite 35%

Montmorillonite 15%

4.4 States of Stress in-situ

The importance of stress history on the strength and deforma-

tion of soils has long been recognised. Very little is known, however,

of the complete states of stress in the natural London Clay in-situ.

While it is often possible to determine the vertical effective

stresses at a site from a knowledge of the soil profile and ground

wat6r conditions there is no direct way of determining the horizontal

effective stress. It is only in recent years that attempts have

been made to estimate the in-situ horizontal stresses in London Clay

(Skempton 1961, Bishop, Webb and Lewin 1965).

Like many other heavily over-consolidated clay and clay-

shales, e.g. Little Belt Clay (Hvorslev 1960, Brinch Hansen 1961),

BearpaeClay shale (Peterson 1954, 1958, Terzaghi 1962) and Fort Union

Clay shale (Smith and Redlinger 1953), London Clay has undergone a

number of distinct phases of deformation and stress changes during

its geological history. An element of the Clay as deposited by a

large river in marine conditions is represented by point A in Fig.

4.4a. As more deposition followed the vertical pressure on the

61.

element gradually increased and the Clay consolidated to point B

-4en the maximum over-burden pressure was reached. At the same time

horizontal pressures were also developing in the Clay, the magnitude

of which depended on its shear and deformation properties, as shown

by the curve ab in Fig. 4.4b.

After consolidation the Clay was left for a considerable

period of time with little or no change of load. During this period

of sustained loading many physical and chemical changes are thought

to have taken place in the Clay. The process described by Bjerrum

(1965) as diagenesis, enabled the Clay particles'physically to con-

form' to each other and develop a certain amount of adhesion and

cementation, called the 'diagenetic bond'.

In a later geological period the pressure on the Clay was

reduced as a consequence of the removal of over-burden by erosion,

and the Clay tended to swell. The amount by which it was able to

expand during the process depended on the recoverable part of the

strain energy that was absorbed during consolidation (Bishop et al,

1965). According to the hypothesis of Bjerrum (op. cit.) this re-

coverable strain energy depended on,

(i) the elastic deformation of the flexible flake-shaped clay

particles. (Provided the particles were not strained beyond their

elastic limit the stored energy would be released on unloading).

(ii) the extent of the diagenetic bond. (Bjerrum describes

three types of bond - weak, medium and strong - in descending order

62.

of recoverability).

Following the above hypothesis, it is believed that London

Clay, after deposition had developed sufficient digenetic bonds so

that on unloading it rebounded along the curve BD and not along BC

which would have been the case had only a weak bond developed. The

Clay near the surface, however, may subsequently have expanded more

owing to the release of the locked-in energy caused by breakdown of

the bond due to weathering.

Now, during rebound the clay had freedom to expand in the

vertical direction, but was restrained in the horizontal direction.

As a result the effective vertical stress decreased relatively more

than the effective horizontal stress (curve bd) . The latter was

always greater than that corresponding to the same vertical stress

during deposition. With increasing over-consolidation the ratio

Ko of the effective horizontal stress to the effective vertical

stress gradually increased and beyond some point (X) the horizontal

stress exceeded the vertical (K0 ) 1). Had only a weak bond developed

the horizontal stress would be even higher because the Clay would

then try to expand more, yet being restrained in the horizontal

direction.

At the present time (1968) no method is available of

measuring the in-situ stresses in London Clay. It has already been

mentioned that where there are no artificial complications due to

tunnelling etc., the effective vertical stresses can be calculated

63.

reasonably accurately from a knowledge of the soil profile, ground

water conditions and unit weight of the materials. Skempton (1961)

and Bishop et al (1965) have estimated the in-situ horizontal stresses

of London Clay and Bradwell and Ashford Common from measurements of

initial suction in undisturbed samples (see Chapter 8). These two

sites, situated, as they are, at the two ends of the London Basin,

have been subjected to estimated vertical pre-consolidation pressures

of 220 and 600 lbs./in2 respectively.

The stresses that may have developed during deposition of

London Clay can be determined from laboratory Ko tests on remoulded

samples. Jaky's expression Ko = 1 Sin 0', though originally

obtained from tests on granular material is found to give satisfactory

results on Clay (Bishop 1958, Simons 1958), though the expression

Ko = 0.95 - Sin 0' has been found to be closer to cohesive soils

(Brooker and Ireland 1965).

The only Ko tests on remoulded London Clay loaded to as

high a pressure as 2,000 lbs./in2 (well above the maximum pre-

consolidation pressure) have been reported by Brooker and Ireland

(1965) who obtained a constant value of Ko = 0.64 for the entire

stress range (the expression Ko = 0.95 - Sin 0' (0' = 17.50) gives

the value of 0.65). Their results are replotted in Fig. 4.5 and

assuming that the same stress changes occurred when London Clay was

deposited, the complete stress histories of the Clay at Bradwell and

Ashford Common have been reconstructed. (Fig. 4.5).

64.

During deposition the stresses increased along the curves

oa and od, a and d being the maximum pressures to which the

Clay had been subjected, at the two sites. On subsequent removal

of the over-burden pressures by erosion the stresses followed the

curves abc and def, the sections be and of indicating the

estimated in-situ stresses today at various depths of the two sites.

The stresses c and f represent points nearer the surface while

b and e are for the deepest points for which data are available.

The decrease of Ko with depth reflecting the influence of over-

consolidation ratio can be easily established from the curves cb

and fe . (A detailed discussion of this will be presented in

Chapter 10).

It is appreciated that the general picture presented herein

is only an approximation and even somewhat idealised because it is

probable that some minor cycles of loading and unloading have not

been taken into account. Yet it is believed that until a more

accurate method of measuring the insitu stresses evolve, the above

gives a reasonably correct assessment of the stress history of

London Clay, (Skempton 1961).

Site Liao

TABLE 4.1

Activity Sp. Ir. Gs. Reference WPI%

Clay Fraction

< 21116

Bradwell 95 30 65 52 1.25 2.75 Skempton (1961)

Malden 85.1 31.4 53.7 56.5 0.95 2.70 Bishop and Little (1967)

Paddington 79 24 55 Skempton

Victoria 73 29 44 ) ac Hankel (1957)

St. Paul's 73 24 49

South Bank 78 28 50

Waterloo 77 27 50 53 0.95 Cooling & Bridge 3kompton

(1942)

Ashford Common

67 27 40 54 0.76 2.74 Bisliop et al (1965)

Wraysbury 74.0 28.5 45.5 57 0.80 2.68 Agarwal (1967)

Ongar 67.5 26.5 41 48 0.85 2.71

Oxford 62.5 26.7 36 48 0.75 2.63 Circus

66.

CHAPTER 5

STRESS-PATH

The Concept of Stress-Path

A basic proposition examined in this thesis is that the

deformation of an element of soil is a function not only of magnitudes

of the applied stresses but also of the manner of their application.

In other words, a knowledge of the magnitudes of stress increase is

not, in itself, sufficient to indicate precisely how a soil element

is going to deform. To obtain a more complete picture we have to

consider also how the applied stresses change, at what rate and in

what relation to one another.

A stress-path is essentially a line drawn through points on

a plot of stress changes. It shows the relationship between the

components of stresses at various stages in moving from one stress

point to another. In the present thesis, however, consideration is

given only to cases where by virtue of symmetry the intermediate and

minor principal stresses are equal and where the vertical and hori-

zontal stresses are the principal stresses, e.g. in the laboratory

triaxial test or along the centre line beneath a loaded circular area.

There are, of course, many ways of plotting stress-paths, the one most

widely used in the study of shear strength of soils being a plot of

shear stress vs. normal stress. But in studying the deformation of

67.

soils, a simple plot of vertical stress versus horizontal stress has

been found to be more convenient and that is the system used through-

out in this thesis.

Definitions

The three types of stress-path relevant to a study of the

effect of stress-path on the deformation of Clays are:

(i) the effective stress-path (ESP)*

0-1 vs. (7 I or v/ vs (T

1 3 (ii) the total stress-path (SSP)*

or1 vs 0

3 or 0- vs h

(iii) path of total stress minus static pore pressure

[(T - us)SP] *

( 0- - us) vs (Cr3 - us) or (Cr us) vs (6'h - us) 1 v

5.3 Stress-path in Laboratory Tests

To illustrate the concept of stress-path the following

standard laboratory tests on saturated clay will be considered:

(i) Isotropic (all-round) consolidation test

(ii) Consolidated drained test

(iii) Consolidated undrained test with pore pressure measure-

ment

These notations were used by Lambe (1967).

68.

(iv) Anisotropic consolidation and Ko tests.

In Fig. 5.1 is plotted the stress-path for simple isotropic consolida-

tion in the triaxial apparatus. At the start of a test the effective

and total stresses are represented by A and A1, ,1`o being the

back pressure. With the drainage valves closed the total stresses

are increased to B1 and the sample is then consolidated against

the original back pressure uo whereby the effective stress changes

from A to B. The same stress-path is obtained if the cell

pressure is increased slowly at a constant rate without allowing any

excess pore-pressure to develop. The sample will compress in all

directions as shown.

In a drained test (Fig. 5.2) a sample is first consolidated

to effective stresses represented by the point A. The axial stress

is then increased slowly holding both the cell pressure cr and 3 back pressure uo constant. cr-

3' thus remains unchanged while the

0- 1 increases until the sample reaches failure at B. The corres-1

ponding total stress-path is A1B1. The paths for a test with

cri constant, cr3 decreasing are also shown.

In a consolidated undrained test (Fig. 5.3) a sample after

being consolidated isotropically to point A is loaded axially under

undrained conditions while (53 is kept constant. Excess pore-

pressure is set up and if the pore-pressure parameter A is positive,

but less than 1, ' will increase and cr3' decrease until Cr

1

approaching failure when due to a sharp drop in the value of A (as

69.

with heavily over-consolidated clays) the latter may begin to rise.

The total stresses during the process move from Al to B1. In

this type of test there is no volume change during shear and the

sample will deform as shown in Fig. 5.3.

In most practical problems, however, consolidation does not

occur under isotropic stress conditions. For example, the con-

solidation of a natural soil during deposition is under conditions

of no lateral yield for which the stresses increase in the ratio

C71/(7 = Ko. The same applies to the triaxial Ko test where 3

compression is only one-dimensional. Effective stresses in such

tests will increase along the Ko

line (Fig. 5.4). Tests may also

be performed to reproduce conditions where an element of soil having

been deposited under Ko

conditions is stressed along a path AC

or AD. If the stress-path lies above the Ko line the sample will

compress vertically and expand laterally while a path lying below

the Ko line will produce both vertical and lateral compression.

5.4 Stress-path in the Field due to Foundation Loading

The present section will be devoted to the consideration

of how a soil in the field is stressed when it is subjected to a

foundation loading.

In its natural condition, before any load is applied, an

element of soil is in a Ko state of stresses. Depending on

whether the soil is normally consolidated or over-consolidated the

70.

horizontal stress in-situ may be smaller or greater than the vertical

stress (see Chapter 4). Because in-situ London Clay is heavily

over-condolidated Ko is usually greater than 1.

Let us now consider an element of London Clay beneath the

centre of a uniformly loaded circle. The in-situ effective stresses

(p and Kop) are represented by the point A and the corresponding

total stresses (effective stresses plus the piezometric pressure) by

Al

in Fig. 5.5. Due to the foundation pressure q the stresses

on the element increase by Aar and A.6h1

. If the pressure is

applied sufficiently fast so that no drainage occurs during the load

application, the element will deform without any volume change and

any vertical compression will be associated with a lateral ex-

pansion.

Now, the increase of stresses Ckciv and 4011 - which

are increments in the principal stress directions in this instance

will set up an excess pore-water pressure in the element according

to the equation (Skempton 1954)

Li u = B Q.c5- + A( &cry ) hi hi (5.4.1)

If the clay is saturated, as all clays below the water table are,

B = 1. Then,

,Lu = LK)"hi + A(O.crv 6o- hl) (5.4.2)

Therefore, immediately after the load application the effective

stresses are:

(v' )o = p + tsv - u

h1)o = K

op CS.(7-

111 - 4Nu

Since for most clays, and certainly for London Clay, the value of

A is positive in the range of stresses normally encountered in

practice, the excess pore-pressure Qu is greater than cr h1

(see eqn. 2). So while the effective vertical stress increases

on load application the effective horizontal stress decreases and

the stress point moves from A to B. The vertical strain during

loading is, therefore, a function of the stress-path AB.

The element now begins to consolidate. At the early

stages the increase in effective horizontal stress is only a re-

compression until the original value Kop is restored. Beyond

this point any further increase of horizontal stress is net while

the element is subjected to a net increase in vertical stress during

the entire process of consolidation.

Now during load application, in undrained conditions, a

saturated clay behaves as an incompressible medium with Poisson's

ratio 1 = 0.5. As the excess pore-pressures dissipate, however,

Poisson's ratio decreases and finally drops to its fully drained

71.

72.

value at the end of consolidation. This change in Poisson's ratio,

however, is not likely to have much effect an the vertical stress

(for an elastic, isotropic, homogeneous medium, the vertical stresses

are independent of the material parameters) but the horizontal stress

will decrease by an amount to its new value Kop + 6crh2

where

46-12 6.611 (5.4.4)

So during consolidation the element will follow the

effective stress path BD while it would have moved from B to C

had the total stresses remained unchanged. After full consolidation,

therefore, the stresses are:

and

( cg"Of = p + A crv

(6'h') f ') = K o h1 p + (dtgr - ) = K p A

h2

and the change of stresses during consolidation:

A-crv c 1 = au ) ) )(5.4.6) )

crh'c = (L1u - ) )

73.

5-5 Stages of an "ideal" Experimental Programme for Settle-

ment Analysis

An ideal settlement analysis should take into account the

complete pattern of stresses an element of soil will be subjected to

in the field. In order to determine the relevant vertical strains

laboratory tests should be performed under identical stress con-

ditions and an integration of all such vertical strains beneath a

loaded area would give the settlement of the structure.

Fig. 5.4 shows the stages of experimental programme that

should ideally be followed in the laboratory. When a sample is

removed from the ground without mechanical disturbance or change in

water content the total stresses are reduced to zero and a negative

pore-pressure is set up (Skempton and Sowa 1963). (A detailed study

of the state of stress after sampling is presented in Chapter 8.)

In the first stage, therefore, the in-situ stresses will be restored

so as to get back to the condition before sampling. Next a set of

stresses, identical to those the sample will be subjected to due to

the foundation load, will be applied under undrained conditions.

Both the vertical strain and the pore-water pressure will be measured.

The sample will then be consolidated against a back pressure equal to

the original pore-water pressure uo while at the same time the

horizontal stress will be decreased to its final value. The vertical

strain in this last stage together with that during undrained loading

will give the total strain of the element in the field. The

74.

settlement will then be obtained by integrating the vertical strains

thus obtained of all the elements beneath the loaded area.

5.6 Methods of Settlement Analysis

The ideal method of settlement analysis postulated in the

preceding section, although making use of the true pattern of stresses

and strains, has never been used in practice. First, the ex-

perimental programme is rather complicated and secondly, a number of

samples from various depths should be tested in order to obtain the

actual variation of strain with depth before any meaningful in-

tegration can be made. Lambe (1964, 1967) proposed the selection

of an "average" element at some depth below the loaded area on

which a limited number of triaxial stress-path tests could be per-

formed and the strain thus obtained assumed constant throughout the

depth of the clay stratum.

The methods most widely used in practice, however, are the

ones based on the oedometer test. Table 5.1 summarises all these

methods and in Fig. 5.5 are shown the effective stress-paths

associated with each of them. In plotting the stress-path AG in

Fig. 5.5 it has been assumed that an undisturbed sample when sub-

jected to the in-situ vertical stress also restores the in-situ

horizontal stress so that subsequent stress-path for loading starts

from the point A. How far this assumption is valid will be dis-

cussed in Chapter 10.

75.

Method 1, described by Skempton and McDonald (1955) as the

conventional method, was first proposed by Terzaghi (194) and later

used by Taylor (1942, 1948). It assumes that all settlement occurs

from one-dimensional compression and that the excess pore-pressure

is equal to the increase in vertical stress. The corresponding

stress-path is, therefore, AF.

Method 2 recognises that the soil undergoes shear deforma-

tion during undrained loading (path AB), and this causes the

immediate settlement, but still assumes that the excess pore-pressure,

which should be a function of the induced shear stress, is equal to

the vertical stress increment. Consolidation settlement, therefore,

occurs along the stress-path AF.

This inconsistency in Method 2 has been overcome by

Skempton and Bjerrum (1957) in a very important contribution des-

cribed in Method 3. According to this the immediate settlement is

a function of the stress-path AB while the consolidation settlement

is based on the increase of stresses along the path EF. The latter

will therefore be only a part of the total strain along the path AF

depending on the magnitude of the excess pore-pressure set up

during loading.

While the method of Skempton and Bjerrum introduces for the

first time the concept of stress-path in settlement analysis the

assumption is still implicit that during consolidation all strain is

one-dimensional which requires the horizontal stresses to adjust

accordingly. There is, therefore, a discrepancy between the field

stress-path BD and the path EF used in the analysis. The con-

dition of no lateral strain may be approximately true in cases like

that of a loaded area which is very large compared to the thickness

of the clay layer. But in the majority of field problems the above

condition may be far from true. Moreover, while computing the

immediately settlement method 3 obviously accepts that the clay

undergoes lateral deformation during loading (with constant volume,

there would be no settlement, otherwise) but this is neglected in

estimating the consolidation settlement. Consequently, of course,

the effect of the horizontal stress is completely ignored except in

determining the excess pore-water pressure.

The experimental work reported in this thesis has been

directed towards investigating the effect of different stress-paths

on the axial and volumetric strains of London Clay and the results

so obtained have been compared with those from a parallel series of

oedometer tests.

AF Pc

Cc =roed Terzaghi (194) Taylor (1948)

..„1 mvAcr2.dz

= roed From Skempton and

= 0-2.dz

Cc =14- Foed Skempton and Bjerrum (1957)

f----SmvAu.dz

f(Bu)

Bjerrum (1957)

TABLE 5.1. METHODS OF SETTLEMENT ANALYSIS

Immediate Settlement

Consolidation Settlement

Method Stress-Path (See Fig.5.7)

Amount

Excess Pore

'Stress-

Pressure c Path (See Fig.5.5)

Amount

Reference

2

3

4 (Field stress path)

AB

AB

AB

•••

P• = c1B°-ID

_ B 1*

f(AB)

Au = Arrir

6u .74 cs- v

A u is + A(Acr -AO")

v h

Au =

ga Cr cs

v h

AF

EF

BD

78.

CHAPTER 6

STRESS-DISTRIBUTION IN SOIL MEDIA

6.1 Introduction

An essential step in settlement analysis is the determina-

tion of the magnitude and distribution of stresses that are

developed in the soil due to the application of the structural load.

It is these stresses which cause not only the initial, elastic

settlement but also the consolidation settlement, which is a con-

sequence of the dissipation of the excess pore-water pressures.

The stresses and strains in a mass of soil depend on the

stress-deformation characteristics, anisotropy and non-homogeneity

of the medium, and also on the boundary conditions. But the task

of analysing stresses taking all these factors into consideration is

extremely complex and, therefore, the attempts that have been made

to date are based on certain simplifying assumptions. The most

widely used of these is the case of the homogeneous, isotropic

elastic medium,

It is well understood that the assumption of linearity of

the stress-strain relationship is a questionable simplification

because soils in their behaviour are essentially non-linear. But

no other theories have yet been developed to describe the response

of soils to stress changes, and within the comparatively small range

79.

of stresses that are normally imposed by structural loads, the

assumption of linearity, for most soils, may be considered to be

reasonably valid. Also, limited field evidence reported by

Plantema (1953) and Turnbull et al (1961) show that measured stresses

correspond fairly well with those predicted on an elastic basis.

Therefore, refinement of the methods of stress analysis based on

the theory of elasticity - still assuming the validity of the

linearity of stress-strain relationship, but taking into considera-

tion the variations of properties within the soil mass - seem to be

justified. In this chapter a detailed study is reported of the

stresses and displacements in non-homogeneous elastic soil media.

A short section on the effect of non-linearity of stress-strain

relationship on the distribution of stresses is also included.

6.2 Soil as a homogeneous, isotropic, elastic medium

The assumption that is most widely made of soils in

determining the stresses beneath a foundation is that of an elastic

medium (i.e. linear stress-strain relationship) with uniform pro-

i perties at all points and in all directions.athouGh/pr

n actice a real

soil can hardly be even approximated to such an ideal medium,

the mathematical solution to this problem was the only one available

to engineers for a very long time. Boussinesq (1885) (see Terzaghi

19'+3) was the first to obtain expressions for the components of

stresses and strains within a semi-infinite homogeneous elastic

80.

medium due to a point load acting on the surface and perpendicular

to it. Since the principle of superposition holds for such a

medium it has been possible to use these expressions to determine

the stresses and deflections caused by loads applied over finite

areas on the surface. Love (1923) gave equations for stresses and

deflections caused by a loaded circular rigid plate and Newmark

(1942), by integrating Boussinesq's equation for vertical stresses,

derived the expression for the stresses under the corner of a

uniformly loaded rectangular area. The tables and charts prepared

by Newmark and later by Fadum 948) are now almost universally used

to calculate the vertical stresses beneath a foundation. The case

of a uniformly loaded strip was solved by Carothers (1920) and

JArgenson (1934) and Bishop (1952) used stress functions and re-

laxation technique to calculate the stresses in and underneath a

triangular dam. The most complete pattern of stresses, strains and

deflections beneath a uniform circular load on a homogeneous half-

space can be obtained from tables prepared by Ah73an and Ulery (1962).

From all these results it can be seen that the vertical stresses in

a homogeneous, isotropic elastic body is a function only of the

dimensions of the loaded area and independent of the elastic pro-

perties of the soil. This is not true, however, of the lateral

stresses and displacements.

81.

6.3 Non-homogeneity in soils

It has been mentioned above that engineering properties of

a soil are not normally uniform throughout its mass. This non-

uniformity may manifest itself in both spatial (non-homogeneous)

and directional (anisotropic) variations of the modulus of deforma-

tion. Consideration will be restricted in this chapter to isotropic

non-homogeneity i.e. to cases where elastic parameters of the soil

are not uniform with depth.

The variation of soil properties with depth may be due to

many factors. Often the subsoil consists of different geological

formations with very different characteristics e.g. a clay deposit

underlain by sand or rock. If the underlying stratum is well below

the surface of the clay relative to the size of the loaded area, its

influence may only be marginal. On the other hand, even in a deep

layer of apparently homogeneous material, the rigidity of the soil

generally increases with depth due to the increase in effective over-

burden pressure.

In dealing with the first type of non-homogeneity mentioned

above, a subsoil is often considered as a layered system. Much work

has been done on this subject in recent years, particularly in con-

nection with the design of pavements and runways and in the follow-

ing section of this chapter a review is presented of the available

results.

In the case of continuous variation of elastic parameters

82.

with depth Gibson's analysis (Gibson 1967, 1968) has been used to

compute stresses and deflections beneath circular and strip footings.

6.4 Two-layer systems

A simple two-layer system (Fig. 6.1) may consist of either

1) two elastic layers with different engineering properties

or 2) a single elastic layer on a rigid base.

6.4.1 Two layers with different elastic parameters

This situation is often encountered in the case of pavements

where a stiffer surfacing is placed on a less stiff subgrade. In the

case of foundations, however, the situation is often reversed and one

may encounter a layer of soft material overlying a stronger deposit.

Biot (1935) and Picketts (1938) were among the first to

attempt to solve the problem of stress distribution in the two-

layer rigid base system, fig. 6.1b. Their results could only be

used, however, to determine the stresses at the surface of the base

layer. In a series of papers in 1943 and 1945 Burmister (Burmister

1943, 1945a, b, c) presented the general theory of stresses and

displacements in layered soils from which exact solutions could be

obtained for axi-symmetric loading. Using Burmister's analysis

Fox (1948) published tabulated values of stresses due to a uniform

circular loading with or without friction at the interface for the

case of Poisson's ratio "Z) = -. The case of the line or strip

83.

loading was analysed by Lemcoe (1961) who developed equations of

stresses for a general two-layer system and tabulated numerical

values for the particular case of E1/E2 = 50 and "")1 = v = 4.

In the general two-layer system for a circular load the

stresses depend on the values of11 1 and on the two parameters

(see Fig. 6.1a).

a = ri and K = E1

2

where b = radius of the loaded area

h = thickness of the top layer and

E E3

are the elastic modulii of respectively the top and

bottom layers.

In Fig. 6.2a are plotted the distribution of vertical

stresses beneath the centre of a circle for the special case of

a = 1 and where the upper layer is stiffer than the lower. It can

be seen that the presence of the stiff upper layer has a considerable

influence on the stresses, particularly in the vicinity of the inter-

face. For example, a rigid upper layer which is five times stiffer

than the subgrade (i.e. E1/E2 = 5) reduces the stress at the inter-

face to 60% of the Boussinesq value. This load spreading capacity

of the stiff upper layer has been successfully employed in the

design of pavements on soft subgrades.

The effect of relative size of loaded areas, and thickness

84.

of the upper layer on the vertical stresses at the interface is

shown in Fig. 6.2b. The upper layer is most effective is spread-

ing the load when its thickness lies between b and 3b while for

very thin and very thick layers the stresses approach the Boussinesq

values.

The case of a foundation where a Ooft layer is underlain

by a stiffer deposit (E1/E2 s 1) has not been evaluated but from

extrapolation it can be concluded that the stresses in the upper

layer will, if anything, be greater than those for a homogeneous

medium.

6.4.2 Single elastic layer on a rigid base

This is a special case of the above problem with the elastic

modulus of the bottom layer E2

=c).9. The problem was first solved

by Burmister (1956) who extended his earlier work to analyse the

stresses and strains in the upper layer of a two-layer rigid base

system. From his influence charts it is possible to obtain the

complete pattern of stresses and displacements under the corner of

a uniformly loaded rectangle for Poisson's ratio s"4- = 0.2 and 0.4.

The same problem was elaborated by Poulos (1967) who used

Burmister's theory to compute a set of influence factors for stresses

and surface displacements due to a point load, for values of Poisson's

ratio = 0, 0.2, 0.4 and 0.5. By integration of these point

load factors he then calculated the corresponding influence factors

85.

for (a) line loading

(b) strip loading and

(c) sectirr. loading.

Using the values for the sector loading and applying the principle

of superposition it is possible to determine the complete pattern

of stresses and displacements for any shape of the loaded area. With

these results the writer has calculated the vertical and radial

stresses beneath the centre of a loaded circle for values of

= 1, 2, 4 and 8 and for = -2- (see Fig. 6.3). The stresses for

the homogeneous half-space (Boussinesq) are also plotted for com-

parison. It can be seen that the presence of a rigid layer at a

shallow depth relative to the size of the loaded area drastically

alters the stress pattern. For small values of Ili just underneath

the load vertical stresses may even be greater than the applied

pressure. With increasing depth, however, the effect of the rigid

base gradually diminishes and for To- > 8 the stresses are almost

indistinguishable from the Boussinesq values.

An approximate method that is widely used to calculate the

surface displacement of a two-layer rigid base system was suggested

by Steinbrenner (1934) (see Terzaghi 1943). In Fig. 6.4 is shown

a comparison between the theoretical settlement of the centre of a

uniformly loaded circle as calculated by Poulos and the approximate

settlement based on the Steinbrenner method. For 11 = 0, 0.2

and 0.4 the approximate method underestimates the settlement by up

86.

/ to about 15% for shallow layers (—

h \ 0.5) though with increasing b

thickness the error decreases. For an incompressible medium of

shallow depth ( = 0.5), however, the approximate method is

grossly in error overestimating the settlement by as much as 100%

or more for h —< 0.5, but for layers with —,>1 the error is not

more than 10%.

6.5 Three-layer systems

The analyses of three-layer soil systems (Fig. 6.5) are

much more complex than for two-layers and solutions have only been

obtained for stresses and deflections beneath a uniform circular

load. Burmister (1945) was the first to develop the general theory

for such a system with both rough and smooth interfaces. Since then

the problem has aroused great interest amongst highway engineers in

their attempts to determine the stresses under a road section con-

sisting of a surfacing and a base layer overlying the subgrade.

Burmister's equations were used by Acum and Fox (1951) to calculate

the stresses at the interfaces (for "., 1 = = 0.5, and for full

continuity between the layers). Schiffman (1957) presented methods

for numerical solutions of influence values and tabulated results for

a particular case. It is not until recent years, however, that

Burmister's work has been extended to compute the stresses and

deflections for any combination of thicknesses of the individual

layers and size of the loaded area. Jones (1962), Peattie (1962)

87.

and Kirk (1966) have published charts and tables giving the stress

factors for any combination of three-layer systems while De Barros

(1966) and Uneshita and Meyerhof (1967) have published those for

deflection factors.

The stresses and strains in a three-layer medium (Fig. 6.5)

are governed by the following non-dimensional parameters:

h 2

H = k1 =

E

E b a = and k = — 2 h2

h2 E2

E3

where b is radius of the loaded circle

h1 and h

2 are thicknesses respectively of the first and

second layers. (The bottom layer is semi-infinite).

E1, E2, E3

are the Elastic modulii of the 1st, 2nd and

the 3rd layers respectively.

Fig. 6.6 shows the effect of the relative thicknesses of the two

stiffer upper layers on the stresses and deflections beneath the

centre of a loaded circular area, for the particular case of

h1 h2 = 2b. For the purpose of comparison the Boussinesq stresses

and the deflections calculated on the basis of Boussinesq stress

distribution are also plotted. It can be seen that the actual,

values are lower than those given by Boussinesq although the maximum

discrepancy in deflection is no more than 25%.

For the situation where the layers become successively

stiffer with depth (k1 = .2, k2 = .2) (Fig. 6.7), the assumption

88.

of homogeneity will underestimate the stresses by up to 30% for h1

h1

h2

while for smaller relative thicknesses of the top

layer the error is considerably less.

6.6 Multi-layer systems

The problems of multi-layer systems involve immense

complexity and to date no analytical solution is available for any-

thing consisting of more than three layers. Vesic (1963) has

suggested an approximate method of calculating the surface settle-

ment of a foundation on a multi-layered medium assuming Boussinesq

stress distribution but using the proper elastic modulus for the

respective layers. His charts and method of calculation are shown

in Fig. 6.8. Vesic observed that in three-layered systems the

shape of the deflected surface computed by this approximate technique

agrees better with measured deflections of pavements than the more

rigorous analyses.

De Barros (1966) proposed an approximate method of

reducing a multi-layer system to a three-layer one, keeping the

subgrade unaltered, by successively attributing to the two adjacent

layers an "equivalent modulus" according to the equation

h 2)35 3 E2 3 - ,

1 + h2- 2 El' 2

_ h1 + h2

He found that using this technique and reducing a three-layer system

89.

to an equivalent two-layer one the approximate method is correct to

within 10% for 21>-1 and 15% for -1110, 2.

An analogous expression was first proposed by Palmer and

Barber (1940) to reduce a two-layer system to an equivalent homo-

geneous medium which yielded deflections very close to Durmister's

two-layer analysis.

6.7 Non-homogeneous medium

The problem of the non-homogeneous soil medium whose

modulus of elasticity varies as a continuous function of depth has

received only limited attention so far. Korenev (1957), Sherman

(1959), Golecki (1959), Cuban (1959), Lekhnitskii (1962) have

studied particular problems of non-homogeneity, but no comprehensive

theory had been presented until Gibson developed the theory (Gibson

1967, 1968) of stresses and displacements in a non-homogeneous,

isotropic elastic half-space subjected to strip or axially symmetric

loading normal to its plane boundary.

6.7.1 Formulation of theory

Let us consider the problem of plane strain as shown in

Fig. 6.9. The modulus of elasticity (E) of the medium varies as

some function of depth, E = E(z), but the assumption is made that

the other elastic parameter, the Poisson's ratio remains

constant.

Cr- - xx YY zz )

cr ( cr fix) zz

) (6.2a) ) ) ) ) (6.2b) ) ) ) ) (6.2c) )

= xx E(z)

( = = YY E(z)

zz =

E(z) zz - ( + ) xx yy

( 1 - V ) crxx - z z

1 t )zz= zz xx 2G(z)

xx 2G(z)

The equilibrium equations at any point, in the absence

of any change in body forces, are given by

90.

#2) 6—xx -b cr. xz

x -se z

b(7- zz .c-xz z x

-- 0

_ o

The relationships between stresses and strains are:

From (6.2c):

6- = ( 4- ) YY xx zz (6.2d)

Substituting (6.2d) in (6.2a) and (6.2b),

where G -

93..

2(1 + )

Now, the dilation

e = = Z) U " U XX 22 x z (6.3a)

In terms of stresses

e - 1 - 2"1 (Qr C"J"'C"J"') xx zz 2G(z)

(6.3b)

Substituting (6.3b) in (6.2e)

Crxx = 2G(z)( xx 1 e) )

- ) ) (6.2f)

cr zz = 2G( z) ) (6" zz 1 -

)

)

Putting 04 _ 1- 2

Cr xx = 2G(z) ((+ coLe) ) xx ) ) (6.2g) )

0-z = 2G(z) (6 zz +04,e) ) z )

F coc = 2G(Z)L

[ crz = 2G(z) z

and cr z = 2G(z) x

au

w

w z a x

, e . V w + 200 2h(z) w (—

z and ÷ (4\ e = 0

I

V2u (1 + 20) h(z)(ax z

+ aw = o x

e

In terms of displacements u and v

92.

Taking derivatives of CY;x, cr-z and Y in equation (6.4) zz

and substituting in (6.1a) and (6.1b), we get the equilibrium

equations in terms of displacements

where ,c72 62- b2 = and .2 e) z2

h(z) = d [log G(z)-.1 dz

Let us now take the case of an incompressible medium where

e = o, 14 = 0.5 and 04. = 00

2 u+ -f +h ?)w) ax 7Cz ax

=0

+ bf + h(232.11 + f 's) = 0 az Oz

2w

v 2u - h(:11-1+ aw bz

from (6.7b) fh + f w = 2w 2h.

z z

from (6.7a) b f bx

(6.8a)

(6.8b)

The product (0(e) becomes indeterminate, so introducing the

function

f (x,z) = (1 + 20)e (6.6a)

we have for an incompressible medium (e = 0)

f (x,z) = 2cke (6.6b)

Substituting (6.6b) in (6.5a) and (6.5b)

93.

To eliminate f from the above equations,

-e5 2f 2 = ;7,7 -2) x-2) z

Ou tifou ZiwN • —+_ "oz\.oz .bx/

h X22

4. -6 2w) (6.8c)

2>z axuz

f 2f 2 '6 w h . + - - . 2h . tit (6.8d)

ax .bx-bz "ex -t) x"6 z

Substituting (6.8a) and (6.8c) in (6.8d)

h 2u+h(1211 + 26Y)+ - p2 z x L

bu + bir;) z x

h (b21, a2w

z2 a xaz

2(b - 2w + 2h .

xi x.b.z

from which we have

94.

(6.9a) The other relationship between the displacements can be obtained

from equation (6.3a)

.)14 +—.o d x z

(6.9b)

We now consider a semi-infinite medium in which the shear

modulus G increases linearly with depth according to the equation

(see Fig. 6.9):

G(z) = G(o) + mz (6.10)

where G(o) = shear modulus at the surface, z = 0.

Here

h = [ log G(z) dz

m

G(o) + mz

and h2 - m2

[G(o) + mz] 2

dh m2 = dz LG(o) + mz] 2

2 dh So h + = 0 dz

Equation (6.9a), then, reduces to

or

or

where

2

!wx 2h ,72u = 0

2 .Id ,uz) 2 . 2u = 0

0 x G(o) + mz

(2 ow _ Ou) _ 2 2u = o ax az (z+i;

A . G(o) m

Therefore, the two equations governing the displacements in an

incompressible medium with modulus increasing linearly with depth

95.

are: u w + 0 ) x z )

) (6.11) 21 w 2) v 2u = 0 )

ox az (z+ ) )

Equation (6.11), together with the boundary conditions,

C7 =- q on, z = 0 and - b b ) zz )

) (6.12a) )

= 0 on, z = 0 and xt>b )

Cr =0, on z = 0 and \xl>0 (6.12b) xz

96.

as z —) 00 (6.12c)

define the complete problem for a semi-infinite medium subjected to

a strip loading (width 2b) of intensity q.

With the use of Fourier sine and cosine transforms

Gibson (1967) solved the above equations and derived the following

expression for the component of vertical displacement:

00

W (x , q e4z il,(bi) CO (PCb

2,c G(o)

q2 J

0

1i-4 FM 4p p(cy)

F(4) ff3 109 (4) +1+

z9 — (6.13)

97.

where the function F is defined as

F( = e2 Ei( 2 - log

in which the exponential integral function

oo

Ei =

Ja

1 -6 e d

and = z -F13

The expressions for stresses in the plane strain problem are:

00 -Ez FfY [F( y) F (4)] -[F(0)-qfp).}

t+ 2Iy Lo9 ( 'f) + + Ci-iJY)/fp cif (6.143

00 -tz

r-fY +F ()] -EF(Y) -Ffq dc

(6.15) 1, 2 Lo9 (he) + 1 + (1-1yrn p

0*-Kz. K.; A

-C1 z e 1 fY [F(cY) -FM)]

0

98. Where K 2 Sinth ) cos(xl)

s = —

K' = 2 Sin(b ) Sin(x ) — s Tr

= F( () + f log(§ () + 1 + I 1 2 f

The analysis was extended by Gibson (1968) to give the expressions

for stresses and displacements for the axi-symmetric loading

(Dia..2b) of intensity q:

Displacements

1). (r, z) -f z f(3 F (pi) +1 + fp F.(ty)'

+ 2 ff3 [09 0.10 (6.17)

w 7-) F(fp) I (fyicif (6.18)

0

fY [F (C-r) + F(49] —EF (cY) - FU(3)] 8 1

(6.19)

Stresses

i- 2fy1.09 (fy) + 1 + (7 -1-FY)/f p

99.

Crrr li[F( i3)-F(i) 1 +4] 4.1 Ji(rt) 1 iF(q3)+F(t-y)

.J1 L r0-1 L-4- yip +zi_09(3y

(6.2o)

where Kc

z ta_ y[F(&Y)-0011

LT; = b Jo(rf ) Ji(bt )

oo Kc ' cr rz A

(6.21)

= b J1(rf ) J1(14 )

log (q (3) + 2 t

To express the above equations for stresses and displacements in

terms of non-dimensional quantities, the "dummy" variable

been replaced by Q such that

b =(34,

All the expressions now appear in terms of the quantity — b

are re-written in Appendix A.

6.7.2 Limiting values

Two limiting cases of the above problem are:

o0 -- ---> or m = 0 : the problem reduces to that

=

has

and

100.

of the homogeneous (Boussinesq) medium.

(ii) , or G(o) = 0 : the shear modulus at the

surface ,(z = 0) is zero.

The corresponding expressions are derived in Appendix B.

6.7.3 Numerical computations

The writer has used the equations obtained above to calculate

the stresses and displacement at selected points beneath strip and

circular footings.

(a) 00 Displacements : Homogeneous medium :

(i) Equation (6.31) (Appendix B) gives the expression for

vertical displacements in a homogeneous half-space beneath a strip

loading. It diverges for all values of z, giving the well-known

result that the settlement at all points in the medium is infinite.

The relative settlement of any point can, however, be obtained by

using a new origin attached to the point 0 and connecting with it,

which means subtracting the infinite constant

o0

--2h-- Sin (0C)

II G(0) o , 2

from equation (6.31). The relative settlement is then given by

w(2c- 1 21) - w(0, 0) b b 0.0

= __at__ Ti

. cos (zok G(0)

Sin co() o(2

b do(

101.

At

( ( ( ( ( ( ( ( (

x log z2 (b x)2 1 +

) )

) ) )

- x) ) z ) ,

(6.31a) =)

2 b b2

_ z2 4- (b2- x)2

4. 1 i _ a log

211.G(0) 2 b b -

+ 4 tan-1 (1) + -x)+ -z tan-1( -b b z z b z

The settlements calculated from the above expression are given in

Table 6.1 and the shapes of the deformed planes - = 0, 0.25, 0.5

and 1 are shown in Fig. 6.10.

(ii) The settlement of any point beneath an axi-symmetric loading

is, however, finite according to the equation (6.32) (Appendix B):

qb w(- b b 2G(o)

- b o 1 s J ( b c9 J (01 ) e .

c"--\ 0

(.1 + zoc) dcd,

The numerical values can be obtained from Ahlvin_ and Ulery (1962).

(b) Displacements : Limiting case : b --)o

(i) S-triE loading

x z 17 ) The displacement at any point '13 is given by

equation (6.33) (Appendix B),

w 2) = -l- b b m

- -13'4 . Sin ((DC) Cos tli °L) e ok . d 0( (6.33)

0

cZ

w (— .Jo rJ (cA ) . doz.,

b

r

0

—S. b 2m

04, b e

from which the surface settlement

00

102.

0) = —g- Sin (-b.() Cos(Sb 04.

(6.33a)

-a- , - 1 C C 1 2m

b

= 0 , 1-1>1 b

(from Selby and Girling 1965)

The numerical values of the settlement calculated from equations

(6.33) and (6.33a) are shown in Fig. 6.11 and tabulated in Table 6.2.

(ii) Axi -symmetric loading

From equation (6.34) (Appendix B) settlement

and at the surface,

w 0) = 2m 0 ,(

.1:0( (0C) do``0 1 b

= 0 c -11 1 2m

= 0 r> 1 b

(Abramowitz and Stegun 1965)

It will be seen that for both plane strain and axi-symmetric problems

103.

in the special case when G or E of the medium increase linearly

with depth according to the equation

G(z) = mz

the settlement of the surface is given by

w(x , 0 = , within the loaded area ) 2m

) = 0 , outside the loaded area )

Since the principle of superposition holds this result will also

apply to a uniformly loaded area of any shape.

The above relationship that settlement at the surface is

directly proportional to the load intensity and independent of the

dimensions of the loaded area is also identical with the behaviour

of the Winkler spring model (Winkler 1867, Gibson 1967). The concept

of the coefficient of subgrade reaction (Terzaghi 1955) defined as

k = w 0

and used in many foundation problems has thus a theoretical basis.

It is applicable, however, only to a medium which is incompressible

= 1) and whose modulus of elasticity E increases linearly 2

104.

with depth from zero at the surface according to the equation

E(z) = 3mz

The coefficient will, then, have a value equal to 2m.

(c) Displacements : 0<

The vertical displacements of any point in a non-homogeneous

medium subjected to strip or axi-symmetric loading are given by

equations (6.22) and (6.26) (Appendix A). The settlement of the

surface can be obtained by putting i73- = 0 :

= Sin (04.) Cos( Stri w 21 d °I\

( 1 ) ( p: 211G(0) 0 04, 2 A

ok) Circle: w (7- , = qb fr4) ji

b 4G(0)

(1) d c A

L.( v- ... log —a. +1 + b \ b 04, \ b 2 1-7-

where A

The settlements calculated from the above equations are tabulated

in Tables 6.3 and 6.4 for — = 0.1, 0.5, 1.0, 5.0, 10.0. The

shapes of the deformed surface are shown in Fig. 6.12 and 6.13.

(d) Stresses : 0

It has already been mentioned that in the limits -s- 0

and 0.0 all the stresses for both strip and circular loadings have

identical values (see Appendix B). The stresses for intermediate

105.

values of — (0 K C-- ) can be obtained from equations (6.23)

-(6.25), (6.28) - (6.30) (Appendix A).

Tables 6.5 - 6.8 give the components of vertical and

lateral stresses calculated from equations (6.23), (6.24), (6.28)

and (6.29) beneath the centre and edge of strip and axi-symmetric

loadings for the values of .14- = 0.1, 0.5, 1.0, 5.0 and 10.0. The

results are plotted in Figs. (6.14) - (6.18) and the variation of

stresses with the parameter b shown in Figs. (6.19) and (6.20).

Some notes on the method of calculation can be found in Appendix C.

6.764 Discussion of results

From the numerical values of stresses and displacements

presented in the previous section a number of important results

emerge.

(1) As mentioned already, a semi-infinite incompressible medium

whose modulus of elasticity increases linearly with depth from zero

at the surface (i.e. () = 0) behaves as a Winkler spring model. In

other words, the surface settlement of a uniformly loaded area on

such a medium is directly proportional to the applied pressure and

independent of the dimensions of the load.

(2) The distribution of stresses in a semi-infinite medium

is not significantly affected by the type of non-homogeneity considered

the analysis. Indeed, the two limiting cases, = 0 and

7 = oo provide exactly the same stresses, while in the intermediate

106.

range 0 C --b- C-00 , both the vertical and horizontal stresses tend

to be a little higher* than the corresponding stresses for the homo-

geneous medium, though the difference is never greater than 10%.

Over most of the range, however, (0 < <7 0.5 and 5; )

the discrepancy is less than 5%. This observation is true of

vertical and horizontal stresses due to both axi-symmetricand strip

loadings.

Extensive field measurements of stresses reported by

Plantema (1953), Turnbull et al (1961) and the Waterways Experimental

Station (1953, 1954) have shown very close agreement with the pre-

dictions based on Boussinesq analysis even though the soil media

varied from non-homogeneous deposits (Plantema) to fairly homogeneous

test sections of clayey silt and sand (Turnbull et al). This is

certainly in agreement with the theoretical results presented here.

(3) If the foregoing conclusions are correct, the settlement

of a foundation on a non-homogeneous medium may be calculated

reasonably accurately by assuming that the stress distribution could

be obtained by Boussinesq analysis. Fig. 6.21 shows a comparison

between the actual settlements of the centre of a uniformly loaded

circle obtained from rigorous computation and the approximate settle-

ments calculated from Boussinesq stress distribution. The latter

* Except the horizontal stresses beneath strip loadings for high values of

107.

underestimates the settlement for all values of though the

maximum error (in the range 0.5 C IT. < 1.5) is no more than 10%.

(4) In order to obtain some indication of the effective depth

that contributes towards most of the settlement beneath a loaded

circle, Fig. 6.22 has been constructed by successively integrating

the vertical strains for various depths, using Boussinesq stress

distribution. It is observed that 80% of the total settlement is

contributed by a depth of only 1.5b for = 0.1 and 5b for

= dJ

b

6.8 Non-linear soil medium

The problem of stress - analysis in a soil medium with a

non-linear stress - strain relationship is immensely complex and

no analytical solution is yet available. Only a few attempts have

been made to use numerical methods to calculate the stresses and

displacements in such media.

Phukan (1968) while investigating the non-linear deformation

of rocks used the finite element method to analyse the stresses and

displacements beneath a strip loading for two particular types of

non-linearity. His results are shown in Figs. 6.23 and 6.24.

while a depth of 8b accounts for as much as 90% for

0 all values of -T. This is consistent with the stress-distribution

in a two-layer rigid base system (see section 6.4.2) where, it has

been shown, the presence of a rigid layer at a depth of 8b does

not significantly affect the stresses.

108.

Phukan found that the two different models produce almost identical

vertical stresses which agree remarkably well with the homogeneous

elastic (Boussinesq) solution while there is a difference of about

10% in the corresponding horizontal stresses and 5% in the shear

stresses. The settlements, however, are no longer proportional

to the applied pressure (Fig. 6.24).

Huang (1968) presented a method of analysing stresses and

displacements beneath a circular load in a non-linear soil medium

whose modulus of elasticity is a function of the stresses:

E = Eo [ + C (Crz (Zrr 0-0 + C )S . LE)]

where (the non-linear coefficient) and C (the body force

coefficient) are material parameters. He divided the semi-infinite

medium into a multi-layer system assuming a rigid base at a depth

of 100b (see Fig. 6.25) and assigned to each layer a modulus

corresponding to the stresses at the mid-point. Employing Burmister's

boundary and continuity conditions (Burmister 1943, 1945) and using

thegiethod of successive approximations Huang calculated the stresses

and displacements until two consecutive iterations gave the same

modulus. His results are shown in Figs. 6.25 and 6.26. Again a

close agreement with the Boussinesq stress distribution is noted.

A comparison between the actual settlements calculated rigorously

by Huang and the approximate settlements calculated by the witet

109.

on the basis of Boussinesq stress distribution but using the proper

variation of E with depth (Fig. 6.26) shows that the two methods

do not differ by more than 10%.

The assumption made by Huang, however, that each layer has

uniform modulus means that the problem is, in effect, reduced to a

multi-layer Burmister Problem with the elastic modulii determined

by the stresses in the centre of the individual layers as shown in

Fig. 6.26.

6.9 Summary

From a study of the distribution of stresses in soil media

presented in this chapter the following general conclusions can be

drawn:

(1) The presence of a rigid layer at a depth which is relatively

shallow compared to the dimensions of the loaded area drastically

alters the stress - pattern, particularly the lateral stresses.

With increasing depth, however, the effect decreases and if the

rigid layer is at a depth greater than 4 x diameter (for a circular

footing) the soil will behave as a semi-infinite medium in respect

to stress distribution.

*(2) From the practical point of view the most important result

to emerge is that, for a non-homogeneous, incompressible medium whose

modulus of elasticity increases linearly with depth, the stresses due

to a boundary load do not differ significantly from those given by the

110.

classical Boussinesq analysis. The same conclusion has been reached

by Phukan and Huang for cettain types of non-linearity of stress -

strain relationships described in the preceding section.

It appears, therefore, that any deviation from the classical

problem of homogeneous elasticity, either in terms of non-linearity

of stress - strain relationship or in terms of non-homogeneity of the

medium will only have a marginal effect on the stress distribution so

long as the medium is subjected to certain boundary stresses.

The displacements will, of course, be significantly affected but

from the foregoing conclusion it can be deduced that the settlements

can be approximately obtained numerically by-assuming the Boussinesq

stress - distribution and taking account of the proper stress -

strain relationship and/or non-homogeneity in calculating the

strains. Calculations carried out by the writer, presented pre-

viously, confirm this deduction.

Finally, the above study is by no means complete. A

number of important factors have not been considered which may yet

have a significant effect on the distribution of stresses in soil

media e.g. anisotropy, Poisson's ratio other than 0.5( i.e. com-

pressible medium) and the rigidity of the foundation. Analysis of

a problem taking any or all of these into account is complex but

could, perhaps, be performed numerically.

Table 6.1. Relative Settlement Uniformly loaded strip (Homogeneous medium)

w(0, 0) -w U-DC , 17) I cp

)1i 0 .25 .5 .75 1 1.5 2 3 4 5 ii

. ,

0 0 401 .042 .098 - .420 .525 .662 .756 .828 .25 .113 .123 .151 .202 .280 .430 .527 .662 .756 .828 .5 .212 .236 .243 .292 .335 .456 .537 .666 .760 4829 1.0 .360 .365 .377 .402 .432 .500 .568 .674 .766 .835

Table 6.2. Surface settlement Uniformly loaded (Non-homogeneous x z w b , 13) = I

strip

medium: G(z) = mz )

(a.) ° 2m

N x

TO. 0 .25 .5 .75 1.0 1.5 2 3 4 5

0 1.0 1.0 1.0 1.0 1.0 0 0 0 0 0 .125 .917 .914 .895 .829 .479 .062 .026 .010 .005 .003 .25 .843 .827 .780 .705 .460 .135 .054 .020 .010 .007 .5 .704 .691 .647 .594 .420 .187 .094 .040 .022 .013 1.0 .500 .489 .460 .413 .352 .231 .149 .069 .039 .025

112.

Table 6.3. Surface settlement Uniformly loaded strip Non-homogeneous medium : G(z) = G(o) + mz

W(ig o) -'r

X b o 0.25 0.5 1.0 2 4

b

0.1 0.045 0.043 0.043 0.025 0.008 _

0.5 0.171 0.17 0.165 0.103 0.019 0.004 1.0 0.270 0.269 0.263 0.174 0.053 0.022 5.0 0.68 0.626 0.484 0.234 0.096 lox 1.00 0.962 0.673 0.478 0.188

Table 6.4. Surface settlement Uniformly loaded circle Non-homogeneous medium : G(z) = G(o) + mz

r w , = I x {do)

\\\N ,;.N

0 0.25 0.5 1.0 2.0 4.0 b

0.1 0.046 0.041 0.041 0.021 0.0012 - 0.5 0.146 0.135 0.130 0.071 0.008 - 1.0 0.21 0.196 0.189 0.107 0.016 0.0012 5.0 0.367 0.350 0.33 0.204 0.052 0.0115 10.0 0.420 0.403 0.386 0.24 0.071 0.0195 op 0.504 0.491 0.465 0.318 0.129 0.0625

G(o)

Table 6.5. Uniformly loaded circle : centre

(a) Vertical stress c:r 1 Influence values I o z

\',\\ 1)- b 0 0.1 0.5 1.0 5.o 10.0 x)

0 1.0 1.0 1.0 1.0 1.0 1.0 1.0 0.5 0.910 0.890 0.927 0.936 0.935 0.912 0.910 1.0 o.646 0.662 0.696 0.696 0.679 0.655 0.646 2.0 0.285 0.302 0.317 0.319 0.315 0.300 0.285 4.o 0.0854 0.0874 0.0948 0.0966 0.0955 0.0928 0.0854 6.o 0.0397 0.0403 0.0435 0.0451 0.0453 0.011110 0.0397 8.o 0.0225 0.0228 0.0243 0.0250 0.0258 0.0252 0.0225

(b) Radial stress Cr : Influence values I zr r r

0 1.0 1.0 1.0 1.0 1.0 1.0 1.0 0.5 0.382 0.397 0.412 0.402 0.381 0.382 1.0 0.114 0.132 0.150 0.149 0.132 0.131 0.114 2.0 0.0164 0.0191 0.0265 0.0279 0.0229 0.0203 0.0164

113.

114.

Table 6.6. Uniformly loaded circle : edge

(a) Vertical stress C-z : Influence values Ic5-z

.\\\

- 73

0 0.1 0.5 1.0 5.0 10.0 op

0 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.416 0.413 0.420 0.420 0.420 0.414 0.416 1.0 0.336 0.336 0.352 0.352 0.343 0.340 0.336 2.0 0.196 0.199 0.213 0.214 0.212 0.205 0.196 4.0 0.075 0.077 0.082 0.084 0.083 0.081 0.075 6.0 0.037 0.038 0.041 0.042 0.042 0.042 0.037 8.0 0.0216 0.022 0.0233 0.0240 0.0250 0.0244 0.0216

(b) Radial stress Cr : Influence values I 0

' - 0 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.285 0.277 0.288 0.284 0.275 - 0.285 1.0 0.150 0.147 0.167 0.167 0.157 0.152 0.150 2.0 0.041 0.044 0.052 0.053 0.048 0.046 0.041

01 = q I crz

o-, = q I Tr.

Table 6.7. Uniformly loaded strip : centre

(a) Vertical stress crz : Influence values I 0

0 0.1 0.5 1.0 5.0 10.0 oo

co O.

o o

OO

OO

OIn

n 1.0 1.0 1.0 1.0 1.0 1.0 1.0 0.952 0.944 0.959 0.966 0.959 0.947 0.952 0.818 0.817 0.848 0.843 0.845 0.829 0.818 0.551 0.553 0.582 0.583 0.583 0.578 0.551 0.306 0.308 0.321 0.326 0.327 0.323 0.306 0.207 0.210 0.218 0.221 0.228 0.223 0.207 0.158 0.159 0.164 0.167 0.171 0.169 0.158

(b) Lateral stress 07 : Influence values icr

0 1.0 1.0 1.0 1.0 1.0 1.0 1.0 0.5 0.452 0.448 0.445 0.427 0.412 - 0.452 1.0 0.183 0.194 0.203 0.196 0.177 0.170 0.183 2.0 0.043 0.049 0.057 0.059 0.044 0.039 0.043

q I CrZ

CI 1 Crr

115.

Table 6.8. Uniformly loaded strip : edge

(a) Vertical stress 0-z : Influence values 10-z

,* 0 0.1 0.2 0.3 0.4 0.5 0.6 Ta-

Lf1 0

0 0

0 0

••

••..

0 0

r- N —I- %.0

c0

1

0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.502 0.499 0.506 0.506 0.510 0.506 0.502 0.486 0./48 0.491 0.488 0.495 0.493 o.486 0.410 0.406 0.422 0.421 0.428 0.422 0.410 0.274 0.277 0.285 0.287 0.292 0.288 0.274 0.199 0.200 0.207 0.210 0.214 0.212 0.199 0.152 0.153 0.159 0.161 0.165 0.165 0.152

,

(b) Lateral stress cr : Influence values I 0-

0.5 0.342 0.329 0.314 0.300 0.287 0.342 1.0 0.228 0.223 0.224 0.220 0.206 0.203 0.228 2.0 0.093 0.100 0.104 0.101 0.093 0.092 0.093

az

(r q (Tx

116.

117.

CHAPTER 7

STRESSES DURING CONSOLIDATION IN THE FIELD

7.1 Development of pore pressures in saturated clay

The application of a structural load causes the total

stresses in the ground to increase, the magnitudes of which can be

determined by methods described in Chapter 6. If the subsoil

consists of clay of low permeability and construction is sufficiently

rapid, these changes in total stresses occur under conditions of

no volume change and are associated with simultaneous development

of excess pore water pressures which subsequently dissipate with

the passage of time.

Skempton (1954) developed the expression for pore pressure

change within an element of soil in undrained conditions, for axially

symmetric stresses ( A cs---2 0-- = A ) in terms of the pore pressure 3 coefficients A and B,

u = B [Q AO-3 + A(A C5-1 - O cr (7.1.1)

where Lc'1 and ti.0'3

are respectively changes in the total

major and minor principal stresses.

For saturated clays, B = 1. Equation (7.1.1) then,

reduces to

!em u= A 3 + A(.6 - 1 3) (7.1.2)

Au=B[Acr1 - (1 - A)(Acr1 - 3

1 - (1 - A) - -L-1-2)- \L

or u = = B 6(5-

118.

This relationship between total stresses and excess pore

pressure does not take into account the value of the intermediate

principal stress Lcr2 and is strictly applicable only to cases

where A.0-2 = LIcr-3, and the stress increments are in the prindipal

stress directions. The validity of this relationship in the field,

where such conditions are satisfied, e.g. beneath the centre of a

uniformly loaded circle, has been demonstrated by Gibson and

Marsland (1960) and Lambe (1962)* who found that the in-situ

measurements of pore water pressures agree well with the predictions

based on the laboratory determination of the parameter A.

For problems of earth dams, Eqn. (7.1.1) is more con-

veniently expressed as

The "overall" coefficient E which expresses the excess pore

pressure in terms of the major principal stress is useful in the

stability analysis of earth dams, particularly in conditions of

rapid draw-down (Bishop 1954, 1957). The prediction of pore

Lambe investigated the case of a preload of the shape of a truncated cone.

119.

pressures according to equations (7.1.3a) and (7.1.3b) and based

on laboratory determination of B has been found to be in good

agreement with many field measurements (Sheppard and Aylen 1957,

Nonveiller 1957, Bishop and Vaughan 1962, Delory, Gass and Wong

1965, Rivard and Kohuska 1965).

It will be evident from equation (7.1.5b) that the

magnitude of excess pore water pressure is, in general, influenced

6 °---3 -,by the stress increment ratio . In certain conditions, 1

however, such as for a saturated normally consolidated clay, all

the quantities A, B and B approach unity and the ratio cr. has no,: more than a marginal influence. The pore pressure developed

is, then, approximately equal to the increase in the major principal

stress etoung and Osler 1965).

The basic pore pressure equation (Eqn. 7.1.2) for a

saturated clay under axi-symmetric stress conditions can also be

expressed in terms more comparable to the theory of elasticity

/:111 = 1.(e_Scr1 3 + 260-3) + (A - 1.)(6c 1 - acr3

) (7.1.4) 3

1 For a perfectly elastic material (A = 3 ' the change in pore

pressure is equal to the average increase in the principal stresses.

In practice, however, the condition of axial-symmetry is

not often satiefied, i.e. ac--2 6. 0-3. For such cases, to take

account of the influence of the intermediate principal stress, a

120.

more general relationship, in terms of the stress invariants has

been proposed (Skempton 1960, Henkel 1960):

u 21-3(zI cr1 2 3 +Lao- + au-)

+01. -Ac72)2 + (40-

2 3 -A 3)2 + (6cr3 -6.05-)2

(7.1.5)

It can be readily seen that, for the particular case of axial

symmetry (a d'2 Acr3 cr or A -1 6cr2). Equation (7.1.5)

reduces to

:317(6(7-1 + 60-2 Acr-3) +cK A/2 (acr-1 -6a- 3)

(7.1.6)

Although it is evident that equation (7.1.5) is of more

general validity than equation (7.1.1), (Hvorslev 1960), experimental

evidence is, so far, too limited (WU et al 1963, Shibata and Karube

1965) to demonstrate its applicability to all types of soil.

7.2 Pore prqssures beneath a circular foundation

Let us now restrict our consideration to the elements

of a saturated soil beneath the centre of a uniformly loaded circle.

Here, by virtue of axial symmetry, the vertical and horizontal

(radial) stresses are the principal stresses. Equation (7.1.6)

can then be written as:

121.

= Acr + A(&cr _ Lcr ) (7.1.9)

where tS0-v

increase in vertical stress

A.0-h = increase in horizontal strew

For over-consolidated clays where the in-situ horizontal stresses

prior to load application may be the major principal stresses,

equation (7.2.1) still holds in so far as only stress increments are

concerned.

The distribution of 460-v and 4crh

in a homogeneous

or a non-homogeneous elastic medium can be obtained by the various

methods of analyses described in Chapter 6. It has been shown

that the distribution of stresses in a iron-homogeneous half-space

is not significantly different from the Boussinesq stresses for a

homogeneous medium. We shall, therefore, use the latter in the

analysis that follows.

Fig. 7.1 shows the distribution of pore pressures in an

incompressible medium - Poisson's ratio = z - beneath the

centre of a uniform circular load, for different values of the

parameter A. It is interesting to note that for the two cases

A = 0 and A = 1 the changes in pore pressures are respectively

equal to the changes in horizontal and vertical total stresses.

If the parameter A changes with depth it is possible to determine

the distribution of A u by taking the variation of A with

depth inco account.

122.

7.3 Stress changes during consolidation

It has been described in Chapter 5.4 that during load

application most of the deformation takes place under conditions

of no volume change, implying a value of Poisson's ratio throughout

the medium equal to 0.5. As the excess pore pressures dissipate,

s'\) decreases and finally drops to its fully drained value

at the end of consolidation. It is well known that for an isotropic

homogeneous elastic medium the vertical stresses are independent of

the elastic parameters and are, therefore, unlikely to be significantly

affected by this decrease in On the other hand Poissonts

ratio has a considerable influence on the horizontal stresses as

can be seen from Fig. 7.2 where radial stresses are plotted as a

function of for the Boussinesq case beneath the centre of a

uniform circular load.

It is clear, therefore, that the problem of consolidation

in the field is very much interlinked with the problem of stress

distribution, a rigorous analytical treatment of which is extremely

complex and only a few specific solutions have so far been obtained

(Gibson and McNamee 1957, 1963, de Josselyn De Jong 1957, McNamee

and Gibson 1960). One of the purposes of the present investigation

is to study the influence of the horizontal stresses on the process

of consolidation and for this an approximate analysis will be made

on the assumption that the horizontal stress at the end of con-

solidation will be the same as the Boussinesq stresses for the

appropriate value of 11 =

With this in mind, let us now consider an element of

clay at a certain depth where, during load application,

A‘r-v increase in vertical stress ) for

At-7-111 increase in horizontal stress

The corresponding change in pore pressure will be

u = cr-h1 A("-v c5-h1)

(7.3.1)

At the end of consolidation, the vertical stress increment will

remain unchanged and the horizontal stress will have decreased by

an amount

_ (.3-h2 h1 (7.3.2)

where Cr = increase in horizontal stress for -) =

So, the changes in effective stress during consolidation will be

vertical: Acr =Au v c horizontal: A(5)

rc = L.12 S

) (7.3.3)

The ratio of the effective stress changes (K') is then given by

6_cr I K' = h c - 1 -

A c>" Au v c (7.3.4)

123.

(1 + 2 ) (17z ) 3 —

z3 b3 (7.3.6)

3 1

L.

-z-b z2 )3/2

1 +b

(7.3.5) 0- = q

124.

(A fuller discussion of this point in relation to the stress path

appears in Chapter 5).

Now, for the Boussinesq problem the general expressions

for stresses beneath the centre of a uniform circular load

(diameter 2b, load intensity q) in polar co-ordinates are,

(Wu 1966),

which can be more conveniently written as

Ciz = q [1 - 1 1 (7.3.5a)

cril = [(1 + 211) - 2(1 +\) ) + 11 3 (7.3.6b)

where 1 z b

(1 + b2

Using these expressions in our problem,

125.

)

6o-v = (c5 2). = = q(1 - '13) ) ) ) )

La-h1 = (611-1 )- = = •g• (2 - 31I 1 + 1 \ 3) ))(7.3.7) i 2 • ) )

(fr-h2 = ( a- )1= 2 1 L(1 + ' - 2(1 + ') .1 +11)

h V )

Combining equations (7.3.1), (7.3.2), (7.3.4), and (7.3.7) we have

K' = 1 - 3 + [(1 + 2-N)I) - 2(1

2 [(2 - 311 + 13) ± 4(2 - 3 ) (2 - + 3 )

which, on simplification, reduces to

(7.3.8) + ) (3A - 1)

Equation (7.3.8) can then be used to determine the ratio of the

stress increments during consolidation at any depth beneath the

centre of a uniformly loaded circle. Fig. 7.3 gives the relation

between K' and - as calculated from eqn. (7.3.8) for a wide

range of values of A and 11 .

It will be noticed that in Fig. 7.3 the values of K'

are plotted for values of up to 3.0 only. Beyond this depth

the horizontal stresses imposed by a circular load are insignificant

and the ratio K' may be taken as 1.

K' =1 - 1

126.

The values of K' plotted in fig. 7.3 will be used to

study the experimental data and in settlement computations later

in the thesis.

127.

CHAPTER 8

SAMPLING, PRELIMINARY MEASUREMENTS AND

EXPERIMENTAL PROGRAMME

8.1 Location of sites

All the experimental work reported in this thesis was

conducted on undisturbed samples of London clay from two different

sites - Oxford Circus and High Ongar, Essex, (see Fig. 4.1). Two

block samples were obtained from the Victoria Line underground

tunnel at Oxford Circus in March 1965 and the first series of tests

were performed during the next nine months. Because of complica-

tions with the tunnelling operations it was not possible to return

to Oxford Circus for more samples and the second site in Ongar,

Essex, was chosen. Three block samples were obtained from this

site in March 1966 on which the second series of tests were con-

ducted until the completion of the experimental programme in Nov-

ember 1967.

8.2 Description of sites, sampling and storage

(i) Oxford Circus

The geological section of the site (obtained from Sir

William Halcrow and Partners, Consulting Engineers) is shown in

Fig. 8.1a. The London clay is overlain by 41.1 ft. of gravel, the

water level being 9.6 ft. above the top of the clay. The samples

128.

were obtained from a depth of 83.6 ft. below ground level l , i.e.

42.5 ft. into the London clay.

9 in. x 9 in. x 9 in. samples were trimmed with sharp

spades from approximately 3 ft. cube blocks detached from the face

of the tunnel by drills and a portable chain saw. The samples,

which were marked previously for cbrientation, were covered with

polythene bags and sealed in suitable size tins with Scotch tape

before they were transported to the laboratory.

(ii) High Ongar

The site in Ongar, used by Messrs. W. F. French Ltd. for

the manufacture of aggregate from burnt clay, consists of a pit of

London clay overlain by various mixtures of gravel, ballast and

brick earth (Fig. 8.1b). The samples were obtained from the

bottom of a cutting 27.5 ft. below the top of the clay. The upper

part of the soil profile is somewhat approximate in that it had to

be extrapolated from a series of borings made by W. F. French Ltd.

in the surrounding area a few years earlier.

Three columns of clay approximately 18" x 18" x 12 ft.

high were first isolated from the mass of clay with picks and

trenching tools and left attached to the base of the pit. These

were trimmed carefully by hand with sharp spades and knives to

approximately 9 in. square x 1 ft. high, after which the blocks

were covered with tinfoil and polythene bags, and suitable size

boxes were slid over from the top. The samples were then separated

129.

from the base by a wire cutter, sealed with Scotch tape and trans-

ported to the laboratory. The whole operation took the best part

oT a day.

Storage

Immediately on arrival in the laboratory the samples were

taken out of the boxes and given a thick and thorough coating of

wax and petroleum jelly mixture. These were then covered in poly-

thene bags and sealed, put into the boxes and lids replaced. The

blocks which were to remain untouched for a long time (no. 2 from

Oxford Circus and nos. 2 and 3 from Ongar) were covered with an

additional layer of wax poured on top of the sealed samples before

the lids were placed.

When specimens were needed for testing, small prisms

(2" x 2" x 4" for triaxial samples and 4" x 4" x 1-i" for conventional

oedometer samples) were cut out with thin wire cutters and the rest

of the block was re-coated with wax and stored, as described above.

8.3 Index properties

The Atterberg limits, the specific gravity of the solid

particles and the clay-fraction were determined for a number of

samples following procedures given in B.S.1377(1961). Table 8.1

summarises the average data for each block. It may be noted that

the clay fractions for both Oxford Circus and Ongar were substantially

the same, yet the Ongar clay was more active because of its higher

130.

plasticity.

A comparison of these values has been made in Fig. 4.3

with the corresponding properties of London clay from a number of

other sites.

8.4 Moisture content

The natural moisture contents have been determined from

initial measurements made on pairings taken during preparation of

individual triaxial and oedometer specimens. In Fig. 8.2 these

values are plotted against time after sampling. It will be seen

that the scatter is large, but not untypical of London Clay (Ward,

Marsland and Samuels, 1965). There is, however, no systematic

decrease of water content with time, as would be the case if there

was loss of water by evaporation due to poor sealing and storage

over such a long period of time (up to 600 days).

Also shown in Fig. 8.2 are the field moisture contents

determined from bulk samples transported to the laboratory in air-

tight bottles. These show no significant difference from the

laboratory moisture contents and lie within the range of scatter,

although being a little above (0.5%) the average.

The inital degree of saturation of n11 the triaxial and

oedometer specimens calculated from their weight and dimensions,

is shown in Tables 8.2 and 8.3. The samples from the Oxford Circus

blocks indicate almost 100% saturation while the later triaxial

131.

specimens and some oedometer specimens of the Ongar clay show a

slightly lower degree of saturation. One of the reasons for this

discrepancy must be the noticeably greater opening of fissures of

the Ongar clay during preparation of the specimens while the Oxford

Circus clay tended to remain more intact. It was often difficult

to push an oedometer ring into the Ongar clay without breaking it

up along fissures and joints. Also, the calculation of the degree

of saturation based on an average value of the specific gravity may

be somewhat in the error for individual specimens containing various

amounts of silty material with different specific gravities. How-

ever, on the average the degree of saturation was between 98-100%

and the clay could, therefore, be considered fully saturated.

8.4 Stresses in the ground and after samplinE

The in-situ vertical stresses of the samples as calculated

from the soil profiles shown in Fig. 8.1 are given below.

Total vertical stress ( cr,)

lbs/in2 T/ft2

Pore-water* pressure (u

o)

lbs/in 2 T/ft2

Effective vertical stress

(p) 2 lbs/in T/ft2

Oxford Circus 69.2 4.45 22.6 1.45 46.6 3.00 High Ongar 48.5 3.11 16.5 1.06 32.0 2.05

Assuming a hydrostatic increase in porewater pressure below the ground water table.

132.

The expression for the state of stress in a soil after

sampling has been derived by Skempton and Sowa (1963).

Let us consider an element of saturated clay in the ground

under a total vertical stress rr and two equal horizontal stresses - v

h If the pore pressure is V

o the vertical and horizontal

effective stresses are, respectively,

p= Tv - uo

K p = - u 0 - h o

) (8.4.1)

where Ko is the coefficient of earth pressure at rest, in terms of

effective stresses.

When this element is removed from the ground, as a sample,

without any change in water content, it is relieved of its total

stresses and a negative pore pressure uk is set up. Since the

total stresses at this stage are equal to zero, the sample must be

under an all-round effective stress pk numerically equal to uk.

From equation (7.1.4) in chapter 7.1 the change in pore

pressure due to sampling can be expressed as

uo = (C7v crh) (As - 31)( Cry

= - Fp . 1 + 2K

o

3 1 + u0 + (A 3 s - —)(1 o)p

or pk=p K

0 -A s (K0 1)

=K - A (K -1) o s o

and

K

(8.4.3)

(8.4.4)

(8.4.5)

Pk or

p

133.

uk = - P

1 + 2:o (As - i)(1 Ko) I

(8.4.2)

or

3

where As is the pore pressure parameter corresponding to the re-

moval of shear stresses.

Therefore,

Pk = uk = [ i 3 0 (As

1 - Ko)

1 + 2K

Pk Fig. 8.3 shows the relation between Ko and — for different

values of A. It may be noted that for normally consolidated clays

(K0 1), pk is less than 1, i.e. the effective stress after

sampling is less than the in-situ vertical effective stress. For

heaving over- consolidated clays, however, (K0 > 1) pk will be

greater than p for all values of As (See also Noorany and

Seed 1965).

It is also of interest to note that for a material that

1 behaves elastically As 3 = - and therefore,

P

Pk = 1 3, 2K) (8.4.6)

134.

i.e. the effective stress in an undistrubed sample of such a soil

is equal to the average of the in-situ effective stresses.

The above analysis assumes "perfect" sampling which means

that the eleme-it has neither undergone any change in water content

nor has it sufered any mechanical disturbance during the process

of removal fro;: the ground. In actual samples, however, varying

degrees of disturbance may result in effective stresses which could

be different from those in "perfect" samples.

Measurement ofpk

The direct measurement of the initial suction (which

is numerically equal to the effective stress P of an undisturbed

sample is difficult unless it is less than one atmosphere. Lambe

(1961) has described several methods for measuring such small values

of u. But for deeper samples of London clay which have been under

very high in-situ effective stresses the negative pore pressures

may be too high to measure directly. For such cases of saturated

samples S,:empton (1961) LaL; suggested four methods for the determina-

tion of u,

(i) F,:.cm -no consojdation stage of drained or consolidated

undradned triaxial tests

The cell. press) re at which a sample neither swells nor

consolidates gives ',he value of u . This can be obtained by

plotting cell pressure against volumetric strain for individual

samples and inLerr:;lating the pressure corresponding to zero volume

135.

change. This method has been used by Bishop, Webb and Lewin (1965).

(ii) From oedometer tests

The so-called "swelling" pressure measured in the oedometer

by finding the load which prevents any volume change of a specimen

submerged in water is sometimes considered to give the initial suction.

However, as will be shown later this method has certain drawbacks

and is less accurate.

(iii) From undrained test with pore pressure measurement

A specimen is subjected to an all-round pressure n7 3

which is higher than the initial suction. If the pore pressure

measured under undrained conditions is u, then the effective

stress in the sample will be given by pk = (13 - u (since B = 1)

(iv) The initial suction can also be deduced from the un-

drained strength determined from conventional quick trig dal tests.

A knowledge of c', 0' and the pore pressure parameter Af at

failure is necessary.

London Clay from Oxford Circus and Ongar

The initial suctions of the Oxford Circus and Ongar clays

were determined by method (iii) mentioned above. Each triaxial

specimen was set up under undrained condition and a cell pressure

higher than the estimated suction was applied. The pore pressure

was measured over a period of 24 hours or more, until equilibrium

was reached. The effective stress (pk) in the specimen was then

given by the difference between the cell pressure and the pore

136. water pressure. Detailed procedures of these measurements are

described in Chapter 11.

The reliability of this method of measuring pk has been

demonstrated by Mr. A. E. Skinner (Skinner 1967) at Imperial College.

He consolidated two 4" dia. x 4" high samples of London clay in the

triaxial cell to all round effective stresses of 70 and 119 lbs/in2

respectively, then reduced the total stresses to zero to simulate

sampling, removed the samples from the cell and cut two smaller

specimens approximately 1.5" dia. x 1.5" high without change in water

content. The effective stresses of these specimens were then

measured.

Sample No.

Before Sampling After Sampling

Sample -

Specimenjai) Size of

Back pressure

ub (psi) p

(psi) Size of Specimen

Cy' 3

(psi)

Measured u

(psi) p,

(psi) P k p

B7/1 4.03"dia x 4.24"

100 30 70 1.48"dia x 1.505"

100 31.5 68.5 0.98

B7/9 4.03"dia x 4.02"

149 30 119 1.52"dia x 1.565"

149 41.1 107.9 0.91

The measurements are in good agreement with the theoretical prediction

(Fig. 8.3) that for samples consolidated under isotropic stresses

Pk (C0 = 1) the ratio — after sampling should be 1.0. P •

Attempts were also made to determine pk from the

oedometer tests set out in method (ii) above. But as will be

explained in chapter 10 this method had several drawbacks and did

not give reliable results.

137.

Fig. 8.4 and Table 8.3 show the measured values of pk

from all the triaxial specimens. Because of a limited supply of

samples only 8 tests were possible on the Oxford Circus clay. On the other hand a total of 36 tests were performed on the Ongar clay

over a period of about 600 days. The average values and the

standard deviations are also shown in Fig. 8.4. It may be seen

that there is no significant change in the value of pk with time,

again confirming that there has been no loss of water by evaporation

during storage. Table 8.4 summarises the results and shows that

pk is considerably higher than the vertical effective stress p

supporting the findings on the London clay from Bradwell (Skempton

1961) and Ashford Common (Bishop, Webb and Lewin 1965). Measure-

ments on normally consolidated Kawasaki clay and lightly over-

consolidated Boston Blue clay at M.I.T. (Ladd and Lambe 1963) have

shown that excessive sample disturbance may lower the measured value

of pk to 30% of the theoretical value for perfect sampling.

However, block samples are thought to suffer considerably less

disturbance than ordinary piston samples.

It is now possible, using equation 5, to determine the

in-situ horizontal stresses of the samples from Oxford Circus and

Ongar. A value of As = 0.4 corresponding to unloading tests

(reported in chapter 11) was used in the calculations. The

results are given in Table 8.4 which also includes for comparison

some results from Ashford Common and Bradwell.

138.

It has already been noted (Fig. 8.3) that pk is highly

dependent on Ho. For a heavily over-consolidated clay, like

London clay, Ko is essentially a function of the overconsolidation

ratio (O.C.R.). The higher the O.C.R., the greater is the value

Pk of Ko (Brooker and Ireland 1965) and also of the ratio — . This

Pk is clearly shown in Fig. 8.5a where — has been plotted against the

O.C.R. on a log scale for both Bradwell and Ashford Common. There

is good agreement between the two sites, i.e. Bradwell and Ashford

Common, and the value for Ongar lies close to the average line.

Substantially the same data are plotted in Fig. 8.5b which shows the

Pk decrease of — with the vertical effective stress p. Here, of

course, two separate lines are obtained reflecting the difference in

the maximum past overburden pressures at the two sites.

A more complete account of the in-situ to of London Clay

will be found in chapter 10 where the results of laboratory determina-

tions of Ko are presented.

8.6 Experimental programme

The main purpose of the laboratory investigation reported

in this thesis was to study the compressibility of London clay under

different conditions of time - dependent loading and of stress-path

bearing in mind, particularly, the conditions beneath a foundation

in the field. Accordingly the experimental programme was divided

into two distinct parts:

139

1. Oedometer tests

2. Triaxial tests.

8.6.1 Oedometer tests

The main series of oedometer tests was pefformed to study

the effect of (a) rest period and (b) the rate of loading on the

compressibility of London clay. It is well recognised that a con-

ventional oedometer test with 1-day load duration and a pressure in-

crement ratio of 1 does not reproduce the field condition where often

small load increments ( 21.t. 1) are slowly applied. Moreover, P

the inevitable structural disturbance that takes place due to the

release of stresses by sampling may have some effect on the com-

pressibility obtained from conventional tests. A number of samples

were allowed to rest for various periods of time at the in-situ

overburden pressure (2 T/ft2 for Ungar and 3 T/ft2 for Oxford Circus)

before any further loads were applied in order to determine whether

such sustained loading influenced the subsequent compressibility.

In addition, special oedometers were built for the follow-

ing investigations:

(1) Consolidation characteristics of the clay were studied in

the high pressure oedometer, in which specimens were loaded by

hydraulic pressure instead of dead weights, (11100 tests)* Settle-

ment, volume change and pore pressures were measured during con-

solidation.

140).

(ii) The effect of rate of loading on the compressibility was

studied by loading specimens at slow rates of strain, and preventing

any excess pore pressures to develop (CRS tests). The behaviour of

London clay at high pressures was also studied in the CRS-oedometer

by stressing samples to 7,000 lbs/in2

(iii) In order to define the stress path followed during one-

dimensional consolidation the lateral stresses were measured with

strain gauges fitted to special oedometer rings (0-SG test).

The complete schedule of eoedometer tests is given below:

(A) London clay from Oxford Circus

Standard oedometer tests

(a) Conventional tests : Daily load increment = 1

Initial swelling allowed :

Test nos: 0LOC-1, 2, 3 tSID

Conventional tests : Daily load increment - P

Initial swelling prevented :

0-00-4, 5

(c) Initial swelling prevented - 7 days' rest at 3.0 T/ft2 -

Then 3.0 - 6.0 T/ft2 as follows:

in 1 increment : 0-0C-6

in 5 daily increments, i.e. 0.6 T/ft2 per day : 0-0C-7

in 20 daily increments, i.e. 0.15 T/ft2 per day : 0-0C-8

Further loadings at daily increments and pressure increment

ratio of 1.

(b)

141.

(d) Initial swelling prevented - 90 days' rest at 3.0 T/ft2 -

Then 3.0 - 6.0 T/ft2 as follows:

in 1 increment : 0-0C-9

in 5 daily increments, i.e. 0.6 T/ft2 per day : 0-0C-10

in 6 daily increments, i.e. 0.15 T/ft2 per day : 0-0C-11

Further loadings at daily increments and pressure increment

ratio of 1.

(B) London clay from Ongar

(i) Standard oedometer tests

(a) Conventional test : Daily load increment - - p P 1

Initial swelling allowed :

Test no: 0-H0-1

(b) Initial swelling prevented - 1 day's rest at 2.0 T/ft2 -

Then 2.0 - 4.0 T/ft2 as follows :

in 1 increment : 0-H0-2 and 3

in 5 daily increments, i.e. 0.4 T/ft2 per day : 0-H0-1

in 20 daily increments, i.e. 0.1 T/ft2 per day : 0-H0-5

(c) Initial swelling prevented - 7 days' rest at 2.0 T/fta - Ap

Then daily load increments at ---- =

Test nos: 0-H0-6, 7, 8

(a) Initial swelling prevented - 90 days' rest at 2.0 T/ft2 -

Then 2.0 - 4.0 T/ft2as follows :

in 1 increment : 0-H0-9

in 5 daily increments, i.e. 0.4 T/ft2 per day : 0-H0-10

1

142

in 20 daily increments, i.e. 0.1 T/ft2 per day : 0-H0-11

(e) Consolidation from a slurry : 0-H0-15

(f) Special "swelling" tests (see chapter 10) : 0-H0-121 13, 14.

(ii) High pressure (hydraulic) oedometer tests

HPO-H0-1 and 2

(iii) Controlled rate of strain tests

CRS-H0-1, 2

(Two tests, CRS-W-1 and 2 were also performed on the London

clay from Wraysbury)

(iv) Tests in oedometers fitted with strain _gauges

0-SG-1, 2 and 3

8.6.2 Triaxial tests

The main objects of the triaxial tests were:

to determine the stress - strain and pore pressure (a)

behaviour of undisturbed London clay for the range of stresses

applied in the field and

(b) to study the consolidation characteristics of London Clay

for different stress paths.

It has been explained in chapters 5 and 7 that the stress-

path of a typical element of clay beneath a foundation consists of

undrained loading which causes the shear deformation, followed by

consolidation (see Fig. 5.5). The stress changes during the latter,

however, depend on the pore pressure parameter A and Poisson's

ratio J (see chapter 7) and may be quite different from that for

143.

one-dimensional consolidation in the oedometer. The best procedure

for a settlement analysis would, obviously, be to apply to a specimen

the same set of stresses that will be applied in the field and to

measure the corresponding deformations. This means that the

sample must first be brought back to the state of stress that pre-

vailed in the ground and then subjected to the stresses that will be

applied on loading in the field.

It has been shown in the previous section that the effective

stress in an undisturbed sample of London clay is greater than the

in-situ vertical effective stress. Therefore, after placing a

specimen (axis vertical) in the triaxial apparatus under an all-

round pressure the axial stress should be reduced without change in

water content until the effective stresses are what they were in the

ground. Then the stress path corresponding to that in the field

should be followed.

It was the purpose of the triaxial tests to study the

effect of stress paths, such as described above, on the deformation

Of London clay and also to determine the effect of isotropic and

anisotropic stress changes on the axial and volumetric strains

during consolidation.

The different types of triaxial tests carried out in the

investigation are described below. The graphical representation

of the stress path associated with each type is shown in Fig. 8.6.

144.

(a) Undrained tests to failure

Type Al. Unconsolidated undrained compression tests with measure-

ment of pore water pressure,

Test nos: T-0C-1, 2, 3

T-H0-1, 2, 3, 12

Type A2. Unconsolidated undrained extension tests with measurement

of pore water pressure,

T-H0-4, 5

Type A3. Axial stress on the specimen decreased until the effective

vertical stress was nearly equal to that in-situ, followed by com-

pression to failure - all under undrained conditions,

T-H0-6, 7

(b) Consolidation tests

Type Bl. Isotropic consolidation,

T-H0-8, 9, 21/1, 21/2, 24/1, 26/1, 26/2, 33/1

Type B2. Anisotropic consolidation,

T-H0-20/1, 20/2, 22/1, 22/2, 22/3

Type Cl. Undrained loading followed by isotropic consolidation,

T-OC-4, 5, 6

T-H0-10, 11, 13, 17, 19, 27, 28

Type C2. Undrained loading followed by anisotropic consolidation,

T-110-29, 30

Type D. Axial stress reduced until vertical effective stress was

nearly equal to the in-situ stress, then undrained loading, followed

145-

by isotropic consolidation,

T-HO-11F, 15, 16

Type El. Drained compression test with constant stress increment

ratios,

T-HO-31, 32, 33

Type F. Unconsolidated or consolidated drained compression tests,

Vertical samples : T-HO-24, 25, 26

Horizontal samples : T-H0-34, 35, 36

(Note: 1. 'OC' denotes Oxford Circus

'HO' denotes High Ungar

2. avi denotes stage 1 of test no. 21).

146.

Table 8.1

INDEX PROPERTIES OF LONDON CLAY FROM

OXFORD CIRCUS AND HIGH ONGAR, ESSEX

Site Block No.

Specific Gravity Gs

Liquid Limit L.L.%

Plastic Limit P.L.%

Plasticity Index P.I.%

Clay Fraction < 2/1.t. %

Activity P.I./Clay Fraction

oxford 1 2.67 61.0 26.5 34.5 48.0 0.72 Circus 2 2.69 64.0 27.0 37.2 48.0 0,78

Average 2.68 62.5 26.5 36.0 ;48.0 0.75

High 1 2.70 69.0 26.2 42.8 52.5 0.82 Ongar 2 2.71 66.3 27.4 38.9 48.5 0.80

3 2.72 67.2 26.7 40.5 42.5 0.95 Average 2.71 67.5 26.5 41.0 48.0 0.86

TABTN, 8.2

OEDOMETER TESTS

Sample Time from Initial Bulk Initial No. sampling water Density Degree of

to start content X

Saturation* of test (days) Wo% lbs/c.ft. %

Oxford Circus o-oc-1 7 22.9 126.2 97.5

2 8 22.9 126.1 98.o 3 45 22.8 124.6 97.5 4 45 22.7 126.5 98.o 5 102 23.6 126.5 99.5 6 124 23.2 126.1 98.o 7 91 24.5 125.8 100.0 8 91 23.6 125.9 98.o 9 91 24.6 125.3 99.o 10 13o 23.o 124.5 94.5 11 139 23.o 126.2 97.5 12 140 23.4 127.1 101.0

Average 23.3 125.9 98.0

Ongar 0-H0-1 36 27.7 125.2 104.5

2 36 27.7 127.1 107.5 3 36 27.9 122.4 100.0 4 86 26.6 118.4 94.o 5 86 26.8 120.4 95.o 6 36 27.9 118.3 93.0 7 113 26.9 122.4 98.o 8 113 27.4 121.2 96.5 9 113 27.5 118.0 93.0 10 149 26.5 121.1 96.5 11 149 27.1 124.1 101.5 12 593 27.o 123.3 98.o 13 6o5 27.3 121.7 96.5 14 610 26.9 122.6 97.o

Average 27.2 121.9 98.0

* Rounded off to nearest 0.5%

147.

148.

TABLE 8.3

TRIAXIAL TESTS

Sample No.

Time from sampling to start of test (days)

Initial Water Content

Wo%

Bulk Density

i lbs/cat.

Initial Degree of Saturation*

(%)

Pk

lbs/in2

Oxford Circus T-OC-1 - 23.0 - 100.0

2 111 24.6 128.5 105.0 104.0 3 114 23.8 128.0 102.5 109.0 4 51 23.o 127.2 99.5 70.2 5 81 24.3 127.2 102.0 66.5 6 133 22.1 128.3 95.2 118.0 7 136 22.0 128.2 99.5 107.0 8 312 22.4 128.4 100.0 100.8

Average 23.2 128.0 100.0 96.0

Ongar T-H0-1 39 28.6 124.7 103.5 43.8

2 131 27.8 124.7 102.5 63.5 3 173 26.8 128.5 107.5 52.0 4 173 26.3 125.3 101.5 45.o 5 210 27.1 125.8 103.5 50.5 6 185 25.3 126.7 101.5 57.2 7 222 25.8 124.0 93.5 53.2 8 117 27.3 121.4 95.5 58.4 9 96 26.7 122.4 96.o 56.0 10 150 25.9 125.9 101.0 64.2 11 15o 26.7 125.0 101.0 58.o 12 6 27.7 123.7 100.0 47.8 13 185 25.9 124.3 98.0 62.8 14 5o 25.o 125.2 98.o 52.0 15 6o 27.1 124.2 100.o 62.o 16 83 27.o 120.7 95.o 55.5 17 210 26.0 124.0 98.o 53.8 18 39 28.o 125.8 100.0 45.8 19 232 26.5 125.2 101.0 47.4 20 409 26.5 121.6 94.5 84.o 21 413 26.9 118.9 93.5 70.2 22 434 26.5 120.5 95.0 58.o

(Continued)

Sample Time from sampling to start of test (days)

TABLE 8.3

(continued)

Initial Bulk Water Density Content

" lbs/c.ft.

Initial Degree of Saturation*

(c/O lbs/in2

149.

Pk

23 555 26.o 122.6 95.5 56.o 24 439 27.5 122.1 97.0 64.o 25 461 25.6 124.1 97.o 80.0 26 507 25.6 124.0 99.5 70.0 27 512 26.5 121.0 92.5 49.o 28 529 27.1 121.7 96.0 53.0 29 512 26.4 122.0 95.0 57.o 3o 529 26.7 121.8 95.o 57.o 31 555 26.o 122.6 95.0 56.o 32 558 26.9 123.8 99.0 56.o 33 577 25.4 124.7 92.5 61.0 34 579 26.5 120.7 94.5 60.0 35 597 26.2 123.2 97.o 62.o 36 600 26.2 123.4 97.o 66.5

Average 26.5 123.5 98.0 58.5 (26.7**)

* Rounded off to the nearest 0.5% ** Taking oedometer test data (Table 8.2) into account

150. TABLE 8.4

STRESS CHANGES DURING SAMPLING

Site Depth Water Effective 0.C.R.* pk K- E Pk o op

Content Vertical Stress

(ft.) Wo% p(psi) (psi)

Oxford 23.3 46.6 96 2.06 2.77 129.0 Circus

High - a6.7 32.o 8.5 58.5 1.83 2.38 76.2 Ongar

Bradwell 10 36.0 4.8 44 8.75 1.82 2.17 10.42 (Skempton 15 34.3 6.75 32 14.45 2.15 2.64 17.82 1961) 20 33.0 8.75 25 19.8 2.26 2.80 24.5

30 32.0 12.70 17 2.22 2.74 34.8 40 31.1 16.8 1334.880 2.07 2.53 42.5 50 30.4 21.0 11 2.29 39.9

4 48.1

60 29.9 25.1 9 1.901 2.10 52.7

70 29.6 29.2 8 48.3 1.65 1.93 85.6 8o 29.3 33.3 7 51.7 1.55 1.79 59.6 90 29.o 37.6 6.5 55.o 1.47 1.67 62.8 100i 28.8 41.8 6 57.8 1.38 1.54 64.4 110 28.7 46.o 5.5 6o.5 1.32 1.46 67.2

Ashford A 30 22.4 17 36.0 46 2.7 3.4 58 Common B 50 25.8 26 24.0 54 2.1 2.6 68 (Bishop, C 66 24.8 34 18.5 65 1.9 2.3 78 Webb and D 91 22.8 45 14.0 76 1.7 2.0 90 Lewin E 114 24.2 56 11.5 100 1.8 2.1 118 1965) F 138 23.6 66 10.o 110 1.7 2.0 132

Reductithn of effective overburdenpressure by erosion, (a) Ongar : 235 lbs/in (ee chapter 10) (b) Bradwell : 208 lbs/in (Skepton 1961) (c) Ashford Common : 600 lbs/in (Bishop et al 1965)

*

151.

CHAPTER 9

EQUIPMENT AND PROCEDURES OF TESTING

9.1 Oedometer tests

9.1.1 Standard oedometer

The main series of one-dimensional consolidation tests,

were performed in standard 3" Bishop-type fiaped ring oedometers

(Fig. 9.1) which have been in use at Imperial College for more than

a decade. A specimen, 3" dia. x ,-.1" thick is held in a brass ring

between two porous stones and the pressure is applied to the upper

stone by placing dead weights on the hanger at the end of a lever-

arm. The settlement is measured relative to the base of the cell

- there is no lateral deformation - by a micrometer dial gauge.

The cutting edge of the ring itself was used to "push"

a specimen into the ring. A small block of the clay approximately

4" x 4" x 11." thick was cut out of the main sample with wire cutters.

The inside of the ring, which was greased with vaseline to reduce

side friction, was then placed on the block with its cutting edge

facing down and, using it as a guide, the clay was trimmed with a

sharp knife until it was just over 3" in diameter. The ring was

then gently and evenly pushed into the sample, the excess clay

being automatically trimmed by the cutting edge. Finally the

specimen was cut on both sides of the ring to its proper thick-

152.

ness.*

The specimen thus obtained was placed in the cell between

the lower and upper porous discs - this latter was fixed to the

loading plate - which were pre-saturated by soaking in water. Two

Whatman's No. 54 filter papers were placed between the sample and

the discs to prevent clogging of the clay into the porous stones.

The cell was then mounted into the loading frame and the lever-arm

balanced with the fine adjustment weight. The reading on the dial

gauge was noted.

A small pressure, usually 0.05 - 0.1 T/ft2 was applied and

the cell filled with water. Soon the sample started to swell. In

tests where this initial swelling was allowed, the sample was left

overnight to complete the process. The next and nominally each

subsequent day the pressure was raised in the sequence 0.1 -->0.5

-->1.0--2.0 (and to3..0T/ft2 for Oxford Circus only) allowing the

sample to consolidate under each load. The dial gauge was read

on the usual d t basis.

For the majority of the tests, however, the initial swell-

ing was prevented by successively adding small weights to allow no

Breaking up of the clay along fissures and joints during trimming and pushing of the ring was an ever recurring problem. Though every care was taken to keep this to a minimum and badly disintegrating samples were always discarded, it may not have been possible to eliminate completely the possibility of incomplete saturation in some cases.

153.

change in the dial gauge reading until equilibrium was reached. The

sample was left overnight under this "swelling" pressure (average

2 T/ft2 for Oxford Circus and 0.5 T/ft2 for Ongar). The following

day the Oxford Circus clay was raised straight from 2.0 --5.0 T/ft2

and the Ongar clay from 0.5---'fr1.0 p2.0 T/ft2 in two days:

The different periods of rest were then allowed according

to the Schedule given in Chapter 8. For long duration tests,

readings were taken daily during the first week, thereafter once

every week till the end of the rest period. After this the loads

were doubled every day until the full capacity of the machine was

reached. Cycles of unloading and loading then followed - all at

the usual pressure increment ratio of 1 or pressure reduction ratio

of i applied daily:* At the end of all tests, when the specimens

had completed the final swelling at zero pressure, they were re-

moved and their moisture contents were determined.

In test nos. 0-H0-12, 13 and 14 the "swelling" pressures

were obtained in the same way as described above, but instead of

loading the samples any further they were quickly removed from the

press. The water was first siphoned out of the cell, the pressure

released and the samples pushed out of the rings. Their water

contents were then measured.

*

pressures **

16 T/ft2.

3 T/ft2 and 2 T/ft2 being the in-situ effective overburden of the Oxford Circus and Ongar samples respectively.

First unloading in some tests was commenced at 12 or

154.

The slurry tests were performed on remoulded samples of

Ongar clay prepared from the air dried material mixed thoroughly

with freshly boiled distilled water to a consistency well above the

liquid limit.

9.1.2 High pressure (hydraulic) oedometer

(a) Description of the apparatus (Fig. 9.2)

The high pressure oedometer was designed to apply pressures

of up to 10,000 lbs/in2 on 4,1 dia. x 1" high soil.samples for one-

dimensional consolidation, though in the present investigation the

maximum pressure used was only 500 .

The oedometer essentially consists of a 6" I.D. x al"

thick x 1" deep mile steel annular ring sandwiched between a top

cap and a base, also made of mild steel. The sample is contained

in a stainless steel internal cell, 4" I.D. x 1" thick, which is

locked to the outer ring- by a circular key at the base. The cell

and the sample rest on a 6" dia. x 5" deep stainless steel "insert"

fitted into the mild steel base. There is a recess for a 4" dia.

xi" thick porous stone beneath the sample which communicates to

two drainage holes leading down to the bottom of the insert. The

volume change is measured by connecting one of these leads to a

5 c.c. reversible type volume gauge (Bishop and Donald 1961) in

conjunction with the self compensating mercury control to maintain

constant back pressure.

155.

The chamber inside the top cap serves to contain the

equipment and connections for measuring the pore water pressure and

deformation of the sample (see below). It is filled with water or

oil to transmit the pressure onto the sample through a membrane

which is sealed at the outer edge by pressing an 0-ring from a

stainless steel flange clamped to the top of the cell. The top

cap is held down to the base by 12 no. dia. holding-down bolts.

Pressures of up to 500 lbs/in2 have been applied through

a Klinger valve connected to the bottom of the top cap. Constant

pressures during consolidation were maintained with the extended

range self-compensating mercury control (Bishop and Henkel 1962).

(b) Equipment for measuring deformation

The settlement of the sample is measured over only a part

of its area. If, as is likely, the effect of friction between the

sample and the confining:xing is restricted to a zone near the

surface of contact, measurement over a small area at the centre

will be relatively free of this effect.

The settlement is given by the vertical displacement of

a 3" dia. stainless steel disc* which has a spigot at the centre.

The membrane with a hole in the middle is slid over from the top

and sealing is effected with an 0-ring pressed by a sleeve which is

clamped by screwing a nut to the spigot. Both the disc and the

Originally a 2" dia. disc was used to test the apparatus but later a 3" dia. disc was chosen to gain more room for the pore pressure connections.

156.

sleeve are sloped at the edge for smooth fitting of the membrane.

The settlement is measured by an inductive displacement

transducer type F.52 supplied by Boulton Paul Aircraft (Electronics)

Ltd. The armature of a transducer is screwed to the top of a

small column fitted at the centre of the spigot and is engaged into

the core which is held in a central position by means of a brass

"spider". The latter, in fact, is a three arm cross-head secured

at the ends on top of three studs screwed firmly into the insert.

At the centre of the "spider" is an adjustable transducer-holder to

which is screwed the core which can thus be held in a fixed position

relative to the movement of the disc and the armature.

The four connecting wires which are soldered to the core

of the transducer are led out of the top cap through a 5" long x

-" O.D. steel tube into which they are sealed by loctitet The tube

itself is fixed to the top cap by an i" Ermeto stud coupling.

The actual settlement readings are taken by connecting

the transducer leads through a Plessey Mk.IV 6-way plug to a B.P.A.

transducer meter type C.61. This is a 5-decade null balance in-

strument which enables the output of an A.C-energized bridge to be

measured. The meter has a normal bridge energising supply of 5v,

200 mA at 1 Kc/s and the decade switches, when adjusted for zero

meter deflectionm indicate the magnitude of the out of balance

All+•••-•••1ill

This is done by passing the leads through the tubing placed in a horizontal position and filling it with loctite.

157.

signal. At the high sensitivity range a minimum armature dis-

placement of 1 x 106 in. can be measured. For overnight record-

ing a speedometer H Automatic Recorder connected to the C.61

meter has been used.

The calibration of the transducer type F.52, serial no.

229, used in the present series of experiments was supplied by

Boulton Paul Aircraft Ltd. At a scale factor setting 100/0.0.0.

the meter reading directly gave the displacement in micro-inches.

(c) Measurement of pore water pressure

The dissipation of pore water pressure during consolida-

tion is measured at the top of the sample as shown in fig. 9.2.

The small porous stone, i" dia. x i" thick, recessed into the

bottom of the settlement - measuring disc communicates to an

0.D. copper tubing extending over the entire hole in the spigot

where it is sealed with loctite. A straight coupling connects

the other end of the double bend to an ill saran which is made into

a spiral inside the chamber to allow free movement of the disc

with settlement of the sample. The saran tubing is taken out of

the top cap through another hole and a drilled stud coupling pro-

vides the sealing. Outside the oedometer a standard null indicator

is connected to the saran through a Klinger valve.

(d) Procedure of testing

With the oedometer dissembled, the drainage connections

at the base were de-aired by flushing water and a porous stone,

158•

previously saturated with boiling water, was placed in position.

The measuring disc which had already been fitted with the copper

tube and a porous stone at its base was then assembled with the

membrane and connected to the saran while the top cap was sus-

pended from the lifting crane. Water was then flushed through

the tubing and the porous stone until all the air was driven out

of the system.

To place an undisturbed sample into the cell, a specimen

was first trimmed into a cutting ring, 4" I.D. x li" high, in ex-

actly the same way as described in the previous section for the 3"

standard oedometer, and then pushed into the cell with the aid of

two perspex guide discs as shown in Fig. 9.3. It was finally

made flush with the top and bottom of the cell.

The ring and the cell containing the sample were placed

on the base of the oedometer, with a saturated Whatman's no. 54

filter paper separating the soil from the porous stone, and screwed

in position with six Allen keys. The measuring disc was then

placed on top of the sample and the spider with the central trans-

ducer core, previously adjusted to give suitable meter readings to

allow for maximum settlement of the sample; was engaged to the arm-

ature of the transducer and finally secured on top of the studs.

This automatically ensured a central position for the measuring disc.

This was done by having a trial run with a steel "dummy" sample which was, in any case, necessary to check the electrical connections.

159.

The space between the membrane and the sample was then de-aired

by flushing a little water through the porous stone and spreading

it outwards. The flange was lowered in position and sealing was

effected by clamping it to the cell with six screws. Final ad-

justment of the meter reading was made by adjusting the transducer-

holder.

The top cap was slowly lowered with the help of three

guide pins bolted into the base of the oedometer. When it was in

position the guide pins were removed and the cap was secured to the

base by the holding down bolts. A complete layout of the apparatus

is shown in Fig. 9.4.

The chamber was then filled with water through connection

at the side, at the same time keeping a check on the transducer

reading so it did not show any undue fluctuations. After sealing

the bleeding valve a pressure of 15 lbs/in2 was applied keeping both

the drainage valve and the pore pressure connection closed. The

sample was left under this pressure overnight, during which a small

deformation occurred, possibly due to the solution of any air that

might still be present.

The following morning the pressure was raised to 30 lbs/in2

and a back pressure of 15 lbs/in2 applied. After the sample had

fully consolidated the effective stresses were increased, in steps,

to their maximum values and then reduced, according to the schedule

given below:

160.

Test No. Loading Schedule (lbs/in2)

HEO/H0/1 ) 15 - 30 - 60 - 120 - 240 - 360 - 485 - & )

HP0/H0/2 ) 360 - 240 - 120 - 60 - 30 - 15 - 0

(Back pressure during consolidation was always 15 lbs/in2)

The following procedure was adopted for each loading

stage:

When consolidation was complete under any effective

stress, indicated by the measured pore pressure of 15 lbs/in2, both

the drainage and pore pressure connections were closed and the cell

pressure was raised to the desired value (i.e. the pressure in-

dicated above plus 15 lbs/in2). Usually it took considerable

time for the pore pressure to reach its peak and at least 12 hours,

often more, had to be allowed for equalisation, before a response

of 90 - 100% could be achi:eved! Very little deformation occurred

during this operation.

Consolidation was begun by opening the drainage valve to

the volume gauge. During the early stages when consolidation pro-

gressed rapidly, readings of pore pressure and volume gauge were

taken on a t basis while the settlement was recorded automatically

on the speedometer H recorder. The latter was afterwards dis-

connected and the transducer read directly on the C.61 meter.

--------- * For a detailed discussion on the response time of the pore pressure measuring device, see Chapter 10.

161.

It took normally between 24 and 48 hours for full con-

solidation to take place. The samples were left at least until

the pore pressure .equalled the back pressure before the next in-

crement was applied.

At the end of each test when the specimen had finally

swelled under zero effective stress the chamber was emptied, the

top cap lifted, the water arm m rcury siphoned out, the flange

unscrewed and the measuring disc removed. The sample was then

pushed out of the ring and its final moisture content determined.

9.1.3 Controlled rate of strain oedometer

This oedometer has been used to study the compressibility

of London Clay at slow (strain-controlled) rates of loading and its

behaviour under very high effective stresses.

(a) Description

The oedometer (Fig. 9.5) is built almost entirely of

brass and is designed to withstand pressures of up to 10,000 lbs/i2

The sample, 3" dia. x i" high is confined in a stainless steel ring

1" thick* and rests on a brass base with an arrangement for measur-

ing pore water pressure. The ring is clamped to the base by means

of a brass flange and six studs, and sealed at both top and bottom

by 0-rings. The load is applied through a thick top cap at the

After the first two tests the inside of the ring was found to be slightly scored by the soil and was replaced by a thin brass inner ring inserted into the outer stainless steel.

162.

base of which is recessed a a" porous stone communicated through

drainage channels to the outside water, thus keeping the sample

submerged. Two dial gauges screwed diametrically opposite to an

annular plate sitting on the top cap measure the settlement. The

entire assembly sits on the pedestal of the 50 T loading machine

which can be selected to apply the load at a predetermined rate of

strain. The load on the sample is measured by a 25 T load cell

placed between the head of the machine and a steel spacer sitting

on the top cap.

(b) Calibration of load cell

The load cell was an N.C.B./M.R.E. Type supplied by

W. H. Mayes & Sons (Windsor) Ltd. A total of 16 electrical res-

istance strain gauges incorporated within the cell are connected

through a Plessey Mk.IV 6-way plug to a C.61 transducer meter

(described in section 9.1.2(b)) to measure the strain of a loaded

cylinder.

The load cell was calibrated in an Amsler 35 T Universal

testing machine. A number of loading and unloading cycles showed

negligible zero shift and little hysteresis. A sensibly linear

relationship between load and meter reading was obtained (Fig. 9.6)

(c) Procedure of testing

Two Klinger valves were connected to the outlets from

the base of the oedometer, to one of which was attached a standard

null indicator. De-aired water was flushed through the system and

163.

after closing the flushing valve, a saturated porous stone, li"

dia. x thick, was placed in the recess provided in the base.

The method of preparing a sample was similar to that for

the High Pressure oedometer described above (see Fig. 9.3). A

specimen was first trimmed into a cutting ring (3" I.D) and then

pushed into the testing ring with the help of perspex guides. Two

saturated Whatman's No. 54 filter papers were placed on top and

bottom of the sample which was then placed in position on the base

and clamped with the flange and studs. The centering of the

sample was thus automatically achieved. The top cap containing

the saturated drainage stone was lowered onto the sample and the

annular plate with the dial gauges was positioned from the top.

The spacerand the load cell were located and the pedestal of the

machine was raised to bring the assembly into contact with the head

of the machine. After adjusting the gear boxes to give the des-

ired rate of loading the test was started. Soon after, water was

poured to submerge the sample. A photograph of a test in progress

is shown in Fig. 9.7.

The rate of strain was so chosen as to produce negligible

pore pressure at the base of the sample which was measured through-

out each test to ensure that this was in fact the case. However,

this had to be achieved by trial and error. At the start of the

first test (CRO/H0/1) a rate of movement of the pedestal of

0.00004 in/min was selected, giving a rate of strain of .0045% per

164.

minute. A maximum pore pressure of 3% of the total vertical

stress was recorded after two days as a result of which the rate

was successively reduced to 0.000025 and .00001 in. per min. giving

a final rate of strain of approximately 0.0012% per min. at which

the maximum pore pressure was 1.5% of the vertical stress. For

subsequent tests a constant pedestal velocity of 0.00002 in/min

was found satisfactory, giving a rate of strain of approximately

0.0025% min. At this rate the pore pressure at the base never

exceeded 2% of the vertical stress* which was accepted as the upper

limit.

In all, 4 tests were performed on the London Clay, two

from Ongar and two from Wraysbury!* Except for the second Ongar

test (CRA/H0/2) all the others were loaded to maximum stresses of

about 450 T/ft2 (7,000 lbs/in2) followed by unloading at the same

rate. Test No. CRA/H0/2 had two cycles of loading and unloading,

the first one to a pressure of 160 T/ft2 (3,000 lbs/in2) and the

, second one to the maximum of 450 T/ft2 (7,000 lbs/in2 ).

9.1.4 Oedometers fitted with strain gauges

In order to determine the complete stress-paths for

standard oedometer consolidation special tests were performed on

Over most of the * * A detailed study bury clay has been made by

range, however, it was considerably less.

of the undrained strength of the Wrays-Agarwal (1967).

165.

the Ongar clay in oedometer rings (3" dia. x 1" high) with strain

gauges fitted on the outside. Three different ring thicknesses

1 1' and $") were adopted ( d Tr) and one test was performed with each. '32llt 16

(a) Description

The rings were all made of brass and cut out of 3" I.D.

tubes. The thicknesses were uniform and there was not cutting

edge.

The strain gauges used were of the type PLS -10 supplied

by Tokyo Sokki Kenkyujo Co. Ltd. Each gauge consisted of a grid of

resistance wire supported on a base impregnated with polyester resin

and was of nominal gauge length 10 mm,-nominal gauge width 1.5 mm,

nominal resistance 120 - 0.3 gauge factor 2.07 and base

dimension 25 x 4 mm. The gauge factor was checked by Phukan (1968)

and was found to be sensibly constant at the manufacturer's value

of 2.07.

Each ring was mounted with four transverse gauges, as

shown in Fig. 9.8. A full bridge circuit was completed with four

auxiliary gauges mounted on a "dummy" ring to compensate for

variations of temperature. The four leads from the circuit were

connected to a Plessey Mk. IV 6-way plug and readings were taken on

a transducer meter Type C.61 described in section 9.1.2(b).

The mounting of the gauges needed particular care. The

location for a gauge on the ring was first polished with emery

paper and cleaned with acetone. A thin coat of a mixture of P-2

166.

and PS adhesives was applied to the face of the gauge and to the

ring. The gauge was then placed and covered with sheet teflon and

a uniform clamping pressure was applied by suspending a small weight

at the end of a ring of polythene sheet covering the gauge. The

adhesive was allowed to cure for at least 2 hours. Waterproofing

was achieved by applying Gagekote 4%5, a two-component rubber line

epoxy resin, supplied by Welwyn Electric Ltd., on the gauges and

their connections. As an extra precaution, when all the connections

were made, the entire ring was covered with "Glasticord" waterproof-

ing tape, supplied by Kelseal Ltd.

To check the effectiveness of the waterproofing both the

test and the compensating rings were immersed in water and readings

were taken over a period of 3 - 1+ days. Practically no fluctuations

were noticed.

(b) Calibration of oedometers

Each oedometer ring, connected in full bridge circuit, was

calibrated with water pressure as shown in Fig. 9,9. Essentially,

the ring was sealed at both ends by 0-rings fitted on two end plates

which were bolted together. The pressure was applied from a

Budenberg dead-weight calibrator, in steps, to a maximum of 500

lbs/in2 and the corresponding readings were taken on the meter. A

number of loading and unloading cycles were performed. Each time

the difference between the initial and final zero readings was small

and there was little hysteresis. The average calibration data for

167.

all the rings are shown in Figs. 9.10, 9.12 and 9.13. It can be

seen that linear relationships existed between pressure and the

meter readings.

At the end of each test the ring was re-calibrated. As

the charts show there was no significant difference between the

before- and after-test calibrations for any ring,

(c) Procedure of testing

At the beginning of each test a reading on the meter was

taken with both the active and compensating rings under water and

this was taken as the zero reading. They were then temporarily

removed to place the specimen into the test ring.

The sample was prepared in exactly the same way as des-

cribed for the controlled rate of strain tests. The specimen was

mounted on the standard 3" Bishop type oedometer press and a small

load was applied. Both the test ring and the compensating ring

were then submerged in water, the latter in a separate bowl, and

the sample was prevented from swelling by successively adding small

weights. When equilibrium was reached the meter was recorded and

the sample was left overnight, during which little change occurred.

Subsequently the procedure for a standard oedometer test was

followed, doubling the pressure every day as described in section

9.1.1. Both the settlement and meter readings were taken on the

JT basis. Records of temperature were kept throughout each test

which showed a normal variation between 18°C and 21°C and there was

168.

usually i.°C difference between the test and the compensating rings.

The final zero reading at the end of each test, when the sample was

taken out of the ring, was compared with the initial readings and

the difference was found to be small.

9.2 Triaxial tests

(a) General

All the triaxial tests were performed in a new batch of

1.1" standard triaxial cells, described by Bishop and Henkel (1962).

The cells were all fitted with rotating bushings to minimise ram

friction and there were arrangements in the base for outlet of the

leads from drainage connection to the top of samples. Constant

cell pressure was maintained with self-compensating mercury control

(range 0 - 150 lbs/in) and the axial stress was applied at controlled

rate of strain by a loading system consisting of a screw jack

operated by an electric motor and gear box which could be adjusted

to give a wide range of strain rates. The load was measured by

7" O.D. high tensile steel proving rings (capacity 300 lbs) which

were calibrated from time to.time with the Budenburg dead-weight

calibrator.

(b) Measurement of pore water pressures

The pore pressures of all samples were measured at the

base by the standard Perspex null indicators (Bishop and Henkel,

(1962)). When properly de-aired these null indicators have a low

169.

volume factor (1.0 x 10-6 cu.in per lb/in

2 ). The null balance of

the water mercury surface was observed with a low - power cathe-

tometer.

(c) Measurement of volume change

One of the features of the triaxial testing on undisturbed

specimens of Lohdon clay was the measurement of small volume changes

during consolidation or drained compression. Often total volume

changes of the order of 1 c.c. had to be measured over a number of

days. The correct determination of the consolidation properties

from time vs. volumetric strain characteristics, therefore, called

for the accurate measurement of the volume of water expelled by a

specimen. The standard 5 c.c. volume gauges used for the early

tests on the Oxford Circus clay were not found satisfactory and the

following method was used for the Ongar tests.

In essence, the volume gauge was replaced by an in O.D.

Saran tubing connected between the sample and the back pressure

control and clamped horizontally on the bench against a yard scale

(Fig. 9.13). A length of air bubble was introduced in the other-

wise water-filled tubing and its movement gave a measure of the .

amount of water coming out of a specimen. The tubing was calibrated

in two ways:

(i) connecting the saran to a 5 c.c. volume gauge and measur-

ing the displacements of the air bubble for known changes of volume

and

170.

(ii) by measuring the weight of a known length of water or

mercury expelled out of the tubing.

The results of the calibrations are given below:

Tube No.

Calibration Factor : in/c.c.

By volume gauge

By weight of water expelled

By weight of mercury expelled

Average

1 (M/C 1)

22.67 22.60 22.53 22.60

2 (M/C 2)

21.6 21.54 21.59 21.58

It will be found that all three methods of calibration gave very

consistent results although there was a slight difference between

the two batches of saran used for the two machines. Moreover this

calibration was found to be sensibly constant over the range 0 - 150

lbs/in2. This was checked by enclosing an approximately 15" length

of mercury into the tubing and increasing the pressure, in steps,

from 0 to 150 lbs/in2 with the end sealed - at which almost negligible

change in length of the mercury occurred. (From 18.90 in to 18.80

in for Tube 1 and 13.42 in to 13.40 in for Tube 2).

The effett of the pressure difference between inside and

outside of the saran - the outside pressure was atmospheric - was

171.

considered insignificant because of the above observation. Also,

the back pressure being typically as low as between 15 and 30

lbs/in2 the pressure gradient was never too great to cause any

significant loss of water by diffusion.

By adopting the system of measuring volume change as

described above a high degree of accuracy - a displacement of 0.02

in of the air bubble was equivalent to volume change of only 1 mm3

- was achieved.

(d) Application of extension load

The triaxial tests of Types A2, A3 and D necessitated

the application of upward forces to the top caps. A loading frame,

consisting of two beams and tie rods was used to pill the ram, the

load being measured by the compression proving ring placed on top

of the testing machine as shown in Fig. 9.14. The ram was passed

through a hole at the centre of the lower beam, the upward force

being applied to the dial gauge holder pinned to the top of the

ram.

Inside the cell, a special stainless steel rod with a

horizontal pin was screwed to the other end of the ram. The top

cap had a vertical slot to receive the pin and two diametrically

opposite horizontal grooves into which the pin was engaged by

rotating the ram through 900 before an extension test was due to

start.

172.

(e) Drainage connections

In the early consolidation tests the specimens were

drained from the base only but later, to minimise the duration of

a test, drainage from both ends was effected. The top drainage

lead, sealed to the loading cap by loctite, was made into a spiral

round the sample. Outside the cell this was connected, together

with the base drainage lead, to the saran volume gauge through two

Klinger valves and a T-joint, thus allowing both the top and bottom

of the sample to be maintained at the same back pressure.

(f) Leakage

The average duration of a test being of the order of 3

weeks, it was necessary to ensure that leakage of water into the

sample through the rubber membrane would not cause any appreciable

change in effective stresses. Also if there was excessive leakage

through valves and connections measurements of volume change would

be in error. Poulos (1964) made an extensive investigation into

the problems -of leakage in triaxial tests and concluded that for

long duration tests special precautions might be necessary.

Leakage of cell water through rubber membranes may occur

due to

(i) the hydraulic pressure gradient between two sides of a

membrane and

(ii) the osmotic pressure difference caused by the different

salt concentrations of the water in the cell and in the sample.

173.

No direct investigation was carried out to determine the

extent of leakage due to each of the above causes. However, in

order to estimate the amount of leakage that would cause appreciable

changes in effective stresses the initial swelling ratio of the

Ongar clay, as defined by Poulos (1964), was determined, (For

details see Appendix D). The results showed that for a typical

test lasting 30 days, in order that the change of effective stresses

was not to exceed 2%, the maximum permissible rate of leakage would

be 5.3 mm3/day while a rough estimate of leakage through a 0.016"

thick Membrane .(used in the present series of tests), based on

Poulos' results, yielded a rate of only 1.67 mm3/day (see Appendix

D). So a membrane which was properly checked for holes and

punctures prior to a test would be capable of preventing excessive

leakage.

Another possible source of error in the measurement of

) volume change was leakage through connections and creep of the

saran tubing. To determine the magnitude of this error a special

test was run which showed that the maximum error from these sources

alone would be no more than 2% of the measured volume change (see

Appendix D). Since it was not possible to eliminate entirely these

leakages an error of 2% was considered an acceptable upper limit.

The measured volume changes of all the consolidation and

drained triaxial tests were checked with the changes in weight of

the samples determined at the end of each test. The results are

174.

summarised in 74. 9.17. It can be seen that the average ratio of

the two measurements in 0.96 which can be considered satisfactory.

(g) Procedure of testing

All specimens were prepared in the N.G.I. type soil lathe.

A prism, approximately 2" x 2" x 4" in size was cut out of a main

block and clamped in the lathe between two end plattens which could

be rotated while a series of thin slices were trimmed with a fine

wire saw, the two walls of the frame acting as guides to give the

sample its final diameter of 14". Initial moisture contents were

determined on pairings. The specimen was then reduced to its

nominal length by holding it in a q" I.D. x 3" long split mould

and trimming the ends. The weight and final dimensions of the

specimen were recorded.

Prior to preparation of a sample, all connections to the

triaxial cell were thoroughly de-aired. Flushing air-free water

was usually enough and when properly de-aired a null indicator would

show only a slight movement under a pressure of 150 lbs/in2. All

drainage connections were also de-aired and an air bubble introduced

into the saran volume gauge!

A little water was allowed to flow to the top of the

pedestal by driving the air bubble to near the drainage valve thus

giving it the maximum travel. A saturated porous disc was slid on

In fact, a bubble once introduced into the saran could be used for a'number of successive tests.

175.

the top of the pedestal and covered by a Whatman's no. 54 filter

paper. The bottom of the sample was smeared with a drop of water

and placed on the porous disc. Another filter paper and a porous

stone were placed on top of the sample, which had also been smeared

with a drop of water. A rubber membrane, previously checked for

holes and punctures, was placed to cover the sample, and the top

loading cap brought into position: The membrane was sealed to the

pedestal with 2 or 3 0-rings. To remove air from between the

sample and the membrane, the latter was gently stroked upwards, and

finally sealed to the top cap with 2 or 3 more 0-rings.

The cell was assembled and filled with de-aired water with

about 1" of castor oil on top. The proving ring was placed in

position. A cell pressure higher than the initial suction was then

applied (60 or 70 lbs/in2 for Ongar) keeping ali-drainage valves

closed.

Initially a fairly high pore pressure (20 - 30 lbs/in2)

was measured (because the base of the specimen was already smeared)

but it soon began to drop until equilibrium was reached. This

usually took about 24 hours but often 48 hours or more were allowed.

The effective stresses of the samples (cell pressure minus the

equilibrium pore pressure) at the end of this stage were taken as

the post-sampling effective stresses pk (Chapter 8).

A3 and D. Special extension caps were used for test types A2,

176.

The cell pressures were then increased in small steps to

the desired value (each step lasting one or two days) allowing the

pore pressure to come to equilibrium under each increment.

values of almost 1.0 were usually measured which indicated that the

samples were fully saturated. The build-up of pore pressures with

increase of cell pressure in test nos. qTAH10-18.J and 9 are shown in

Figs. 9.17 and 9.18.

The following procedures were then followed for the

different types of tests.

Type Al

The specimen was loaded axially, undrained, until failure.

For accurate measurement of pore pressure during shear very slow rates

of strain (approximately 0.0004% per minute) were chosen (Bishop

and Henkel 1962) such that failure was reached in 8 - 10 days.

Pore pressures were measured throughout each test.

Type A2

Axial extension was applied, undrained, until failure was

reached. The loading frame was assembled as shown in Fig. 9.14.

The base of the cell was clamped to the screw jack which was then

set to travel downward at a rate of strain of about 0.0004% per

minute, giving a time to failure of about 8 - 10 days. Pore

pressures were measured throughout, about 3 or 4 times a day.

Zree_i_12

Early stages of these tests were similar to the A2 tests

177.

except that the extension was stopped at some desired effective

stress ratio (= c 0.55) and reloading then followed. This (711

was done simply by reversing the direction of travel of the screw

jack. When isotropic stress conditions were re-established, the

extension frame was dismantled leaving the lower beam on top of the

cell and the proving ring was brought back to its usual position.

Axial compression was then applied until failure, as in the Al tests.

Pore pressure was measured throughout.

Type B1

The consolidation tests of the types B1 and B2 did not

involve any axial compression. Once the pre-determined cell

pressure was reached and the equilibrium pore pressure measured the

samples were consolidated against a back pressure. Before opening

the drainage valve, however, the air bubble in the saran volume

gauge was allowed to come to equilibrium under the back pressure

for at least 30 minutes. The top cap was brought into contact

with the ram and drainage was started by opening the drainage valves.

The displacement of the air bubble was measured by the attached yard

scale while axial deformation was measured by the dial gauge. As

the sample consolidated the loading platform was raised by turning

the handwheel to maintain a constant reading on the proving ring

- at a value slightly higher then the initial reading in order to

ensure that the ram was always in contact with the sample. By the

evening, the rate of consolidation was sufficiently slow, and a smsll

178.

extra axial load was applied by raising the screw jack to ensure

that with overnight deformation the sample would remain in contact

with the ram. During the following days only occasional adjust-

ment was necessary.

Type B2

Exactly the same procedure as the B1 tests was followed

except that the cell pressure was also reduced at a steady rate to

simulate conditions in the field where the lateral pressure during

consolidation decreases due to the drop in Poisson's ratio. The

steady decrease of cell pressure was achieved by continuously

lowering the upper mercuty pot of the self-compensating mercury

control with the help of a motor and gear arrangement which could

be selected to give various rates of decrease. For a typical test

lasting 4 - 5 days, the cell pressure was decreased during the

first two days only - at a faster rate during the first few hours

of consolidation followed by a slower rate - and held constant

afterwardst To compensate for the loss of axial stress due to

the reduction of cell pressure the screw jack was regularly ad-

justed and the corresponding axial load was applied through the

proving ring.

Type C1

Both the cell pressure and the axial stress were in-

* To reproduce field conditions properly the rate of de-crease should strictly follow the average rate of consolidation.

179.

creased under undrained conditions, the latter at a rate of strain

of 0.0004% per minute, as in the Al tests and pore water pressures

were measured. When the desired effective stress ratio was

reached, loading was stopped and the sample was consolidated against

a back pressure (usually between 15 - 25 lbs/in2). Both the axial

and volumetric strains were measured, the constant axial stress

being maintained in the same way as described for the B1 tests.

Type C2

The undrained loading stage was similar to the Cl tests

while during consolidation the lateral stress was decreased as

described for the B2 tests. Both the axial and volumetric strains

were measured during consolidation, which took place against a

back pressure.

Type D

The initial undrained extension and compression stages

were followed in the same way as described for the A3 tests except

that the loading was stopped at some pre-determined stress-ratio,

before failure was reached. The specimen was then consolidated

against a back pressure, as for the Cl tests, during which both

axial and volumetric strains were measured.

Type E

After a sample was set up with all drainage valves closed,

the cell pressures applied and the equilibrium pore pressures

measured, both the cell pressure and axial stresses were increased

180.

under fully drained conditions at very slow rates such that no ex-

cess pore pressures developed. The rate of increase of the cell

pressure was so chosen as to give a constant stress increment ratio.

Test no. T-H0-31 had three stages with three different stress in-

crement ratios before, in the fourth and final stage, it was loaded

to failure as a conventional drained test. Test no. T-110-32 had

only one stage with constant stress increment ratio before it was

sheared to failure by axial compression. Test no. T-H0-33 was

similar to T-HO-32 except that the shearing stages were preceded

by isotropic consolidation against a back pressure.

Type F

These were conventional unconsolidated or consolidated

drained compression tests against back pressures. The rates'of

axial strain were so chosen (0.0004% per minute) as to produce

negligible excess pore water pressuring during shearing. A

typical sample failed in 8 - 10 days.

At the end of each test the specimen was removed as

quickly as possible from the triaxial cell and its final weight

measured. The difference between the initial and final weights

was then compared with the measured volume change. It can be seen

from Fie4.907that the agreement was, on the whole, good with the

average ratio of the two measurements equal to 0.96.

181.

CHAPTER 10

RESULTS OF OEDOMETER TESTS

10.1 Tests in the standard oedometer

10.1.1 Determination of swelling -pressure with standard oedometers

The swelling pressures: of the London clay from Oxford Circus

and Ongar, as determined from the initial stages of the standard

oedometer tests, were approximately 2.0 T/ft2 and 0.5 Tifft2

respectively. It was, however, difficult to maintain the dial

gauge readings absolutely constant during the determination of the

swelling pressures and deformations of between - 0.0002 in and

- 0.0014 in were observed. To get a more accurate measure of the

swelling pressure, therefore, the dial readings have been plotted

against vertical stresses in the vicinity of the swelling pressure

for all tests including those where swelling was permitted (curves

1 in Figs. 10.1 and 10.2) from which the swelling pressures corres-

ponding to zero change of height have been interpolated. The values

of swelling pressure thus obtained•for Oxford Circus and Ongar are

1.7 T/ft2 (26.5 p.s.i.) and 0.45 T/ft2 (74.s.i.) respectively. In

table 10.1 these values are compared with the initial suction

measured in triaxial tests and reported in chapter 8. It can be

seen that the swelling pressures measured in the standard oedometer

are too low, being only —288(.12 of the corresponding suctions

measured in the triaxial tests.

182.

In order to find an explanation for such discrepancies

three special tests (Nos. T-HO -12, 13 and 14) were performed where

the samples were removed from the oedometer press at the and of the

"swelling" tests and their height, weight and water contents deter,-

mined. These results are summarised in Table 10.2. It can be

seen that even though the dial gauges showed an average settlement

of 0.0005 in.the specimens had, in fact, swelled 0.0180 in and

gained 2.28 gms. in weight resulting in an increase in water content

from 27.1% to 29.4%. So, far from remaining at constant volume the

samples actually sucked in water thus giving considerably smaller

values of the swelling pressure.

Possible reasons for this could be:

(a) degree of saturation of the specimens less than 100%

(b) deformation of the loading platform and

(c) bedding error at points of contact between the sample and

the porous stones and between different parts of the apparatus itself.

It is true that the average degree of saturation of the

specimens before test was 97.5% which rose to almost 100% after the

tests were complete. This would account for an increase in weight

of less than 1 gm and no change in height. So the actual increase

in weight of 2.28 gm and considerable increase in thickness must be

attributed to causes (b) and (c) above.

Toms (1966) observed from loading tests with steel

"samples" that the oedometers used in the Chief Engineer's laboratory

183.

of the Southern Railways showed considerable elastic and inelastic

deformation. In order to determine the magnitude of such deforma-

tion of the oedometers used in the present investigation, these were

calibrated with steel "dummy" samples set up in exactly the same way

as in actual tests. The results are shown in Fig. 10.3.

Large inelastic deformations and hysteresis are observed

for all three frames. The measured deformations consist of the

elastic deflection of the loading platform which supports the con-

solidation cell and is held by two nuts screwed to the tie rods at

the end and bedding error at points of contact between different

parts of the apparatus. This latter error which is not wholly

recoverable is, perhaps, the cause of the hysteresis.

An approximate calculation* showed that the elastic de-

flection of the loading plate alone could be 0.0030 in.for the

maximum pressure of 32 T/ft2 i.e. 0.0001 in.per T/ft2. But the

total deformation of each frame was 0.0130 in. indicating that the

remaining 0.0100 in must have been due to inadequate bedding at the

various contant points, e.g. between the screw jack and the re-

action, between the plunger and the nose of the loading cap, and

between the "sample" and the porous stones.

However, even applying all these corrections for the

corresponding stresses, from Fig. 10.3, the swelling pressures as

Assuming that the plate behaved as an elastic beam, simply supported at its ends.

184.

determined from curves 2 in Figs. 10.1 and 10.2 are 2.4 T/ft2 and

0.5 T/ft2 for Oxford Circus and Ongar respectively which are still

considerably lower than the triaxial measurements. These dis-

crepancies which are due to increase in height of the samples un-

accounted for by the deformation of the apparatus must be the con-

sequence of inadequate contact between the sample and the porous

stones and perhaps between the sample and the confining ring.

This increase in height of a sample during the determina-

tion of the swelling pressure has a marked influence on the later

consolidation characteristics as will be subsequently demonstrated.

10.1.2 Time - settlement relationships

The time - settlement diagrams for some of the consolida-

tion stages with pressure increment ratio of 1 are given in Appendix

E. It will be seen that the initial deformations just after load

application are quite large. For each curve three values of zero

reading are shown, obtained as follows:

(a) Initial reading on the dial gauge before pressure incre-

ment

(b) Zero reading corrected only for apparatus deformation -

as from Fig. 10.3, and

(c) Zero reading obtained by Casagrande construction (Taylor

1948), from the shape of the early portion of the curve.

It is obvious that for most curves the corrected readings

185.

(b) and (c) do not coincide. This indicates that deformation of

the apparatus alone cannot account for the large initial deforma-

tion and that inaccurate bedding and, perhaps, the effect of the

shock loading may be important. The amount of discrepancy

between the two readings may be judged from Figs. 10.4 and 10.5 in

which the ratio of the settlements calculated on the basis of the

two zero readings (b) and (c), (hereafter referred to as Method 1

and Method 2 respectively), are plotted against pressure. It is

clear that on first loading the difference between the two methods

is considerable at lower pressure (40% at 2T/ft) but decreases at

higher pressures (10% at 32 T/ft2). On reloading, however, method

1 is, on average, only 5% greater than method 2 over the entire

stress range. This shows that inadequate bedding, is primarily

responsible for the initial deformation. At higher pressures on

first loading and during the second loading when the bedding error

has largely been eliminated the two methods give almost the same

settlement.

10.1.3 Pressure - void ratio relationships

The pressure - void ratio relationships for all the

standard oedometer tests have been calculated from settlements

obtained by methods 1 and 2 described above, but for reasons stated

in the preceding section, method 1 is not believed to give the true

consolidation settlement and, therefore, the e vs log p diagrams

186.

obtained by method 2 only are shown in Figs. 10.6 - 10.14. For all

tests, the void ratios have been calculated, working backwards, from

the final void ratio determined at the end of the tests. For com-

parison, the initial void ratio of each specimen calculated from

the initial moisture content is also shown with an open circle.

Oxford Circus

Fig. 10.6 shows the e vs log p diagrams for test nos.

0-0C-1, 2 and 3 in which the specimens were allowed to swell under

pressures of 0.4, 0.25 and 1.0 T/ft2 respectively before further

load increments were applied. It will be noticed that the magnitude

of the pressure increment has no significant effect on the slope of

the curves.

In Fig. 10.7 are plotted the pressure - void ratio re-

lationships for test nos. 0-00-4 and 5. These were conventional

oedometer tests in which initial swelling was apparently prevented.

A comparison between the actual void ratio determined from the

initial water content and that calculated from the final measurements

shows that considerable swelling had, in fact, taken place.

The results of the tests in which the samples were nilowed

to rest for 7 days (0-00-6, 7, 8) at the in-situ effective over-

burden pressure of 3.0 T/ft2 are shown in Fig. 10.8. It may be

noticed that the magnitude of the pressure increments subsequent to

this rest period does not significantly affect the e vs log p

diagrams.

187.

In contrast, longer rest periods of 90 days (0-0C-97 10,

11 : Fig. 10.9) show some influence on the shape of the pressure

void ratio diagram in the subsequent range 3 - 6 T/ft2. The

specimens have all undergone some secondary consolidation (see

enlarged detail in Fig. 10.15b) during the long rest periods and

have developed marked resistance to further compression, as can be

seen from the very flat shape of the curves for small pressure in-

crements.

Ongar

Fig. 10.10 shows the e vs log p relationship for test

no. 0-H0-1 where the specimen was allowed to swell under a small

pressure of 0.1 T/ft2 before it was loaded conventionally and in

Fig. 10.11 are plotted the pressure - void ratio relationships for

conventional tests (0-H0-2 and 3) where swelling was apparently pre-

vented but had in fact taken place. The effect 02 rate of loading

after the conventional rest period of 1 day at 2.0 T/ft2 (0-H0-4

and 5) is shown in Fig. 10.12.

The results of the three tests (0-H0-6, 7 and 8) in which

the specimens were allowed to rest 7 days at the in-situ vertical

effective stress of 2.0 T/ft2 but otherwise loaded in the conven.1,.

tional manner are plotted in Fig. 10.13.

Finally, Fig. 10.14 shows the e vs log p diagrams for

test nos. 0-H0-9, 10 and 11 in which rest periods of 93 days were

allowed at 2.0 T/ft2 before the samples were subjected to further

188.

load increments - at different rates. Again, the main effect of

the prolonged rest period is found to,ghe greatly increased res-

istance to compression for subsequent small pressure increments

(see Fig. 10.15a).

The pressure - void ratio relationship of the Ongar sample

no. H0-15 which was consolidated from a slurry (initial moisture

content 89.6%) is shown in Fig. 10.16. Here the correlation

between the initial void ratio and that calculated backwards from

final measurements is quite satisfactory which indicates that if

settlement is large the effect of inadequate bedding and the

apparatus deformation is not significant.

10.1.4 Compressibility characteristics

The coefficient of volume compressibility (mv) at a

pressure po is defined as

m = v A

Lip 1 eo

where eo is the void ratio corresponding to the pressure po and

e is the change of void ratio due to a pressure increase from

po to (po +.6 p).

In Figs. 10.17 - 10.20 are plotted the relationship

between compressibility and effective stress for the normal pressure

increment ratio of 1 (i.e. --211 = on both first and second

Po

A e 1

189.

loadings. For each case the range of scatter is indicated by the

shaded areas.

In the case of Oxford Circus, samples 0-0C-1, 2 and 3

which were allowed to swell initially under small pressures, ex-

hibited consistently higher compressibility than the others over the

entire stress range (Fig. 10.19). Only one corresponding test was

performed on the Ongar clay (0-H0-1) the results of which fell within

the range of scatter for the other tests. As is usually found with

clays, the compressibility on reloading is considerably lower than

on first loading particularly in the low stress range. In fact,

in the vicinity of the pressure at which reloading is started the

compressibility is initially very low as indicated by the very flat

slope of the e vs log p diagrams. It then increases slightly

before starting to decrease with increasing pressure.

The magnitude of pressure increment after rest periods'

of 1 day and 7 days at 2.0 T/ft2 (Ongar) and 3.0 T/ft2 (Oxford

Circus) did not have any significant influence on the compressibility.

The results fall within the range of scatter shown in Figs. 10.17,

10.18 and 10.20. On the other hand the rest period of 90 days and

the magnitude of subsequent loading have very marked effect on the

compressibility as shown in Figs. 10.21 and 10.22. It has already

been shown in Fig. 10.15 that these samples had undergone some

secondary consolidation during this prolonged rest period and at the

same time developed marked resistance to further deformation. As a

190.

result, the compressibility is very low for small pressure changes

but increases as —EA increases. Yet a pressure increment ratio

of 1 (test nos. 0-H0-9 and 0-0C-9), after these rest periods, pro-

duces a compressibility which is very close to the average for the

standard tests. This shows that a large increment of pressure

completely destroys the extra resistance the clay may develop after

a period of sustained loading, which can, therefore, only be de-

tected by small pressure increments.

10.1.5 Coefficient of consolidation

The coefficient of consolidation (cv) for each loading

stage has been calculated for t50, obtained in the usual way from

the settlement vs log time plot (Appendix E). In Figs. 10.23 and

10.24 are shown the variation of cif with effective stress, cv

being plotted against the average pressure for the corresponding

consolidation stage. In general, cv decreases with pressure on

first loading though the scatter is usually large. On second

loading, however, cv is considerably lower than on the first and

increases slightly with pressure until the stress from which un-

loading started is reached, beyond which it again decreases mith

pressure.

That the lower values of cv on second loading are pre-

marily due to lower void ratios are demonstrated in Fig. 10.25

where the coefficient of consolidation is plotted against void •

191.

ratio. The results for both first and second loading lie within

the same range and approximately unique relationships between cv

and e are obtained. It can also be seen that at smaller void

ratios cv remains virtually constant indicating that both the

permeability (k) and the compressibility (mv) decrease in such

a way that their ratio remains constant. This is shown in Figs.

10.26 and 10.27 where my and k* are plotted against void ratio.

For e less than 0.65 both log k and log my increase linearly

with e and at the same rate resulting in a constant value of cv.

For e greater than 0.65, however, k increases much faster than

my as a result of which c

v also increases more rapidly (see Fig.

10.25).

The effect of the magnitude of pressure increment on cv

is shown in Fig. 10.28 and 10.29. For small pressure increment

ratios (0.1 <--". <0.2) the coefficient of consolidation is con-

siderably lower for the standard increment ratio of 1. The two

rest periods of 7 days and 94 days show similar values of cv.

10.1.6 Discussion of results

The method of determining the swelling pressures of un-

disturbed samples of London clay, with high initial suctions, in the

standard oedometer, is open to criticism, mainly on two grounds.

Firstly, and more important, unless specimens can be prepared which

Calculated from cv and m

v.

192.

are 100% saturated, and bedding error is completely eliminated therp

may be considerable volume increases of the samples, even if the

dial reading is held constant during the performance of the swelling

test. This migration of water into a sample reduces the suction

resulting in a measurement of the swelling pressure which is con-

siderably smaller than the initial suction.

Secondly, the stress changes that occur in a sample during

a swelling test are difficult to ascertain. It is true that, if

the condition of no volume change can strictly be maintained, a

saturated specimen in the oedometer will be acted upon by isotropic

stresses. In practice, however, due to swelling, the horizontal

stresses tend to increase considerably as a consequence of lateral

confinement and may be much higher than the vertical stress at the

end of a swelling test:

It must be mentioned, however, in this connection, that

Skempton (1961) and Ward, Samuels and Butler (1959) have obtained

reasonably correct indications of initial suction of undisturbed

samples of London clay from swelling tests. It is, therefore,

believed that provided fully saturated samples can be prepared and

bedding errors completely eliminated swelling tests may give

satisfactory results.

The initial swelling mentioned above has an important

bearing on the deformation of the clay during subsequent loading.

Numerical values are presented in Section 10.4.

193.

Firstly, the void ratio of a sample before test, calculated backwards

from the final measurements and changes of height, does not accurately

correspond with the void ratio determined from the initial moisture

content. The former is always higher, althouth applying the

corrections for apparatus deformation, a better correspondence is

achieved.

The initial swelling also affects the compressibility of

the clay in a significant way. This is shown in Fig. 10.30 where

the average e - log p relationship of the Ongar clay, obtained

from all the tests reported previously, is plotted. It will be

noticed that the curve starts at a void ratio which is higher than

the in-situ value and has, therefore, a greater slope than a field

consolidation curve at low stresses. This is easily understood

from a comparison with the e vs log p relationship for the con-

trolled rate of strain tests in which there was little scope for

initial swelling! The compressibility data for first loading pre-

sented in Figs. 10.17 - 10.20 are, therefore, likely to be higher

than the corresponding values in the field. It is noted, however,

that when the in-situ vertical effective stress (2.0 T/ft2) is re-

established the in-situ void ratio is also restored.

In many settlement analyses of structures on over-con-

solidated clays the use of the second loading curve has been pre-

ferred on the ground that sampling disturbances give higher com-

pressibility or first loading (Burmister 1952, Hansen and Nisi 1961).

For details see section 10.3.

194.

It may now be added that the initial swelling is another reason why

the use of the first loading curve may result in the settlement

being overestimated.

The pressure - void ratio relationships are not significantly

affected by rest periods at the in-situ effective stress of 1 or 7

days nor does the magnitude of load increment subsequent to these

rest periods alter noticeably the shape of the curves which are

obtained from standard tests, i.e. doubling the load (i.e. AD

= 1)

every day. This result is in accordance with the findings of

Lewis (1950), Northey (1956), Leondrdd and Ramiah (1959), Simons

(1965) on normally consolidated clays. Lewis used pressure in-

crement duration of 24 hours and 48 hours, Northey 20 minutes and

24 hours, Lennards and Ramiah 4 hours, 1 day and 1 week and Simons

5 hours, 1 day and 1 week. They all came to the conclusion that

the effect of load duration on the pressure.- void ratio relation-

ship was insignificant.

The picture changes considerably with the longer rest

period of 90 days at the effective overburden pressure. During

this period of sustained loading the samples exhibited some second-

ary consolidation but, perhaps of more importance, also developed

marked resistance to further compression. The use of slow rate of

loading in the range 2-4 T/ft2 for Ongar and 3.6 T/ft2 for-OxfOrd

Circus (at 0.4 T/ft2 per day and 0.1 T/ft2 per day for Ongar and

0.6 T/ft2 per day and 0.15 T/ft2 per day for Oxford Circus) showed

195.

that the compression was very small for pressure increase of up to

20%, but increased rapidly beyond this point. However, a pressure

increment ratio of 1 tended to conceal this effect. This phenomena

which has often .been observed in soft normally consolidated clays

(Leonard and Ramiah 1959, Bjerrum 1967) is attributed to the develop-

ment of a rigid bond after a long period of rest (Terzaghi 1941)

which is susceptible to sudden breakdown when a "critical pressure"

or ailluasi-preconsolidation" pressure is exceeded: A similar

argument may conceivably be extended to an undisturbed sample of

clay. While in-situ, this clay has been under the effective over-

burden pressure for many years during which the particles have

developed a certain structural arrangement. On sampling, the over-

burden pressures are removed and, even if there have been no

mechanical disturbances, this release of stress alone will cause

some alteration of the original structure of the clay. When this

sample is reloaded in the laboratory following the normal procedure

of daily load increments, the clay particles may not have the

opportunity to recover their original structural arrangement. This

will be particularly so if the sample undergoes considerable swelling

during the initial stages of a test. When the in-situ vertical

It is interesting to note in this connection that the presence of a threshold gradient has been observed by Raymond and Low (1963) for flow of water in clays. This threshold gradient, which is the hydraulic gradient below which no flow occurs, decreases with decreasing clay concentration and increasing temperature.

196.

effective stress is restored,and the sample allowed to rest for a

long period of time the clay particles may then readjust themselves

to a more stable structure, which may not be effectively broken until

the pressure exceeds a certain value. This phenomena for stiff

clays was first observed by Langer(1936) who found that for small

pressure increments there existed a "threshold" value beyond which

appreciable volume change first occurred. The only field evidence

to support this was reported by Terzaghi (1941) who observed that a

point situated at a depth of /00 ft. below the Charity Hospital,

New Orleans did not settle at all even though the pressure at the

point had certainly increased due to the loading. It is, therefore,

believed that for London clay also a threshold value may exist which

must be exceeded before any appreciable settlement may occur and

from the data presented for Oxford Circus and Ongar a tentative

suggestion of 10% of the in-situ vertical effective stress is made.

The permeability (k) and the coefficient of consolida-

tion (cv) are found to be functions of void ratio and for both

first and second loadings unique relationships are obtained. In

fig. 10.31 the cv data from some other sites in London have been

compared with the results from Oxford Circus and Ongar. The close

correspondence between the latter two is perhaps fortuitious, but

it is nevertheless, believed that at lower void ratios, i.e. higher

effective stresses, the variation between the different sites would

diminish. At void ratios near the in-situ values, however,

197.

different patterns of joints and fissures give rise to different

coefficients of consolidation and it is not until higher pressures

are applied that these fissures close to allow the various samples

to approach uniformity.

198.

10.2 Tests in the high pressure (hydraulic)oedometer

10.2.1 Presentation of results

The procedure of these tests has been described in Chapter

9. Two tests were performed on the Ongar clay, but only one was

successfully completed, the other having to be terminated due to a

leakage in the membrane when consolidation was progressing in the

range 240 - 360 lbs/iZ

The maximum pore pressures recorded on stress increments

under undrained condition before consolidation commenced, are given

in Table 10.3. At low effective stresses pore pressure response of

100% was achieved, which immediately indicated that the samples were

fully saturated. With increasing effective stresses, however, the

maximum pore pressure response steadily decreased. No systematic

record of the build-up of pore pressure with time was kept, but it

normally took a few hours for the pore pressure to reach its peak at

high effective stresses, while at low stresses the response time was

considerably smaller. However, at least 12 hours was allowed for

the pore pressure to equalise before consolidation was started:-

The time/settlement, time/volume change and time/pore

pressure diagrams during consolidation are given in Figures 10.32

- 10.41. The results are presented in the form of time vs. degree

of consolidation and time vs. dissipation of maximum pore pressure

curves for each pressure increment, the back'pgessure during con-

solidation being always 15 lbs/inZ Each pressure was maintained for

199.

top 24 hours or until the measured pore pressure at the/equalled the back

pressure, whichever time was greater, except for the increment

60 - 120 lbs/in2 of test no. HMO/H0/2 which was kept for a period of

5 weeks (Fig. 10.42). For each pressure increment the degree of

consolidation has been expressed in terms of total settlement or

total volume change occurring at the end of the consolidation stage.

In Tables 10.4 and 10.5 are summarised the results of the

two tests. It will be noticed that for both tests the measured

volumetric strains were generally higher than the corresponding

axial strains. In Fig. 10.43 are plotted the ratio of these

strains against time for all the consolidation stages and in Fig.

10.44 substantially the same data are plotted as a relationship

between axial and volumetric strains for different periods of time.

It will be noticed that over most of the range an almost linear re-

lationship between the strains (i.e. a constant strain ratio) is

obtained and no consistent variation with time can be discerned.

The amount of volume change causing this difference between the

axial and volumetric strains is plotted against time in Fig. 10.44.

It is clear that although there is a slight variation with time much

of the discrepancy arises due to the difference in strains during

the first 100 minutes of consolidation. For test no. HPO-H0-2 the

difference between the two strains was even greater. That leakage

and diffusion through the membrane were to a large extent responsible

for this at least in the latter test is clear from Fig. 10.42 where

200.

the plots of both axial and volumetric strains with time follow

substantially the same curve up to about 90 minutes, after which

they diverge giving higher values of the volumetric strain. It

has been mentioned already that this test had to be terminated when

consolidation was in progress in the pressure range 240 - 360

lbs/in2 due to a leakage in the membrane.

The values of the coefficient of compressibility (mv),

calculated from both axial and volumetric strains, and tabulated in

Tables 10A and 10.5, also reflect the discrepancies mentioned above.

In contrast, however, the coefficient of consolidation cv, all

calculated from t50, for axial strain, volumetric strain and

dissipation of maximum pore water pressure; do not show any sig-

nificant difference.

The theoretical curves for Terzaghi one-dimensional con-

solidation, all fitted at t50, have also been plotted in Figures

10.32 - 10.41. Because the cv values obtained from /ill and

by were similar, the corresponding theoretical curves for the

degree of consolidation were almost identical and for simplicity

only those for L\ II are shownT* The theoretical dissipation curves

have been obtained by fitting them at points where half the excess

pore pressures at the impermeable boundary had dissipated.

For 11100-110-2 cv values have been calculated for only

6H and Av. ** There were some differences in the case of HFO-H0-2.

201.

It can be seen that the Terzaghi theory predicts the rate

of settlement accurately for the degrees of consolidation (U) of

up to about 70%, beyond which due mainly to the influence of second-

ary and creep effects, the measured rate of settlement becomes slower.

The dissipation of maximum pore water pressure, on the other hand,

proceeds somewhat faster than predicted for U < 50% and slower

than predicted for U > 50%. The pressure void ratio relationships for the two tests

are shown in Fig. 10.46. While for HPO-H0-1 the void ratios have

been calculated backwards from the water content of the sample

measured at the end of the test, those for HPO-H0-2 have been cal-

culated from the initial measurements. It will be seen that for

the first test the initial void ratio, calculated from changes of

height corresponds favourably with that determined from initial

moisture content of the sample - the small swelling may have been

due to water being sucked in from the porous stone before the start

of the test and while water was flushed from the top to drive out

any air from between the membrane and the sample. On the other hand,

the correspondence of the initial void ratio with that obtained from

volume changes is less satisfactory reflecting the lack of agreement

between the water content of the sample measured at the end of the

test and that calculated from initial moisture content.and volume

changes! This is added evidence to suggest that measurements of

From measurements of volume change during consolidation and swelling the moisture content of the sample should have decreased from the initial 27% while, in fact, it was 27.5% which was in agreement with that obtained from change of height.

202.

volume change were somewhat in error.

10.2.2 Discussion of results

(a) Initial response of pore pressure

The initial response of pore pressure due to loading under

undrained conditions depends to a considerable extent on the flexi-

bility of the measuring system, which gives rise to a number of

important effects. First, even though all drainage connections

may be closed during a pressure increment some water must flow from

the sample into the connecting leads and to the null indicator it-

self - the two form a system which is not absolutely rigid - due to

the difference in pressures in the measuring system and in the pore

water at the instant of load application. This means, in effect,

that some "drainage" takes place before the equilibrium pressure

is attained and this is different from that set up in the specimen

when the pressure is applied. Secondly, there is often a con-

siderable time lag before equilibrium is reached. With high system

flexibility this response time may be very large. Thirdly, ex-

cessive flexibility of the measuring system may change the overall

compressibility of the pore phase thus altering the stress dis-

tribution between the soil skeleton and the pore phase.

Whitman et al (1961), Bishop and Henkel (1962) and Gibson

(1963) have analysed the effect of system flexibility on the pore

pressure response of clay specimens under undrained conditions.

-m2T/4 m= cc

2 7 '\ eXP U.

uo 1 1 ÷

(lo.a.)1) L, =,c,2 ± 7 + 9 2 ) m=1

203.

Christie (1963, 1965), Northey and Thomas (1965) and Tan (1968) have

shown that in the conventional oedometer test with pore pressure

measurement, the time lag gives rise to erroneous measurement of

pore pressure in the early stages of consolidation. It was, there-

fore, decided in the present tests, to allow sufficient time for the

pore pressure to reach equilibrium before consolidation was commenced.

Whitman et al (1961) obtained an expression for the pore

pressure response of saturated clay specimens in terms of the flexi-

bility of the measuring system. A more rigorous analysis was done

by Gibson (1963) who derived the following expression, based on

Terzaghi's theory of consolidation, of the equilibrium pore pressure

as well as the times to reach various degrees of equalisation:

where m etc are roots of 0(mCoto(

m = 0

and

uo is the initial pore pressure in the

equalisation begins

u

is the initial pore pressure in the

ut is the pore pressure observed after

uoo is the equilibrium pore pressure at

1") is the stiffness factor, defined as

sample before

measuring system

time t

t = oe-)

204.

Ahm =

where = volume factor of measuring system (in3 per lbs/in2)

my = compressibility of the sample (in2/lb)

h = height of sample (in)

\ A = cross-sectional area of the porous stone* (in

2 )

The equiliktium pore pressure at time t = is given by,

** U co

uo 1 -

u.

uo (10.242)

U cO.

It is obvious that — decreases with decreasing "7 o ' u so low compressibility coupled with high flexibility results in a

106W pbrd pressure response (Bishop and Henkel 1962).' In the

present tests the flexibility of the measuring system was due to

the null indicator (1 x 10 6 in3/p.s.i.) and 4 ft. of Saran tubing

(1 1 x 10-6

in3/p.s.i. per foot) connecting the null indicator to

the sample - atotal of 5.4 x 10-6 in3/p.s.i. Assuming that there

was no air trapped in the system the value of X can therefore be

taken as 5.4 x 10-6 in3/p.s.i. Taking the values of .Mv from Fig.

The above expression applies strictly to cases where the porous stone is of the same diameter as the sample. In the present case the diameter of the stone is only i" compared to the sample diameter of 4". ** For the particular case of ui = 0, equation•10.2,,j,...167.the same as that given by Whitman et al (1961), i.e. u,o/uo = 1/1 + B where B = 1/-9.

205.

10.48, the peak pore pressure response for each pressure increment

has been calculated and shown in Table 10.3. It will be seen that,

theoretically, the maximum response should decrease from 98% at the

effective stress of 15 p.s.i. to 84.3% at 360 p.s.i. while the

measured values show a decrease from 100% to 88% (see Fig. 10.47).

The agreement is, therefore, satisfactory.

The theoretical response times have also been determined

from charts given by Gibson (1963). For 98% equalisation (i.e. 1 - ut/u00

equalisation factor = 0.02) the times (see Table 10.3) 1 - u./u

vary from 15 minutes at loW aresses to 6.7 hours at high stresses.

No systematic record of pore pressure vs. time was kept, but, as has

been said earlier, at least 12 hours were allowed to achieve equili-

brium before consolidation was begun.

(b) Measurement of strains

The discrepancies between the axial and volumetric strains

during consolidation are difficult to explain. While for EEO-H0-2

much of these can be attributed to leakage and diffusion of water

through the membrane the same cannot be confidently said of HPO-H0-1.

As shown in Fig. 10.43 axial strains were smaller throughout the

process of consolidation. In a perfect oedometer test on saturated

clay with lateral restraint, the axial strain must equal the

volumetric strain. In practice, however, a number of factors may

influence the strains during consolidation:

(a) incomplete saturation of the specimen

206.

(b) inadequate bedding

(c) lateral deformation due to expansion of the ring

(d) side friction

It has been shown above that almost 100% response of pore pressure

was achieved during the early load increments under undrained con-

ditions. Also for higher stresses the response was close to the

theoretical prediction for saturated samples. This indicates that

the degree of saturation, for all intents and purposes, was almost

100% and could not account for the difference in strains.

Any inadequate bedding would result in a relatively

greater axial strain and would also cause an initial settlement to

be recorded, none of which were noted in the actual tests.

Lateral deformation of the confining ring could be

neglected because a 3i in thick annular ring could hardly be ex-

pected to undergo substantial expansion at maximum pressures of

500 lbs/in

One of the major advantages of the hydraulic loading system

and a flexible membrane is to minimise the effect of side friction

on measurements of axial strain if the latter is made over a central

area of the sample. It is believed that the effect of any side

friction would be restricted to a narrow zone near the periphery

of the sample and a bowl-shaped deformation surface with a flat base

would result. The settlement over much of the central area would

be uniform and greater than that at the edge. This would corres-

207.

pondingly result in higher axial strains at the centre and smaller

volumetric strains. The measurements were, however, to the con-

trary. The discrepancies cannot, therefore, be explained by side

friction.

It has been mentioned before that the final moisture con-

tent of the sample (1110-H0-1) corresponded more closely to the

axial strains during the test. If the measured volume changes

were correct, the moisture content of the sample should have de-

creased from the initial 27.0% to the final 24.5% while it had in

fact gone up to 27.5%. (Calculated from axial strains, it should

have been 27.2%). It is, therefore, believed that the measurements

of volume change were somewhat in the error.

Possible reasons for this could be:

(a) Leakage at various connections

(b) Diffusion of water through the membrane

(c) Flexibility of the tubings connecting the sample to

the volume gauge and/or any air in the system. At

the instant when consolidation is starthd the pressure

in the system drops from the initial equilibrium

pore pressure to the applied back pressure causing

an initial volume change in the system.

(c) Consolidation characteristics

The rate of settlement predicted by the Terzaghi theory,

fitted at t50, is sufficiently accurate for undisturbed samples

208.

of London clay of the size (4" dia. x 1" high) tested in the high

pressure oedometer. In fact, for 1.14( 70% the measured rate of

settlement agrees extremely well with the theory. For U› 70%,

however, the measured rate is slower than that predicted. This is

primarily due to the influence of secondary consolidation. The

dissipation of maximum pore water pressure at the impermeable

boundary on the other hand, is not predicted quite as accurately

over most of the range by the Terzaghi theory. At early stages of

consolidation, dissipation proceeds faster than predicted while at

later stages it is slower.

It is well understood that the Terzaghi theory of con-

solidation is based on the Assumptions that compressibility, (mv),

permeability (k) and the coefficient of consolidation (cv) re-

main constant during the consolidation process. In real soil,

however, not one of these assumptions is strictly correct. In

recent years, many attempts have been made to formulate theories of

consolidation taking, at least, some of the above variabilities into

account. Schiffman (1958) considered the variation of permeability

during consolidation while Hansbo (1960) considered the same problem

for consolidation by sand drains. Davis and Raymond (1965) derived

a theory for non-linear consolidation, assuming the coefficient of

consolidation to be constant and a linear relationship between logarithm

Of pressure and void ratio. Barden (1965) solved the more general

case of the above problem taking the variation of permeability also

209.

into account. Both the above theories, however, strictly apply to

small strains only. Gibson et al (1967) derived equations govern-

ing the one-dimensional consolidation for large strains, taking

variation of compressibility and permeability into account and re-

casting Darcy's law in a form which relates the relative velocity of

the soil skeleton and the pore fluid to the excess pore pressure

gradient. For the London clay from Ongar and Oxford Circus the

change of void ratio during consolidation is small and therefore the

small strain theories can be applied without much error.

Davis and Raymond (1965) have found that if compressibility

does not remain constant during consolidation but varies according

to linear e vs log p relationships, the dissipation of maximum

pore pressure is a function of the pressure increment ratio. But

if cv is constant, which means that both m

y and k decrease in

such a way that their ratio remains the same, the rate of settlement

is independent of the pressure increment ratio and is identical to

that given by the Terzaghi theory. It will be seen from Fig. 10.46

and from the results of the conventional oedometer tests presented

in Figs. 10.4 - 10.14 that for a small stress increment, the e vs

log p relationship is very nearly linear. The coefficient of

consolidation is, however, not constant at lower effective stresses,

though at higher stresses the variation of cv with pressure is not

great (Table 10.4). Over much of the range, therefore, the conditions

of Davis and Raymond's theory are at least approximately satisfied,

210.

and so Terzaghi theory predicts the percentage settlement with

remarkable accuracy for 11( 70%. As the above theories do not

take account of secondary consolidation it is not surprising that

for higher values of U correspondence between measured and pre-

dicted rates of settlement is less satisfactory.

In the case of the dissipation of pore pressure, both

Davis and Raymond's and Barden's work show that the rate of dissipa-

tion is to a great extent influenced by the pressure increment

ratio. For Ap/p = 1 or less, however, the difference between

the above theories and that of Terzaghi is not significant. In

the tests reported here, Lip/p was mostly equal to 1 (except for

the last two increments of HPO-H0-1, where h. p/p were respectively

equal to 0.5 and 0.35). The discrepancy between the theoretical and

observed rates of dissipation cannot, therefore, be fully explained

as being due to the variation of compressibility and/or permeability.

The presence of random fissures within the samples may contribute

towards quicker dissipation in early stages of consolidation because

the water may then have the opportunity to flow along the fissures

having higher permeabilities than the overall sample. This effect

may not predominate at later stages of consolidation when the pore

pressure gradient within the sample is small, and fissures may have

closed up due to the increasing effective stress.

In Figs. 10.48 and 10.49 respectively, the values of mv

and cv obtained from these tests are compared with those from the

211.

conventional tests. For reasons mentioned earlier the mv values

determined from axial strains only are considered. As can be seen

from Fig. 10.48 the hydraulic oedometer gives smaller values of

compressibility than the first loading conventional tests over the

entire stress range, primarily because the specimens had lesser scope

for initial swelling (see section 10.1.6).

The coefficient of consolidation cv from these tests

shows the same type of ,variation with void ratio as from the con-

ventional tests (Fig. 10.49). Also, the cv values calculated

from axial strains, volumetric strain and the dissipation of maxi-

mum pore water pressure do not show any consistent variation and are

similar to those obtained for the conventional tests.

(a) Side friction

It has long been recognised that side friction may have

an important influence on the results of conventional oedometer tests.

This friction arises from the lateral restraint provided by the

oedometer ring during consolidation and acts in opposition to the

applied load thus causing a smaller load to be transmitted to the

sample than actually applied at the boundary.

Taylor (1942), Hansbo (1960), Nakase (1963) have found

that the coefficient of side friction lies generally between 0.15

and 0.30 and that the reduction of the effective consolidation

pressure due to friction in a conventional oedometer depends on the

relative dimensions of the sample and the magnitude of the applied

212.

pressure. For a height to diameter ratio of- this reduction

may be between 2 to 10 percent.

In the hydraulic oedometer both the pore pressure and the

settlement are measured over a central area of the sample where the

effect of side friction may be expected to be very small. That

the measured pore pressures before consolidation were not significant-

ly affected by friction has already been demonstrated. The maximum

response of less than 100% was due to flexibility of the measuring

system and not due to side friction reducing the pressure transmitted

to the sample.

It has Also been shown that any major influence of friction

on the settlement and volume change of a specimen would result in a

smaller volumetric than axial strain being measured. That this was

not so can be considered as added evidence for the insignificant

effect of friction on the process of consolidation in the hydraulic

oedometer. The close correppondence between the compressibility

ddtermined from conventional tests and from the high pressure oedo-

meter tests also suggests that for London clay which undergoes only

small strains during consolidation, side friction plays only a minor

role in influencing the results of the standard oedometer tests.

10.2.3 Advantages of the hydraulic oedometer

The high pressure oedometer and its hydraulic pressure

system have a number of advantages over the conventional oedometer and

213.

its deadload-lever loading arrangement. More important of these

are the following:

(a) Uniform pressure on the sample can be maintained by

applying the pressure on a flexible rubber membrane which allows

the sample to deform freely. In the conventional oedometer, the

load being applied on a rigid plate, uniform displacement does not

ensure uniform pressure.

(b) By measuring the settlement and the pore water pressure

over a central area of the sample, the effect of side friction can,

to a very large extent, be eliminated.

(c) The sample is not subjected to any shock loading which

is inevitable in the dead-load lever system.

(d) The error due to deformation of the apparatus, which may

be significant for stiff samples undergoing small settlement, is

reduced to negligible proportions.

(e) In the conventional oedometer there is little control over

drainage, which begins as soon as the load is applied and, perhpps,

before the pore pressure has had the time to reach its peak. With

the hydraulic oedometer and employing a rigid measuring system, high

pore pressure response can be:achieved after relatively short tithes,

during undrained loading, before consolidation is started. Drainage

can thus be fully controlled.

(f) Consolidation can be carried out against back pressures.

This not only allows tests to be performed under conditions more

214.

similar to those in the field but, for samples not fully saturated,

a sufficiently high back pressure can be applied to Unsure saturation

before consolidation is begun.

(h) By measuring settlement, volume change and the dissipation

of pore water pressure, complete consolidation behaviour of soils

can be studied.

The idea of the hydraulic oedometer is, however, not new.

Lowe et al (1964); Whitman and Miller (1965), Rowe and Barden (1966)

have successfully employed the principle of hydraulic loading system

in the design of consolidation cells which have been satisfactorily

used to study the consolidation characteristics of soils.

cap. Lowe et al applied hydraulic pressure on a rigid loading

215.

10.3 Controlled rate of strain tests

10.3.1 General

The main purpose of these tests, the procedures for which

have been described in Chapter 9.1.3, was to study the compressibility

of London clay when strained continuously in such a way that no excess.

pore pressure could develop. It has been shown in section 10.2, that,

following a long rest period, the rate of loading may significantly

influence the subsequent deformation of London clay. In the field,

the rate of loading is, in general, considerably slower than in the

conventional laboratory tests and in certain conditions - such as

during deposition - the loading rate is too slow to create any ex-

cess pore pressures. Leonards and Altschaeffl (1964) have shown

from experiments on artificially sedimented clays that the rate of

loading has considerable influence on the compressibility of clay

during deposition. It must be emphasized in this connection that

any measurable loading rate that can be produced in the laboratory

is still likely to be several orders of magnitude faster than the

rate at which a clay is deposited in nature. Nevertheless, a com-

parison of results of conventional step-loading tests with those of

controlled rate of strain tests may serve as a useful guide to assess

the difference in behaviour of the clay under different conditions of

loading.

The opportunity was also taken to study the deformation

of London clay at high pressures. Much work has been done in recent

216.

years to study the behaviour of granular materials at high pressures,

but very little data is available of similar work on clay. Terzaghi

and Peck (1948) described the stress - strain relationships for

loose and dense sand at pressures of up to 14,000 lbs/in2 while

similar data were presented for sand and ground quartz by Roberts

and De Souza (1958). De Beer (1963) reported results of compression

and penetration tests on dense sand at pressures of up to 45000

lbs/in2. Vesic and Barksdale (1963) tested samples of dense sand

at 9,000 lbs/in2 and Bishop (1966) presented results of compression

tests on dense and loose sand at cell pressures of up to 4,000

lbs/in2. Recently Vesic and Clough (1968) published results of a

comprehensive series of isotropic compression and triaxial tests on

dense sand at maximum confining pressures of up to 45,000 lbs/in,2

All these works have shown that there is considerable breakdown of

particles at high pressure and that both the Mohr envelope and the

shape of the volume change curves during shear are significantly

affected.

Available data on the behaviour of clay at high pressures

is, on the other hand, limited. Most laboratory oedometer tests on

clay are taken to pressures of about 32 T/ft2 (500 lbs/in2) which is

generally sufficient for purposes of design. Moreover, for normally

consolidated clays, this pressure is usually high enough to define

the virgin part of the pressure - void ratio relationships., For

many over-consolidated clays, on the other hand, this pressure is

217.

often insufficient to define even the pre-consolidation pressure let

alone the virgin consolidation behaviour.

Smith and Redlinger (1953) published a pressure void ratio

relationship for the heavily over-consolidated Fort Union clay shale

for pressures of up to 500 T/ft2 (7,500 lbs/in2) and deduced from it

a pre-consolidation pressure of between 80 - 100 T/ft.2 Similar

high pre-consolidation pressures have been estimated from geological

evidence for the Bearpan shale (Petersen 1954, 1958, Terzaghi 1961),

but no high pressure test has been reported. Brooker and Ireland

(1965) performed a series of one-dimensional consolidation tests on

five remoulded clays, using pressures of up to 2,200 lbs/in2 to

study the relationship between stress history and the coefficients

of earth pressure at rest. Bishop, Webb and Lewin (1965) published

the strength - effective stress relationships of undisturbed London

clay for confining pressures of up to 1,100 lbs/in2 and noticed a

"marked change of slope of the Mohr envelope in passing from low

stress to high stress range".

10.3.2 Pressure - void ratio relationships

The e vs log p relationships of all the controlled rate

of strain tests are shown in Figs. 10.50 and 10.51. As already

mentioned in Chapter 9, two tests were performed on the Ongar clay,

in one of which (ORS-H0-1) the sample was loaded to 386 T/ft2

(6,000 lbs/in2) and then unloaded, while in the other (CRS-H0-2)

218.

the specimen was first loaded to 152 T/ft2 (2,400 lbs/in2), unloaded

to 0.72 T/ft2 (10 lbs/in2) and reloaded to the maximum pressure of

423 T/ft2 (6,600 lbs/in2) before being finally unloaded to zero

pressure. The two Wraysbury tests (CRS-W-1 and 2) were taken

straight to maximum pressures of 435 T/ft2 (6,800 lbs/in2) and 417

T/ft2 (6,500 lbs/in2) respectively and then unloaded.

The first point to note from Figs. 10.50 and 10.51 is that

the final void ratios of the specimens determined from the moisture

contents measured at the end of the tests correspond quite accurately

with those determined from the initial void ratios and the changes of

height* during a test. This indicates that the samples did not

undergo any initial swelling for.whichin any case, they had little

scope, being held between the pedestal and the head of the loading

machine both of which were extremely rigid. This is in contrast

with the results of the conventional oedometer tests reported

earlier in which the considerable initial swelling caused large

discrepancies between the two measurements of the void ratio.

It will be noticed that the early parts of the pressure -

void ratio diagrams are rather flat and when the in-situ effective

overburden pressures are reached only small deformations have taken

place. Thereafter, the relationships follow continuous curvatures

up to pressures of 40 T/ft2 for Ongar and 80 T/ft2 for Wraysbury.

Taken as the average of the two reduced dial gauge readings.

219.

Beyond these points the e vs log p relationships are essentially

straight lines, except for the Ongar specimen CRS-H0-1 which showed

a slight tendency to curve off at pressures greater than 200 T/ft2.

At the end of each loading stage the specimens were left

overnight before unloading was started the following day. During

this period a small settlement was recorded which is believed to be

partly due to the dissipation of excess pore pressures* that developed

during loading and partly due to creep. However, this was only small

and the specimens soon started to swell when unloading began. Each

unloading curve has an initial flat tangent followed by a straight

line which finally ends up in another flat position towards the end

of the unloading stage. The most important thing to notice is that

all the unloading curves are remarkably parallel to one another

although there is a difference in slope between the Ongar and Wrays-

bury clays.

The only reloading curve (CRS-H0-2) indicates the pattern

observed by Schmertman (1953) i.e. a very flat initial position

followed by reloading along a straight line which is almost parallel

to the unloading line, and then continuing into the virgin curve.

The laboratory first loading curves are, in essence, results of

reloading following the geological rebound - but the above phenomena

Maximum pore pressures of up to 14% of the vertical stress was tolerated during loading and the effective stresses were corrected accordingly.

220.

are not so marked for the Ongar clays, although for Wraysbury, a

similar trend can be discerned.

10.3.3 Compressibility characteristics

Fig. 10.52 shows the relationship between effective stress

and volume change for one-dimensional consolidation of undisturbed

London clay for pressures of up to 435 T/ft2 (6,800 lbs/in2). The

volumetric strains have been calculated from the pressure - void

ratio diagrams shown in Figs. 10.50 and 10.51 and using the re-

lationship, 1\ e/1 + eo = AV/Vo for saturated clays.

The first thing to note in Fig. 10.52 is that an almost

unique curve is obtained for the volume change of the clay from

both Ongar and Wraysbury. This may seem surprcising, particularly

if it is considered that the two sites are more than 40 miles apart.

On the other hand the initial moisture contents of all the samples

are virtually the same - as are their index properties (see Table

4.1). Also, compared to the range of stresses considered, the

pre-consolidation pressures are too low (see section 10.3.4) to

significantly influence the deformation characteristics of the clay

at high pressures.

The shape of the volume change curve is similar to those

for dense and loose sand tested in isotropic compression at pressures

eo and Vo have been taken as the void ratio and the sample volume respectively at the beginning of each test.

221.

of up to 600 T/ft2 (Vesic and Clough 1968). The slope of the curve

changes continuously from the very steep initial portion to a flat

section towards the end of the pressure range. This is reflected

in Fig. 10.53 where the compressibility (mv), defined as

mv = 1/1 + e . Ae/Ap and calculated in the usual way from the

pressure - void ratio relationships, is plotted against effective

stress on a log - log scale. For comparison, the compressibilities

determined from conventional and high pressure oedometer tests are

also included. It will be seen that the relationship consists of

a straight line for pressures of up to 80 T/ft2 followed by another

straight line with a steeper slope. It should be noted that the

compressibility decreases 50 times(from 0.01 ft2/Ton to 0.0002 ft2/

Ton) for a pressure increase from 1 to 400 T/ft2 and, while mV is

still decreasing with pressure the rate of decrease is very slow at

the high pressures:

In Fig. 10.54 the variation of compressibility with pressure

has been plotted to natural scale for the stress range 0 - 20 T/ft2,

The results of the conventional and high pressure oedometer tests are

also plotted for comparison. It will be seen that at pressures

greater than about 4 T/ft2 all the methods of testing give essentially

the same compressibilities on first loading. At lower pressures,

however, there is considerable variation in the results, the con-

* The slope on a log - log plot does not give the true rate of decrease.

222.

trolled strain rate tests giving the lowest values of mv. The

reason for this will be clear from the pressure void ratio relation-

ships plotted in Fig. 10.30. Due to initial swelling in the con-

ventional oedometer a specimen consolidates from a void ratio higher

than that in-situ giving a e vs log p relationship which has a

greater slope at small stresses than the controlled strain rate

specimen which, because there has been little initial swelling, has

a much flatter e vs log p curve. Consequently the conventional

tests produce much larger compressibility than the controlled strain

rate tests. The results of the high pressure oedometer tests lie

in between the above two, presumably because these samples swelled

slightly, but certainly less than the conventional test specimens.

At pressures greater than 4 T/ft, however, the effect of initial

swelling is largely overcome and all the methods of testing give

essentially the same results - within the normal range of scatter.

The compressibility on second loading in the conventional oedometer

is, as expected, lower than all the first loading compressibilities

over most of the range except at very low stresses where only the

controlled strain rate tests give slightly lower compressibilities.

From a practical point of view the above results are ex-

tremely important. Beneath a typical foundation, the vertical

effective stress of an element of clay may increase by say

L = 1 (where p is the in-situ effective overburden pressure

before construction). For the Ongar clay under consideration this

223.

would mean an effective stress increase from 2 - 4 T/ft In this

range, the compressibility obtained from the first loading con-

ventional test is as much as 50% higher than that obtained from the

controlled strain rate tests. Therefore, for accurate settlement

prediction it is essential that in any testing programme proper care

be taken to prevent the initial swelling. Lack of this may be one

of the reasons for over-estimating the compressibility of over-

consolidated clays. For soft normally consolidated clays, on the

other hand, the initial swelling is usually small compared to the

total settlement and its effect may, therefore, be proportionally

less significant.

Crawford in a series of interesting papers (Crawford 1964,

1966, Hamilton and Crawford 1963), reported results of a large number

of oedometer tests on undisturbed normally consolidated sensitive

clays. He loaded the specimens both incrementally and at different

rates of strain and found that the pressure - compression relation-

ship was relatively independent of the method of loading, provided

that the average rate of compression was the same From results

presented above the same observation is found to be true for un-

disturbed London clay; except for the low stress range, the com-

pressibility and pressure - void ratio relationships are essentially

the same for both incremental loading and controlled rate of strain

Otherwise, secondary compression plays an important part in the case of sensitive clays.

224.

tests - within the usual range of scatter. The reasons for the

discrepancies at low stresses have already been discussed.

As mentioned earlier, one of the main effects of high

pressure tests on granular materials has been found to be the change

of grading due to particle breakdown. To investigate if similar

trend could be noticed in the case of London clay the grading curves

of the oven dried specimens CRS-H0-1 and 2, at the end of each test,

were determined and compared with that for Block 2, from which these

specimens were obtained. The results are shown in Fig. 10.55. It

is noticeable that both the after-test curves lie a little above the

pre-test grading possibly indicating a small amount of breakdown of

the medium and coarse silt fractions. Too much emphasis need not

be placed on this trend; however, because there was considerable

scatter in the pre-test grading curves of the different blocks.

The identical percentage clay fractions shown by all the curves is

perhaps fortuitious but it is clear that practically no breakdown

has occurred of the fine silt and clay fractions of either specimen.

It is concluded, therefore, that particle breakdown of clay specimens

tested at high pressures does not play a predominant part in their

behaviour.

10.3.4 Determination of pre-consolidation pressures

It has already been mentioned in connection with the

geological study of the London Basin (Chapter 4) that London clay

225.

as it exists today is heavily over-consolidated. Several attempts

have been made in the past to determine the maximum pre-consolidation

pressures of London clay at various sites. Cooling and Skempton

(1942) deduced from a study of oedometer tests on samples from the

site of Waterloo Bridge a pre-consolidation pressure of 20 - 30

T/ft.2 A similar value was proposed for central London by Skempton

and Henkel (1957). A lower pressure of 13.5 T/ft2 was suggested

from geological evidence by Skempton (1961) for Bradwell, to the

east of the London Basin, while Bishop, Webb and Lewin (1965)

obtained a minimum pre-consolidation load of 39 T/ft2 for Ashford

Common, to the west. In none of the above studies were the samples

loaded to sufficiently high pressures to define the virgin consolida-

tion behaviour of the clay and accurate determination of the pre-

consolidation load has, therefore, been difficult.

The significance and a method of determining the pre-

consolidation loads of clay specimens were first suggested by Casa-

grande (1936). The method is based on the observation that a

distinct change of slope of the e vs log p relationships of re-

moulded clay samples occurs in the vicinity of the laboratory pre-

consolidation pressures from which Casagrande suggested an empirical

construction for its determination. The method has been used for a

wide variety of normally consolidated clays. But for heavily over-

consolidated clays the laboratory e vs log p diagrams show con-

tinuous changes of curvature in the normal laboratory stress ranges

226.

(0 - 50 T/ft2) and often no distinct change of slope can be dis-

cerned. This is certainly true of all the oedometer tests on

London clay reported in this thesis which were loaded to maximum

pressures of 32 T/fd In such circumstances the Casagrande method

of determining the pre-consolidation pressure is not satisfactory.

Schmertmann (1953) proposed a method based on an original

suggestion by Rutledge (1942) of determining the pre-consolidation

pressure of over-consolidated clays. It is essentially based on

the assumption that the geologic rebound (i.e. the unloading e vs

log p curve associated with the removal of overburden) is parallel

to the laboratory rebound. However, this method leads to satis-

factory results only if the laboratory rebound is started from a

pressure sufficiently greater than the pre-consolidation load

(Strachan 1960), because at low pressures the rebound curves are not

always parallel (Crisp 1953). In the present thesis Schmertmann's

basic postulate has been used to determine the pre-consolidation

pressure of London clay from Ongar and Wraysbury.

It has already been shown in Figs. 10.50 and 10.51 that

the laboratory unloading curves for the high pressure tests are

remarkably parallel to one another. In Fig. 10.56 the e vs log p

relationship for the Ongar clay has been reproduced as an average

curve for the two tests. In this figure is also plotted the mean

unloading curve for the conventional oedometer tests 0-H0-1 to 11,

reported in section 10.1. The parallel nature of the three sets

227.

of unloading curves is very obvious. From this observation it is

postulated that the geologic rebound also followed a curve which is

parallel to the laboratory rebounds.

The consolidation curve for the slurry sample (initial

water content 89.6%) is also shown in Fig. 10.56. It is noticeable

that this slurry line meets the curve for undisturbed samples almost

as a tangent at a pressure of 40 T/ft. It has already been mentioned

that the undisturbed ellog p curve for the Ongar clay beyond 40

2 T/ft is a straight line and can therefore be considered as the

virgin line. At such high pressures, sampling disturbances may not

have any appreciable effect as can be seen from the remarkable close-

neas of the two curves in Figs. 10.50 and 10.51.

Now the question arises as to the position of the sedimen-

tation curve. In an ideal case this should be an extension of the

virgin line of a truly undisturbed sample. Yet this sedimentation

curve can, perhaps, never be determined accurately from laboratory

slurry tests because any slurry line will depend on the initial

moisture content of the sample. Even for a sedimentation test the

true rate of deposition in the field may be difficult to duplicate

in the laboratory. It is reasonable to believe, however, that

whatever the initial moisture content all slurry lines will tend to

converge at sufficirtly high pressures and ultimately become

tangential to the virgin line. For the Ongar clay, therefore, at

effective stresses greater than 10 T/ft2 the differences between the

228.

possible sedimentation curves will only be small and they will all

tend to converge to the virgin line at 40 T/ft. In the absence of

more information further analysis will be based on the slurry line

shown in Fig. 10.56.

Following the above hypothesis, the determination of the

pre-consolidation pressure becomes a relatively simple matter. In

Fig. 10.56, A represents the in-situ void ratio (e0)* and the

effective overburden pressure (P.) of the Ongar samples. From

A is drawn a curve parallel to the three laboratory rebound curves

to meet the slurry line at the point B. Because of the initial

flat portion of the laboratory unloading curves it is difficult to

define precisely the early part of the geologic rebound. However,

all the initial tangents are parallel to one another and, as shown

in Fig. 10.56, by extending the straight line portion of each re-

bound curve a point of intersection with the initial tangent is

obtained. It has been found that for each rebound curve, if p1

is the pressure corresponding to the point of intersection and p2

is the pre-consolidation pressure, the ratio log p1/log p2 is a

constant and is equal to 0.83. In order to determine the point

B a series of lines are drawn from the slurry line parallel to the

pre-consolidation pressure and the straight line portion of the

geologic rebound is extended. For each point of intersection the

e0 is the average void ratio of all the specimens, oedometer and triaxial, used in the experimental programme.

229.

ratio log p1/log p2 is calculated and where this ratio is 0.83

the corresponding point on the slurry line is taken as the pre-

consolidation pressure Pct. A smooth curve is drawn between B

and the straight line portion of AB to complete the geologic

rebound. Using this method a pre-consolidation pressure of 17

T/ft2 is obtained for the Ongar clay. Since the existing over-

burden pressure at the depth from which samples were taken is 2

T/ft2 erosion seems to have removed a pressure of 15 T/f0

Similar analysis of the Wraysbury test results (see

Fig. 10.57) gives a pre-consolidation pressure of 38 T/ft In

this case no slurry line is available, but the final virgin line

of the undisturbed samples has been extended to construct the

geologic rebound. In the region of high pre-consolidation

pressures such as determined for Wraysbury any deviation of the

slurry line from the one assumed can only be small and consequently

the shift of the point B is not expected to be great.

A comparison with the estimated values of the pre-

consolidation pressure for other sites shows that the pre-

consolidation load of 15 T/ft2 determined for Ongar agrees fairly

well with the value of 13.4 T/ft2 suggested for Bradwell (Skempton

1961), 20 miles east of Ongar. On the other hand, a pre-

consolidation load of 37 T/ft2 for Wraysbury confirms the value

of 39 T/ft2 determined by Bishop et al (1965) for Ashford Common,

230.

only 5 miles south of Wraysbury! These pre-consolidation pressures

correspond to the removal of 550 ft. of submerged sediments at

Ongar and 1,300 ft. at Wraysbury. This seems to substantiate the

suggestion made by Bishop et al (1965) that the thickness of the

eroded sediments may have increased in a westerly direction. On

the other hand geological evidence does not apparently support (or

contradict) the greater thickness determined for the western section

of the London Basin (Fookes 1966). Fookes, working on an early

paper by Wooldridge (1926) showed that the maximum probable thick-

ness of the London clay was approximately 500 ft. in Essex and may

actually have decreased westward. This discrepancy between the

geological evidence and the pre-consolidation pressures determined

from laboratory tests for the western part of the London Basin,

Fookes sought to explain by supposing that the thickness of the

late Eocene formations such as the Barton Beds, may have been

considerably greater in this section before they were removed by

erosion.

The method suggested in this thesis of determining the

pre-consolidation pressures of London clay, may be criticised

mainly on the following grounds:

(a) the position of the sedimentation curve is, to some

extent, uncertain and

For location of all these sites see Fig. 4.1.

231.

(b) how correct is the hypothesis that the geologic rebound

is parallel to the laboratory rebounds.

As already mentioned the sedimentation curve for a

natural clay during deposition is most difficult to determine.

Initial moisture content and/or the rate of loading play important

parts in influencing the deformation of remoulded clays in the

laboratory and unless sedimentation tests can be carried out at

very slow rates of loading and under conditions similar to those

in nature this uncertainty will prevail. But as explained earlier,

at high pressures and near the virgin curve, determined from tests

on undisturbed samples, the discrepancies between the possible

sedimentation curves are likely to be small and it is in this

range that the pre-consolidation pressures of London clay lie.

The phenomena associated with the geologic rebound of

over-consolidated clays have been discussed in Chapter 4. Depend-

ing on the nature of diagenetic bond developed after deposition

(Bjerrum 1965) a clay will have certain unloading characteristics

which may be difficult to reproduce in the laboratory. The only

possible way to verify the correctness of the hypothesis that

geologic rebound is parallel to the laboratory rebound is to check

the latter with the field pressure - void ratio relationships. No

data are available of the water content/depth profile for Ongar or

Wraysbury. However, it is possible to examine the data for Brad-

well and Ashford Common and compare them with the rebound curves

232.

for Ongar and Wraysbury respectively. The results are shown In

Fig. 10.58.

Assuming that erosion has reduced the effective stresses

by 13.4 T/ft2 at Bradwell and 38.5 T/ft2 at Ashform Common a series

of rebound curves* are drawn for various depths of the present soil

profiles. The probable pressure - void ratio relationships are

then plotted for the two sites as shown by the solid dots. From

water content - depth profiles given by Skempton and Bishop et al

the field void - ratio vs effective stress relations are then

plotted - indicated by the open circles. It will be seen that for

Bradwell the field e-log p curve lies somewhat above the labora-

tory prediction, but, except for the points near the surface, they

are remarkably parallel. For Ashford Common, on the other hand,

the field curve lies very close to the laboratory prediction and

is also parallel to the latter. If the pre-consolidation pressures

were accurately determined and the geologic rebound was parallel to

the laboratory rebound the field and laboratory curves would, of

course, have coincided. However, the fact that the field and

laboratory curves are parallel to one another for both sites - in

the range of stresses where direct comparison is possible - indicates

that the hypothesis of the parallel nature of the geologic and

Parallel to the Ongar rebound curves for Bradwell and Wraysbury rebound curves for Ashford Common.

233.

laboratory rebound curves is at least approximately correct. It

appears that for Bradwell the pre-consolidation pressure may have

been a little over-estimated although the possibility is not dis-

counted that the rebound curve at this site had a greater slope

than at Ongar at the higher stress range. For Ashford Common, on

the other hand, the closeness of the field and laboratory curves

suggests that both the pre-consolidation pressure and the rebound

characteristics are correctly determined. In any event, the data

plotted in Fig. 10.58 indicate that the pre-consolidation pressures

for sites to the east of the London Basin are lower than those for

sites to the west.

The field evidence presented here is too limited to allow

any definite conclusions to be reached. Nevertheless, it is clear

that the use of laboratory rebound curves provides a reasonable

means for determining the pre-consolidation pressure of over-

consolidated clays. One major requirement for the success of this

method is that the laboratory rebounds should be started from

pressures sufficiently greater than the pre-consolidation pressure.

Previous studies with Bearpaw shale (Peterson 1958) and Fort Union

clay shale (Smith and Redlinger 1953) have not shown good agreement

between field and laboratory curves. This lack of agreement has

generally been attributed to the large "time rebound" in the field.

It is also possible that this is due to the laboratory rebounds

being carried out from pressures which are too small. Moreover,

234.

softening and weathering may alter the moisture content of the clay

near the surface (Terzaghi 1936, 1961, Skempton 1948 ). To enable

valid comparisons to be made, therefore, water content data from

deeper borings in over-consolidated clays is required, together

with a knowledge of the existing porewater pressure distribution

with depth.

10.4 Tests in oedometers fitted with s:taii-Leages

10.4.1 Introduction

The purpose of these tests was to study the stress path

and to determine the coefficient of earth pressure at rest (Ko)

of undisturbed London clay during one-dimensional consolidation in

the standard oedometer. The equipment and experimental procedures

have been described in Chapter 9.4.

As pointed out by Tschebotarioft (1952) the term coefficient

of earth pressure at rest Ko was originally introduced by Donath

(1891), who defined it as the ratio fo the horizontal to vertical

earth pressure resulting in a soil from the application of vertical

load under condition of zero lateral yield. Initial efforts to

determine the value of Ko were made by Terzaghi (1920, 1925) who,

using the friction tape method obtained typical values of 0.4 to

0.5 for sand and 0.7 for a remoulded marine clay. Kjellman (1936)

found from one-dimensional compression tests on cubical specimens

of sand that the coefficient Ko was a function of the stress

235.

history of the soil. An apparatus for measuring the variation of

lateral pressure during one-dimensional consolidation was developed

by Binnie and Price (1941). Bishop (1958) used the triaxial

apparatus to measure Ko for granular materials. During one-

dimensional compression, he kept the axial and volumetric strains

of a specimen equal - a method also used by Simons (1958). A more

elaborate procedure, using the lateral strain indicator, was

adopted at Imperial College by Fraser (1957) and Sowa (1963).

(See Bishop 1958, Bishop and Henkel 196a). A modified triaxial

apparatus consisting of an inner and an outer cell with mercury

filling the annular space between the sample and the inner cell,

was used for London clay by Webb (1966) (See Bishop, Webb and

Skinner 1965). Kjellman and Jacobsen (1955) tested cylindrical

samples of pebble and crushed rock which were enclosed in a series

of steel rings, the elastic extension of which served to indicate

the magnitude of the lateral stress. The same principle was

adopted by Cebertowicz and Wedzinski (1958) who measured the lateral

deformation of the oedometer ring during one-dimensional compression

of granular materials.

Another method, known as the "cell test', has been used

by Gersevanoff (1936), and Davis and Poulos (1963). This is based

on the principle that, with a loading ram of the same diameter as a

cylindrical sample, the latter could yield laterally during axial

loading only if water was allowed to escape from the cell. The

236.

increase of pressure required to maintain constant volume of the

cell water, therefore, gave a measure of Ko. As pointed out by

Bishop (1958), however, this method would give reliable results if

there was absolutely no leakage, the cell was rigid and the water

in the cell was incompressible and free of all air. Jackson (1964)

tested specimens of clay in the triaxial apparatus using different

stress increment ratios, and interpolated the value of Ko corres-

ponding to E l/cv (the ratio of the axial to volumetric strain)

equal to 1. Hendron (1963) and Brooker and Ireland (1965) used

specially designed oedometer cells to conform to the condition of

zero lateral strain. Hydraulic pressure was applied in a chamber

surrounding the oedometer ring to maintain null balance condition.

All the work mentioned above refer to measurement of K

in the laboratory. Few attempts have so far been made to measure

Ko directly in the field. There is, of course, considerable

difficulty in this because the insertion of a measuring device

inevitably causes disturbances in the soil resulting in a change

of the state of stress! An indirect method of measuring in-situ

he r5.zoritaa stryieses has been proposed by Zeevaert (1953) based on

the determination of the pre-consolidation pressures in horizontal

and vertical directions of undisturbed specimens. A more satis-

* Recently Kenney (1967) has developed a device for ' measuring in-situ stresses in the Norwegian quick clays. It essentially consists of a large open pipe, instrumented with earth pressure gauges, which sinks under its own weight through the quick clay "without displacing the clay outwards".

237.

factory method, also indirect, based on the measurement of effective

stresses in undisturbed samples, has been used by Skempton (1961)

and Bishop et al (1965) to estimate large horizontal stresses in

natural London clay (see Chapter 8).

10.4.2 Variation of lateral stresses during consolidation

The variation of radial stresses during one-dimensional

consolidation in the strain gauge oedometers is shown diagrammati-

cally in Figs. 10.59 - 10.62. The test conditions were similar to

those in a conventional oedometer test. The lateral deformation of

the clay was restricted to the elastic extension of the confining

rings which also served to indicate the magnitude of the lateral

stresses. As already mentioned in Chapter 9.4, three tests were

performed on the Ongar clay, viz. O-SG-1, 2 and 3 in rings 1/32",

1/16" and in thick respectively.

At the end of the "swelling" stages of all tests, lateral

stresses considerably higher than the vertical "swelling" pressures

were measured (see Table 10.6) - the former being, on average, as

much as four times the latter. It has already been shown in Section

10.1 that considerable swelling took place during the performance of

a swelling test. But while a specimen could expand vertically due

to inadequate bedding it was restrained laterally by the confining

ring which caused large horizontal stresses to develop. There was,

however, no systematic increase of the lateral stress with thickness

238.

of the ring, the i" ring giving the minimum value.

As the tests were continued, following the procedures for

a conventional oedometer test, the following phenomena were observed.

At the moment of a load application the radial stress increased by

an amount depending on the vertical pressure increment and then

began to decrease during the progress of consolidation, finally

assuming a constant value at the end of dissipation. The reverse

behaviour was observed during unloading.

The variation of the lateral stress during consolidation

for each load increment of test nos. 0-SG..1 and 2* is shown in

greater detail in Figs. 10.62 - 10.66. It will be noticed that

at small pressures the maximum lateral stress was reached as soon

as the load was applied and then decreased with progress of con-

solidation. At higher pressures the lateral stress increased by

a small amount for a few minutes after the load application before

starting to decrease - a phenomenon which was more noticeable in the

case of the thinnest ring (test no. 0-SG-1). The reason for this

is believed to be the flexibility of the confining ring. A similar

phenomenon has been observed for dissipation of pore pressure in the

standard oedometer tests due to the flexibility of the measuring

system (Christie 1965, Tan 1968).

For 0-SG-3, the ring being j-" thick the variation of strain during consolidation was too small to be measured accurately with time. The initial and final values could, however, be noted correctly.

239.

The development of the maximum radial stress after load

application is shown in Fig. 10.67. It can be seen that except

for the first loading in test 0-SG-1, a maximum response of over

80% was usually achieved. For a perfect test with a completely

rigid system, this response should, of course, be 100%. But in-

complete saturation coupled with the lack of perfect rigidity of

the confining ring allowed a sample to undergo small initial de-

formations resulting in a response which was less than 100%. That

the flexibility of the ring played an important role in this is

indicated by the higher response achieved with the thicker rings.

10.4.3 Stress paths

The relationships between the vertical and radial stresses

measured at the end of consolidation are shown in Figs. 10.68 -

10.70. Except for test no. 0-SG-1 in which the specimen was un-

loaded from a vertical pressure of 250 lbs/in2 and then reloaded to

the maximum pressure of 500 lbs/in all the tests were taken straight

to the maximum vertical stress of 500 lbs/in2 and then unloaded. It

will be noticed that for each test the loading stress path consists

of two distinct sections - an initial flat portion followed by a

steeper straight line relationship between the vertical and radial

stresses. On unloading a familiar curve forming an hysteresis

loop was obtained.

It is necessary, at this stage, to differentiate between

240.

two values of th,"- coefficient of earth pressure at rest for over-

consolidated materials:

(a) Ko

defined as the ratio of the horizontal to vertical

effective stresses, ( 0- 1,/cr ,) and r v

(b) Ko defined as the ratio of the effective stress incre-

ments in the lorizontal and vertical directions required to main-

tain the condition of zero lateral strain, Ko cr t r v

The above distinction is necessary, because, for a

heavily over-onsolidated clay, the ratio of the existing effective

stresses (K ' may be high, depending on the over-consolidation

ratio, but the stress increments necessary to produce consolidation

under condition of zero lateral strain on the reloading cycle may

be quite different.

In Fig. 10.71 are plotted the values of Ko against

vertical effective stress while the variation of the incremental

coefficient Ko with pressure is shown in Fig. 10.72. It can be

seen that Ko

initially is very high but decreases sharply and

attains a steady value for vertical effective stresses greater than

100 lbs/iZ On unloading, Ko is always higher than the corres-

ponding value during loading, reflecting the effect of over-

consolidation. On the other hand, the increase of radial stresses

required to maintain the condition of no lateral strain is small

in the low stress range as indicated by the initial flat portions

of the stress paths shown in Figs. 10.68 - 10.70. Consequently

241.

Ro (see Fig. 10.72) is small for pressures below 100 lbs/in2 but

increases rapidly to attain a fairly constant value, approaching

Ko, over most of the stress range.

Fig. 10.73 shows the variation of Ko with over-consolida-

tion ratio during unloading. For comparison the results of tests

on remoulded London clay (Brooker and Ireland 1965) are also shown.

It will be noticed that Ko

is greater than 1 (i.e. effective

horizontal stresses greater than the vertical) for O.C.R. > 3 and

may be as high as 3.5 for O.C.R. = 40.

Table 10.7 shows the values of Ro obtained from the

three tests described above. Two values are quoted for each test

- one for the initial slope and the other for the final. It is

clear that Ro is small (0.28) for effective vertical stresses

less than 100 p.s.i. although over most of the range (100 ((5"v'

< 500 p.s.i.) a higher value 0.64 is obtained. In Fig. 10.74

these results are compared with the results of tests on remoulded

London clay (Brooker and Ireland 1965) and on undisturbed London

clay from Ashford Common (Webb 1966). It should be emphasised

that Brooker and Ireland performed their tests in a specially

designed oedometer cell where the condition of no lateral strain

was ensured by applying fluid pressure in a chamber surrounding the

oedometer ring to maintain null balance - while Webb performed his

test in the Ko - triaxial cell (Bishop, Webb and Skinner 1965).

In spite of the variations of testing techniques the close agreement

242.

between the results is remarkable. While the test on the remoulded

clay produces a straight line relationship starting from the origin

that on the undisturbed material from Ashford Common* shows a flat

initial portion similar to the ones for Ongar. Over most of the

range, however, all the curves are almost parallel, giving values

of Ro between 0.61 and 0.64.

In Fig. 10.75, the in-situ values of Ko estimated for

London clay at Bradwell (Skempton 1961) and Ashford Common (Bishop

et al 1965) are compared with the results of the laboratory tests

on the clay from Ongar. Although the Bradwell values lie a little

above the laboratory curve, the agreement, on the whole, is con-

sidered satisfactory.

10.4.4 Discussion of results

The principle adopted to measure the lateral stresses

during oedometer consolidation, as reported in this thesis, was

originally used by Kjellman and Jakobson (1955), Jakobson (1958)

and (ebertowicz and Wedzinski (1958). The major criticism of the

method is that the condition of no lateral strain is not strictly

enforced because small lateral deformations must occur to enable

the pressure to be recorded. Lower values of Ko measured for

granular material by Cebertowicz and Wedzinski (op. cit.) was

The initial state of stress in this test was an all round effective stress of 10 lbs/in2 to which the specimen was allowed to tswell before the K

o stage was carried out.

243.

attributed by Rowe (1958) to this lateral deformation.

In the present investigation, three different ring thick-

nesses were used to determine whether the lateral strain influenced

the measured values of the lateral stress in any significant way.

Since the thickness to diameter ratios of all the rings were smnil

(1/24 for the i" ring to 1/96 for the 1/32" ring) their behaviour

could be assumed to be similar to that of thin rings subjected to

uniform radial pressures: For such a ring the lateral strain is

inversely proportional to the thickness (Timoshenko and Goodier

1954) and, therefore, the passage from an -" ring to a 1/32" ring

would mean approximately fourfold increase of the radial strain.

Calculations based on the theory of elasticity and the measured

strains showed that under the maximum vertical pressure of 500 lbs/

in2 the lateral strain would be no more than 1% of the vertical

strain for the 1/32" ring. Thus the deviation from the condition

of no lateral strain, even for the thinnest ring, was very small.

Moreover, the lateral stresses measured at any vertical stress does

not show any significant variation with the thickness of the ring

(see Figs. 10.68 - 10.70) and consequently the values of Ro are

almost the same for all the tests (Fig. 10.72). Also the results

agree well with measurements under strict condition of zero lateral

strain by Brooker and Ireland (1965) and Webb (1966). It is con-

* Assuming, for the moment, that shear stresses due to side friction were absent.

244.

eluded, therefore, that the slight lateral deformation of the

sample caused by the elastic extension of the rings had no measur-.

able effect on the lateral stresses during consolidation.

The other area of uncertainty in the measurement of

lateral stresses with strain gauges fixed to the outside of an

oedometer ring is the influence of the shear stress caused by

friction between the soil specimen and the ring. Three major

effects of side friction may be considered:

(a) The effective stress transmitted to the sample at any

given load is likely to be less than the load applied at the

boundary and consequently the lateral stress set up in the specimen

may be less. However, as discussed in section 10.2.2 this effect

is unlikely to be very significant for London clay over most of

the range except at very low vertical stresses where relatively

high lateral stresses are measured.

(b) The side friction may also affect the measured value of

the lateral stresses because the calibration of the rings was done

with water pressure - i.e. in the absence of any shear stresses.

A rigorous analysis of this error is difficult, but an approximate

analysis can be made (see Appendix F) on the assumption that the

entire side friction can be replaced by a vertical face acting on

top of the inside edge of the oedometer ring. This shows that the

maximum possible error from this source alone would be less than

5%. Since the accuracy of measurement is probably not attainable

245.

to such precision due to such factors as the variations of tempera-

ture and general scatter this error can be neglected.

(c) Perhaps an important effect of side friction is that it

actually changes the distribution of stresses within the sample

thus causing different lateral stresses to be set up at the centre

and towards the edge of the specimen. The effect is likely to be

accentuated by the non-uniform stresses caused by the rigid loading

cap. Little is known on this point, however, but it is believed,

that where side friction does not reduce the average vertical stress

within the sample by any great amount the distribution of lateral

stresses will not be significantly affected.

The variation of the lateral stress during consolidation

(Figs. 10.62 - 10.66) reflects the decrease of Poisson's ratio ( )

with volume change of the clay. It is well known, from the

classical theory of elasticity, that at the moment of load applica-

tion - under undrained conditions - a Poisson's ratio of 0.5 is

implied for a specimen of saturated clay. With dissipation of ex-

cess pore pressures, however, the sample changes in volume which

results in a decrease of ,) until the latter attains its fully

drained value at the end of consolidation. According to the well

known expressiono - , therefore, should decrease

from 1 to its final value, with respect to effective stresses, when

consolidation is complete. The values of calculated from the

final values of Ro (see Table 10.7) show that Poisson's ratio,

246.

under fully drained conditions is small (0.21) in the low stress

range but considerably higher (0.39) in the high effective stress

range. • The influence of stress path on Poisson's ratio is, there-

fore, obvious. (More discussion of this point will be made in

Chapter 11).

The duration of the pressure increments was not long

enough to study the effect, if any, of creep on the value of K.

Some of the increments which were kept for 2 - 3 days, did not

show any variation of lateral stresses after 24 hours suggesting

that Ko does not change significantly with time (see also Bishop

1958a).

The empirical relation between Ko and the angle of

shearing resistance 0' given by Jaky (1944) for normally consolida-

ted soil,

Ko . 1 - Sin 0' (10.4.1)

has been found to give satisfactory results for a wide variety of

sands and clays (Bishop 1958, Simons 1958, 1960b). For normally

consolidated London clay 0' determined from tests on remoulded

samples lies in the range 17o - 20°, giving values of Ro, from

equation (10.4.1), between 0.66 - 0.72 (Bishop, Webb and Skinner

1965). The undisturbed 0ngar clay has 0' equal to 21° (see

Chapter 11), for which the Jaky expression gives R0 = 0.642.

247.

This agrees almost exactly with the value obtained experimentally

for higher pressures, i.e. in the normally consolidated range (see

Fig. 10.74).

Brooker and Ireland (1965) from tests on five widely

different clays concluded that equation (10.4.1) slightly over-

estimates the value of Ko for clays although, for sands, it has

proved satisfactory (Hendron 1963). Based on their results, Brooker

and Ireland found that a slightly modified equation

Ko = 0.95 - Sin 0' (io.4.2)

seemed to fit the test data better than the original Jaky equation.

Using the same results Schmidt (1966) suggested that another re-

lationship

Ko 1 - 1.2 Sin,0' (10.4.3)

was even better both in form and in fit. These modifications of

the Jaky equation may make only small differences in the value of

Ko, but it is perhaps significant that all the clays tested by

Brooker and Ireland show Ko values slightly lower than those

suggested by the Jaky equation - the discrepancy becoming greater

with higher values of the angle of shearing resistance.

The influence of stress history on Ko can be clearly

248.

seen from Fig. 10.73. During rebound, Ko increases steadily with

over-consolidation ratio (O.C.R.) and after the latter has exceeded

a certain value the horizontal effective stress becomes greater than

the vertical. For London clay Ko is greater than 1 when the

O.C.R. exceeds 3 and may be as high as 3.5 for an O.C.R. = 40.

It is believed that the maximum value Ko may attain at

any over-consolidation ratio is the coefficient of passive earth

pressure K which is dependent on the effective stress parameters

of the clay (Terzaghi 1943). Bishop, Webb and Lewin (1965) have

shown that in natural London clay the in-situ stresses may be con-

siderably in excess of the passive pressures based on the residual

strength of the clay but less than those based on the peak strength

parameters.

Recently, Schmidt (1967), from a study of Brooker and

Ireland's data on five clays found that during one-dimensional re-

bound the relationship between Ko and O.C.R. may be expressed by

the following empirical relationship,

K rb K ' 0

0-v'max n

(10.4.5)

Tv '

o where K rb is the coefficient for rebound

Ko' is the coefficient for 0.C.R..1 i.e. at the end of

the loading stage and Tv

' max is the over-consolidation ratio.

249.

According to this equation:: log (Korb) is proportional to log (OCR).

Plotting the average Ko vs O.C.R. relationship for the three Ongar

tests on a log - log basis a straight line is indeed obtained (see

Fig. 10.76) which seems to verify equation (10.4.5). For comparison

the corresponding line for remoulded London clay is also shown.

The only way to check the validity in nature of the re-

lationship between Ko and O.C.R., of the type shown in Fig. 10.73,

is to compare such data with field observations. It has already

been mentioned that direct measurement of horizontal stresses in-situ

presents almost insurmountable difficulties and, therefore, only in-

direct estimates have so far been made. The limited data available

for the sites of Bradwell and Ashford Common (see Fig. 10.75) show

reasonably good agreement with the laboratory results. It is worth

emphasizing, in this connection, that the over-consolidation ratios

of London clay at the two sites were calculated on the basis of

estimated values of the pre-consolidation pressures. The good agree-

ment between the field and laboratory data, therefore, indicates once

again that the pre-consolidation pressures at Bradwell and Ashford

Common have been estimated fairly accurately (see section 10.3).

However, more field data of this type should be collected before any

definite conclusions can be made.

The stress paths shown in Figs. 10.68 - 10.70 and the

variation of Ko

and Ko with pressure, plotted in Figs. 10.71

and 10.72, suggest an important result that may be of great

250.

practical significance. It is easily understood that the con-

solidation of London clay in the laboratory is essentially a re-

loading, following the geologic rebound in the field, until the

pre-consolidation pressure has been reached and this is reflected

in the initial flat portions of the effective stress paths and the

small Ro values measured at the low stress range. Previous work

on sand (Fraser 1957, Bishop 1958) has shown that Ro on re-

loading is considerably less than that on first loading. The

results presented above show a very similar trend. At higher

stresses - that is near the pre-consolidation pressure and beyond

- the clay behaves as a normally consolidated material with respect

to stress changes and consequently the Ro values are almost equal

to those obtained from tests on remoulded clays.

It must be mentioned, of course, that the effective

horizontal stress measured when the in-situ vertical effective

stress of the samples are restored in the oedometer is still con-

siderably smaller than the estimated in-situ horizontal stress (see

Table 8.4), shown by the solid dot in Fig. 10.74. This is thought

to be due to sampling disturbances and the initial volume change

that occurred in the samples during the performance of the swelling

tests. This means that it has not been possible to reproduce in

the laboratory the probable in-situ reloading stress path. Never-

theless, it is believed that the in-situ R0 in the low vertical

stress range will be small but comparable to the values of K0

251.

measured in the laboratory. And it is in this range that the in-

crease of effective stresses beneath a typical foundation problem

will lie.

The use of oedometer test results to calculate settlement

of structures implicitly assumes that the increase of stresses in

the field are similar to those in the oedometer. Taking, for ex-

ample, the points beneath the centre of a uniform circular load,

this means that the horizontal and vertical effective stresses in

the field should increase during consolidation in such a way that

8(Th 1/66vl = Ro = 0.28. It has been shown in Chapter 7 (see

Fig. 7.3) that this ratio of the effective stress changes during

consolidation depends on the pore pressure parameter A and the

Poisson's ratio and over most of the range is considerably

greater then the value of Ro. Therefore the stress paths in the

field and in the oedometer are likely to be quite different and con-

sequently the settlement calculated on the basis of the oedometer

test results, assuming one-dimensional strain, may be far from

correct. (This point has been qualitatively described, in more

detail, in Chapter 5 - see Fig. 5.11).

It follows, of course, that in cases where the ratio

A 0.-- t/Ao-v, in the field is similar to that in the oedometer settlement predictions based on the oedometer test results may be

correct. This is often true of normally consolidated clays having

high values of Ro and may explain the good correlation between

observed and calculated settlements of structures on such clays.

TABLE 10.1

MEASUREMENT OF SWELLING PRESSURES

Location Swelling pressure Initial suction measured from measured in oedometer tests triaxial tests

(ps) (Pk)

T/ft2 lbs/in2 Tift.2 - lbs/in2

Oxford 1.7 26.5 96.0 6.17 0.28 Circus

Ongar 0.45 7.0 58.5 3.76 0.12

252.

P-s Pk

TABLE 10.2

RESULTS OF SPECIAL

Test No.

Measured Dial Gauge Reading Swelling x 10-4 in.

ZS H Pressure (T/ft2) Initial Final (in)

Measured Change of Height (in)

Water Content (%)

Initial Final

T-H0-12 9;50 1542 1531 -.0011 +0.025 27.0 29.6 T-H0-13 0.40 1193 1190 -.0003 +0.008 27.3 29.3 T -HO -14 0.50 1625 1625 0 +0.020 26.9 29.2

Average 0.47 - -.0005 +0.018 27.1 29.4

SWELLING ltSTS

Degree of Saturation Weight of Sample Volume of sample Change of (%) (gms)

(c.c.)

Volume 6.1/

Initial Final Initial Final Initial Final

98.5 100.0 171.25 173.88 86.7 89.7 +5.46 96.6 99.5 169.02 170.39 86.7 87.7 +1.15 97.o 100.0 170.20 173.05 86.7 89.1 +2.77

97.5 99.8 170.16 172.44 86.7 88.8 +2.46

253.

TABLE 10.3

INITIAL RESPONSE OF PORE

Effective Pressure Peak change of ilu/tp stress p Increment pore pressure lbs/in2 LID 2 A.0 lbs/in2

(lbs/in ) HPO-HO-1 HPO-H0-2 HPO-H0-1 HPO-H0-2

15.0 15.0 15.0 15.0 1.00 1.00 30.0 30.0 30.0 30.0 1.00 1.00 60.0 60.0 58.0 59.0 0.97 0.98 120.0 120.0 110.0 106.0 0.91 o.88 240.0 120.0 110.0 - 0.91 - 360.0 125.0 110.0 - 0.88 -

A = Area of porous stone (ill dia) = 0.196 in2 H = Height of sample = 1 in N= Vol. factor of null indicator + 4 ft. of Saran

tubing = 5.5 x 10-6 in5/p.s.i. * * ui = Initial p.w.p. in measuring system before pressure

increment (i.e. back pressure) = 15 p.s.i. uo = Pore pressure in sample before redistribution

= Alp + 15.0 p.s.i. * * * Calculated for test no. HPO-H0-1, values from

Table 10.1

254.

11

WATER PRESSURE

Compressibility my (in2/lb) (Fig. 10.48)

71* Am v ui** -- ct

uo 1 114 /uo For 985 Equalisation

Time t(min)*** — T= X

H 2 = 1 • 7 _17

1 al ( vs \ uo (Gibson 1963)

0.00064 22.7 0.50 0.979 0.18 15.5 0.00050 17.9 0.33 0.965 0.19 36.5 0.00039 13.7 0.20 0.946 0.23 100.0 0.00024 8.6 0.11 0.907 0.29 196.0 0.00016 5.7 0.11 0.860 0.34 276.o 0.00013 4.6 0.11 0.843 0.39 2.106.0

TABLE 10.4

RESULTS OF TEST NO. HPO-H0-1

Initial water content

Effective stress ran:e

Initial void ratio

Axial Strain AH/T4

Vol. Strain Av/v% lbs/in2 I Ilf-TE

Ichange From

of height

From vol. change

27.0% 0-15 0-0.96 0.7408 0.7596 - 0.41 - 0.39 15-30 0.96-1.92 u0.7338 0.7528 - 1.02 - 1.18 30-60 1.92-3.86 0.7162 0.7328 - 1.85 - 2.02 60-120 3.86-7.72 0.6846 0.6985 - 2.43 - 2.91 120-240 7.72-15.43 0.6438 0.6505 - 3.07 - 3.38 240-360 15.43-23.14 0.5936 0.5965 - 2.11 - 2.60 360-485 23.14-31.2 0.5601 0.5561 - 1.80 - 2.05 485-360 31.2-23.14 0.5323 0.5249 + 0.48 + 0.50 360-240 23.14-15.43 0.5395 0.5325 + 0.84 + 0.90 240-120 15.43-7.72 0.5524 0.5465 + 1.80 + 1.84 120-60 7.72-3.86 0.5803 0.5758 + 1.96 + 2.02 60-30 3.86-1.92 0.6111 0.6082 + 1.90 + 2.18 30-15 1.92-0.96 0.6444 0.6442 + 1.64 + 1.57 15-0 0.96-0 0.6691 0.6702 + 4.58 + 4.30 0 0.7452 0.7452

255.

AH

Compressibility my ft2/Ton From LH

From Q V

Coefficient

From 6 H

of Consolidation cy in2/min From AV

From Au

6\T v

1.05 .0043 .006.11. - - - 0.88 .9106 .0123 0.0116 0.0156 - 0.92 .0095 .0104 0.0052 0.0052 0.0039 0.85 .0063 .0075 0.0023 0.00223 0.00233 0.91 .0040 .0044 0.00148 0.00165 0.00148 0.83 .0027 .0033 0.00123 0.00123 0.00158 o.88 .0020 .0025 0.00096 0.00096 0.0014 0.96 .0006 .0006 0.0027 0.0055 0.0028 0.93 .0011 .00116 0.0015 0.0008 0.00115 0.98 .00233 .00238 0.00082 0.00066 0.97 .0051 .0052 0.0005 0.00049 0.88 .0098 .0112 0.00019 0.00021 1.04 .017 .0167 0.00015 1.06 .048 .045

TABLE 10.5

RESULTS OF TEST NO. HPO-H0-2

Initial water

Effective stress range

Initial void

Axial Strain

[ Vol. strain

content lbs/in2 T/ft2 ratio 611/I1 % A V/V % (From change of height)

26.0% 0-15 0-.96 0.7o46 - 0.48 - 0.48 15-30 .96-1.92 0.6964 - 0.73 - 1.15 30-60 1.92-3.86 0.6841 - 1.502 - 2.09 60-120 3.86-7.72 o.6588 - 2.304 - 3.35 120-240 7.72-15.43 0.6200 - 2.67 - 3.06 240 15.43 o.5868 - -

256.

16.H/H Compressibility Compressibility Coefficient of E777 (From E. IT) my from (A V) Consolidation

ft2/Ton (ft2/Ton) cv in2/min From From AH Au

1.00 0.0050 0.005 0.63 0.0076 0.012 0.0099 0.0055 0.72 0.0078 0.0109 0.0079 0.0038 0.69 0.0061 0.0087 0.0032 0.00215 0.87 0.0035 0.0040 o.0014 0.00168

TABLE 10.6

LATERAL STRESSES AT THE END OF "SWELLING" STAGES

OF STRAIN GAUGE OEDOMETER TESTS

Test No. Ring Vertical Lateral Thickness "Swelling" . Stress _ ., Pressure

(in) (lbs/in2) (lbs/in2)

0-SG-1 1/32 9.30 31.7 O-SG-2 1/16 7.8 45.6 O-SG-3 1/8 7.8 24.0

Average 8.3 33.8

257.

TABLE 10.7

VALUES OF Re FOR OEDOMETER CONSOLIDATION

OF UNDISTURBED LONDON CLAY (Fig. 10.74)

Test No. Ring Initial Slope Final Slope Thickness 10 p.s.i.< 120 p.s.i.< (in) Cy < 500 cry' < 120

p.s.i. p.s.i.

Ro ,,) Ro

0-SG-1 1/32" 0.22 0.18 0.63 0.39 0-SG-2 1/16" 0.32 0.24 0.60 0.38 0-sG-3 1/8" 0.28 0.21 0.68 0.40

Average 0.28 0.21 0.64 0.39

Ro = Incremental coefficient of earth pressure at rest =

AT r1/ CrIT I = Poisson's ratio = K/1 + K 0 o

259

CHAPTER 11

RESULTS OF TRIAXIAL TESTS

11.1 Presentation of data

The different types of triaxial tests performed and their

procedures have been described in Chapters 8 and 9 respectively.

In this section all the basic results will be presented. Analyses

of the data will be made in subsequent sections.

Al Tests

Figs. 11.1 and 11.2 show the stress - strain and pore

water pressure relationships for the unconsolidated undrained com-

pression tests. Each specimen was set up in the triaxial apparatus

and its initial suction measured. The sample was then sheared at

a nominal strain rate of 0.0005% per minute until failure. It

will be seen that for both Oxford Circus and Ongar the specimens

failed at strains of about 2% while the pore pressures reached their

peak a little earlier. Failure was usually of the brittle nature

along one or more shear planes. For each specimen the pre-shear

effective stress g r (i.e. the initial suction) has been given.

A2 Tests

The stress - strain relationships of the two Ongar specimsns

T-HO-4 and T-110-5 which were taken to failure in undrained extension

(cell pressure constant, axial stress decreasing) are shown in

Fig. 11.3. The shape of the stress - strain curves are similar

260.

to those for compression reported above, although the change of

pore pressure does not seem to follow the same pattern. It must

be laid, however, that none of these tests were taken to large

enough strains to define the post-failure behaviour of the pore

water pressure

A3 Tests

In Figs. 11.4 and 11.5 are plotted the stress - strain

- pore pressure relationships for Test Nos. T-H0-6 and T-H0-7.

The specimens were first unloaded (cell pressure constant, axial

stress decreasing) until the vertical effective stress was approx-

imately equal to that in-situ, followed by axial compression to

failure - all under undrained conditions. Hysteresis in the stress

- strain relationships is noticeable. Failure occurred at strains

of 3 - 4% which were greater than for the Al tests. The pore

pressure - strain relationships follow the familiar pattern.

Figs. 11.6 and 11.7 show the effective stress paths for

all the tests reported above. From Fig. 11.7 in which the Ongar

test data are plotted the following points may be noted:

(a) The stress paths radiate from the isotropic line at angles

which depend on the pore pressure parameter A. The change of

direction of the individual curves also reflect the variation of

'At during a test.

(b) The estimated in-situ effective stresses are very near to

failure - as indicated by the results of the extension tests. The

261.

effective stresses of the A3 specimens at the end of the unloading

stages were not exactly equal to the in-situ stresses although the

vertical effective stresses were similar.

(c) The stress paths for the A3 tests, corresponding to load-

ing, indicate that the pore pressure parameter 'A' during loading

is higher than during unloading, resulting in lower effective stresses

in a sample when isotropic conditions are restored than at the start.

(d) The shear strength depends on the pre-shear effective

stresses but is generally independent of the type of test - as in-

dicated by the unique failure line (max (T 1 - 0-3) produced by

both Al and A3 tests and the correspondingly similar line for ex-

tension (see section 11.2).

It must be emphasised here that the study of the shear

strength of London clay does not form an important part of this

thesis although from all the tests that were taken to failure it

is possible to determine the peak strength parameters of the Ongar

clay (see section 11.2). The main purpose of the above tests was

to study the deformation of London clay due to stress increase under

undrained conditions, the details of which are presented in section

11.3.

B1 Tests

The results of all the isotropic consolidation tests are

summarised in Table 11.2. No stress paths are shown - they are

along the isotropic line - but all relevant information can be

262.

found in Table 11.2. The variation of axial and volumetric strains

with time are shown in Figs. 11.36 and 11.37.

B2 Tests

Fig. 11.8 shows the stress paths and strains for the two

specimens which were consolidated anisotropically. For these tests

the vertical total stresses were held constant while the cell-

pressure was decreased, thus giving various stress - increment ratios

during consolidation. The results are also summarised in Table

11.3. Fig. 11.39 shows the variation of axial and volumetric strains

with progress of consolidation.

Cl Tests

In Fig. 11.9 are plotted the stress paths for the three

Oxford Circus tests T-00-4, 5 and 6 in which the specimens were

first loaded axially under undrained conditions to various stress

levels and then consolidated isotropically against back pressures.

The corresponding stress paths for the Ongar tests T-HO-10, 11, 13,

17 and 19 are plotted in Fig. 11.10. The first part (undrained)

of these stress paths are, in fact, similar to those for the Al

tests, while during consolidation the stresses increase along lines

parallel to the isotropic line. Table 11.4 gives the measured

strains during consolidation and also summarises all other informa-

tion. The development of axial and volumetric strains with pro-

gress of consolidation are shown in Figs. 11.40 and 11.41.

The stress - strain relationships for undrained loading

263.

for specimens T-H0-27 and 28 are given in Figs..11.9 and 11.10.

These tests were similar to the ones described in the preceding

paragraph except that in the undrained loading stage both the

axial and lateral stresses were increased. Since the samples

were fully saturated the pore pressure parameter B = 1 and it is

possible to separate the component of the excess pore pressure due

to the principal stress difference alone. The loading stages of

these tests were followed by two stages of isotropic consolidation

for which the stress paths are shown in Fig. 11.15. Specimen

T-H0-28 was then subjected to drained compression, increasing both

the axial and lateral stresses but keeping the stress increment ratio

constant, before finally taking it to failure by drained axial com-

pression. The consolidation data are summarised in Table 11.4.

C2 Tests

Figs. 11.14 and 11.15 show the stress - strain relation-

ships for undrained compression - both axial and lateral stresses

increasing - of Test Nos. T-H0-29 and 30. These tests were similar

to T-H0-28 and 29 described above, except that the loae,ing stages

were followed by anisotropic consolidation. The stress paths are

shown in Fig. 11.16 and the consolidation data are presented in

Table 11.5.

The stress - strain relationships for the final stages of

Test Nos. T-H0-28, 29 and 30 - drained axial compression to failure

- are plotted in Fig. 11.17. These data are similar to the stress

264.

- strain relationships for the conventional drained test except that

the pre-shear stress condition was anisotropic.

D Tests

The stress path and strain data for Test Nos. T-H0-14,

15 and 16 are plotted in Figs. 11.18, 11.19 and 11.20. It was in-

tended to bring these specimens to the in-situ stresses, indicated

by crosses in the stress path plots, before subsequent loading,

but, as has been mentioned earlier, the in-situ stresses were too

near failure in extension. Therefore, the unloading stages were

stopped a little short of the in-situ stresses and the specimens

were then loaded undrained to different stress levels, followed by

isotropic consolidation as in the case of the C1 tests. Although

Au shows negative values at the end of each test the pore pressures

were, in fact, positive because of the initially high pore pressure

at the start of each test. The numerical data are presented in

Table 11.5.

E Tests

The stress - strain relationships for Test Nos. T-H0-31,

32 and 33 are shown in Figs. 11.21, 11.22 and 11.23. In these

tests the specimens were loaded under fully drained conditions in-

creasing both the axial and lateral stresses - at various stress in-

crement ratios (K' = (T3'FaNTI ) Test no. T7410-31 had three

stages of different stress increment ratios while the remaining two

specimens T-H0-32 and 33 had only one such stage before all of them

265.

were taken to failure as conventional drained tests. For each test

the ratio of the axial strain to volumetric strain (C31/(;v) are

also shown. To facilitate presentation of data 07 v has been

plotted along the abscissa and rr h, as the ordinates.

F Tests

These were isotropically consolidated drained tests in

which the specimens were first consolidated to different effetive

stresses and then sheared to failure. Fig. 11.24 shows the stress

- strain - volume change relationships for the three vertical

specimens T-HO-24, 25 and 26. Three tests were also performed on

horizontal samples, i.e. the major principal axis in the direction

bedding and the results are shown in Fig. 11.25.

11.2 Shear strength parameters

It has already been mentioned that it is not a major part

of this thesis to study the shear strength characteristics of

London clay. However, from the tests that were taken to failure

it is possible to determine the peak strength parameters of the

clay from Ongar. Fig. 11.26 is a plot of 071, - 0-31Y2 vs

1 + a31/2 at failure for all the different types of test des-

crIbed in the previous section - failure being defined as the point

of maximum stress difference. It will be seen that all the points

lie on a faikly unique straight line giving C' = 3.0 lbs/in2 and

0' = 21° as the peak strength parameters of the Ongar clay. It

266.

appears that neither the type of test, drained or undrained, nor

the manner in which failure is reached has any significant in-

fluence on the shear strength parameters, which are, therefore,

sole functions of the effective stresses at failure. This is in

agreement with the results of tests on normally consolidated

Drammen clay (Simons 1960a) and artificially over-consolidated Oslo

clay (Simons 1960b) as well as undistrubed London clay from the

Ashford Common shaft (Bishop, Webb and Lewin 1965).

It is also of interest to note that the two extension

tests give effective stress parameters which are little different

from those from compression tests, supporting the findings of

Taylor and Clough (1951) and Parry (1956).

In Fig. 11.26 are also plotted, for comparison, the

results of tests on undistrubed London clay from Ashford Common

(Bishop, Webb and Lewin 1965) and Bradwell (Skempton and La

Rochelle 1965). It is noticeable that the Ongar results lie close

to those for Bradwell but well below those for Ashford Common'.

Moreover, the marked change of slope of the envelope for Ashford

Common in passing from low to high pressure range is totally

absent from the Ongar results.

11..5 Deformation under undrained conditions

It has been shown in Chapter 3 that, for structures on over-consolidated clays, the settlement at the end of construction

267.

may be well over half the total settlement and is, therefore, of

considerable importance. For an homogeneous, isotropic, elastic

, medium this immediate settlement (( i), which takes place under

the condition of no volume change (i.e. Poisson's ratio = 0.5),

is given by,

pi-E

_ 221 i (11.3.1)

where q is the foundation pressure, B is some convenient di-

mension of the foundation, E is Young's modulus and I p is the

influence factor given by the elastic displacement theory (Terzaghi

1943). Although it is true that a real soil does not behave as an

ideal homogeneous, elastic medium it has been shown in Chapter 6

that the classical Boussinesq analysis gives stresses beneath a

foundation which are reasonably accurate, even for a non-homogeneous

or a non-linear soil medium. To calculate the settlement from

elastic theory, however, accurate determination of the elastic

modulus E is of the utmost importance.

It has been the common practice for many years to obtain

the value of E from undrained triaxial compression tests on un-

disturbed samples of clay from beneath the foundation. However,

soils being essentially non-linear in their stress - strain relation-

ships the tangent modulus as well as the secant modulus change con-

tinuously with increasing shear stress thus making it imperative to

268.

choose some criterion on which the determination of E can be

based.

The most common procedure is, of course, to obtain the

value of E (the secant modulus) corresponding to a certain level

of stress such as one-half or a third of the failure stress

(Skempton and Henkel, 1957, Ward, Samu&ls and Butler 1959, Simons

1963, Ladd 1964). Some workers have chosen various levels of

strain (1%, 2'4, 5% etc.) as the criterion (Seed and Chan 1957,

Mitchell 1964). Although neither of the above-mentioned criteria

is complete in itself, the stress based one is certainly more satis-

factory as it is the stresses which are known in a foundation pro-

blem and the strains which have to be determined.

The elastic modulus cf a soil depends not only on the

effective normal stresses but also on the shear stresses which act

on it - their effects being in opposition to one another. With

increasing normal stresses an element of soil becomes stiffer, but

increasing shear stresses cause it to deform more. This will:be

clear from the stress - strain relationships shown in Fig. 11.2

where it can be seen that specimens having higher pre-shear

effective stresses have steeper stress - strain curves giving higher

modulus of elasticity at any strain. But with increasing stress

difference the modulus decreases.

The non-linearity of stress - strain relationships of

London clay even at small strains can be clearly seen from Figs.

269.

11.27 and 11.28 where some of the stress - strain data have been

replotted to a larger scale for strains of up to 1%. This shows

that the modulus changes appreciably even at small strains. In

Fig. 11.29 are plotted the values of secant modulus calculated for

-different stress levels from all the Oxford Circus tests. To

facilitate comparison the modulus has been expressed in non-

dimensional terms by dividing it by the pre-shear effective stress

(crr') for each test. The corresponding data for the Ongar tests

are plotted in Fig. 11.30. For the A3 and D tests (stress - strain

relationships shown in Figs. 11.4, 11.5 and 11.18 - 11.20) the

compression modulus has been calculated with reference to the point

corresponding to the end of the extension stage and has been divided

by the mean effective stress ( m )o

. at that point: Although

there is considerable scatter, which is typical of London clay, the

definite trend of the variation of E with stress ratio is clearly

established. The following features of Figs. 11.29 and 11,30 are

of particular interest:

(a) There appears to be definite relationships between the

non-dimensional quantity EMT m ')o or E/( Cr')r and the effective

stress ratio during shear. This means that for any stress ratio

the secant modulus E is proportional to the pre-shear effective

stress (.0 ,)o or CT r

This canbe clearly seen from Figs.

For specimens sheared from initially isotropic condition T' = ( (7 ) r m o

270.

11.31 and 11.32 where E for compression has been plotted againtt

the pre-shear effective stress for stress ratios 1.5 and 2.0.

Similar relationships have been obtained by Ladd (1964) who carried

out an extensive investigation on four undisturbed clays. It must

be recognised, however, that although this is true for the limited

range of stresses considered, the same relationships may not hold

for higher stresses, because the stiffness of a clay increases with

effective stress only at a decreasing rate.

(b) For any value of CT r or ( Cr 1)o the secant modulus

m

decreases steadily with increasing stress level. This is also true

of the compression modulus for increasing ( 3 ' - from an initially

' an-

isotropic (0-31‹ 0'2') stress condition - although the latter

is always higher than the modulus corresponding to initially isotropic

stresses. Beyond Cr 1 - ,/(1-3 = 1, however, the recompression modulus

is similar to those obtained from conventional tests. It is also

noticeable that the extension modulus for any effective stress ratio

is not vastly different from those for compression. (Note that the

extension modulus has been plotted against 0731/0-1' and, there-

fore, to make direct comparison with the compression modulus, the

former has to be referred to the reciprocal of 445-3t/c1 1) This

latter result is in conflict with those obtained for other sites in

London which have shown higher modulus for extension than for com-

pression (Skempton and Henkel 1957, Skempton 1959).

(c) It has already been mentioned that in the conventional

271.

settlement analysis we commonly use the elastic modulii such as

shown in Fig. 11.29 and for the portion 1°- 1'/(1-31 1 in Fig.

11.30. This modulus is usually obtained from the stress strain

relationships for one half or a third of the failure stress. In

Table 11.1 is summarised the values of secant modulus for different

factors of safety (F.S. = CT1 - C13/(Cr1 3)f) taken from the

mean curves of Figs. 11.29 and 11.30, and based on the average stress

ratio at failure. It can be seen that the modulus increases by

40 to 50% in moving from F.S. = 2 to F.S. = 5.

(d) The foregoing results have important practical signifi-

cance. For settlement analysis it is common practice to perform

undrained tests on undisturbed specimens of clay at confining

pressures equal to the overburden pressure. In such cases, the

effective stresses prior to shear are quite different from those

in-situ (see Chapter 8). Ideally, the elastic modulus should be

determined for stress increases corresponding to those in the field

starting from the in-situ stresses. Although the compression parts

of the A3 and D tests reported above do not exactly satisfy this

condition, the stresses in these specimens, prior to compression,

were closerto the in-situ stresses than is the case with conventional

tests. The compression modulus shown in Fig. 11.29 - for 0-3

1/(T1

less than one, but increasing, - is, therefore, likely to be more

applicable to foundations on London clay than the modulus obtained

from conventional tests.

272.

(e) The above point can best be illustrated by taking an

hypothetical example. For the Ongar specimens, the in-situ

effective stresses are,

((rv 1) o = 32 lbs/in2

((Th')o = 76 lbs/in2

giving (CT ')o 4-(32 + 76 + 76) = 61 lbs/in2 m

Average all-round effective stresses in the specimens after

sampling, tr r, = pk = 58.5 lbs/in2 (see Chapter 8). Let the increase of total stresses on the element due to a

foundation loading be

AcTv = 30 lbs/in2

46 Cr h = 8 lbs/in2

which will set up an excess pore pressure, (see section 11.4)

. h A( A(Tv h)

. 8 + 0.5(30 - 8) = 19 lbs/in2

So the effective stresses after load application in the field,

are

273.

Cry' = 32 + (30 - 19) = 43 lbs/in2

G - = 76 4. (8 - 19) = 65 lbs/in 2

giving C73' '

. 0.67 cr cr , 65 1

From Fig. 11.30, corresponding to this value of O"3'/cr t we have

E/(0-111 1)0 = 210 which gives E = 210 x 61 = 12,800 lbs/i/Z

Let us now consider the same element in the laboratory on

which a conventional unconsolidated undrained test has been performed.

Here

(cr o ,) = crr 58.5 lbs/in2 m

After application of total stresses,

' = 58.5 + (30 - 19) = 69.5

CT = 58.5 + (8 - 19) = 47.5

Therefore, Cr , cr

6 - = 1.47

3 TIT 47.5

From Fig. 11.30, for or l,/cr,, = 1.47, we have E/(0,)0 = 90,

from which E = 90 x 58.5 = 5,300 lbs/in2

Thus for the same set of stress increases we have two

distinctly different values of E corresponding to two initial stress

conditions.

Although the above analysis points out the importance of

taking account of appropriate stress paths in determining the elastic

274.

modulus for use in settlement analysis, it must be emphasised that

the experimental data is as yet somewhat limited and final con-

clusions cannot, therefore, be drawn. More work is needed,

particularly in the following fields;

(i) Tests of the type described above should be made on

samples from various depths in London clay. It is worth remem-

bering that in a foundation problem elements of clay at different

depths are subjected to different stress paths. This makes it

difficult to assign a single value of E for the entire depth.

While by virtue of increasing effective stresses E will generally

tend to increase with depth, different levels of shear stress due

to the foundation load will have varying influence and a combina-

tion of these two factors will determine the true values of E

for the problem concerned.

(ii) For the concept of stress path to be successfully em-

ployed in settlement analysis it is essential that in-situ stresses

in London clay be correctly known. The author is not aware of any

direct measurement to date (1968) and recourse has had to be made

to indirect determinations such as those described in Chapter 8.

While there is little doubt that the latter give reasonably correct

results, there are some uncertainties which make it desirable that

they be compared with direct measurements wherever possible. There

are, of course, considerable difficulties, but perhaps model studies

may be useful in this respect.

275.

Apart from the considerations of stress path mentioned

above there are many other factors which are known to influence the

deformation modulus of clay determined from laboratory tests. Rut-

ledge (1942), Terzaghi and Peck (1948), Ladd and Lambe (1963) have

studied the effect of sampling disturbances on the deformation of

clay. Such disturbances not only give rise to effective stresses

in a sample which are different from those in a "perfect" sample but

also increase the deformability of the clay, resulting in low

modulus of elasticity determined from laboratory tests (Klohn 1965).

This has led many research workers to suggest the use of the con-

solidated undrained tests in settlement analysis (Simons 1957, 1963,

Ladd 1964). It is obvious that good sampling is an essential

prerequisite for accurate settlement prediction and in this respect

block samples are believed to be more satisfactory than ordinary

piston samples.

The rate of loading also influences the stress - strain

behaviour of clay during undrained shear. Casagrande and Wilson

(1951), Bjerrum, Simons and Torblaa (1958), Bishop and Henkel (1962),

Richardson and Whitman (1963), and Burn (1965) have shown that

this rate effect is primarily caused by the difference in the excess

pore pressure developed during the shearing process. To correspond

with field conditions it is necessary to run a test sufficiently

slowly so that uniform distribution of pore pressures throughout

the specimen is achieved.

276.

Finally, the accuracy of any prediction of settlement,

based on laboratory data, can only be checked by direct measurement

in the field, such as those undertaken, in recent years, for measure-

ment of heave of excavations (Serota and Jennings 1959, Bozuzuk 1965,

Klohn 1965, Bara and Hill 1967, Resendiz 1967).

11.4 Pore pressure parameters A and B

A general discussion on the pore pressure response of soils

to undrained loading and its importance in settlement analysis has

been made in Chapter 7. In this section numerical results will be

presented of the pore pressure parameters A and B determined

from the triaxial tests reported in Section 11.1.

For an element of clay subjected to axi-symmetric loading

with zero drainage, the increase in pore pressure due to increase in

total stresses can be expressed as, (Skempton 1954),

u= B FACr3 + A(6 - 6 0' 3) 11

(11.4.1)

where AT 3 1 and A J are increases in the major and minor

principal stresses and A and B are the pore pressure parameters.

11.4.1 Parameter B

When a specimen of clay is subjected to all-round stress

increase, (i.e. AT 1 = L10-3) equation (11.4.1) takes the form

Au.B.AT3

from which the ratio

u B 0-3

277.

(11.4.2a)

(11.4.210)

Bishop (1966a) has derived a theoretical expression for the para-

meter B in terms of the compressibility of different phases of a

soil skeleton,

B _

1 (11.4.3)

Cw - C

s 1

C - Cs

1 + n

where n denotes porosity of the skeleton

Cw denotes compressibility of water

Cs denotes compressibility of solid particles

C denotes compressibility of soil skeleton

For a saturated clay, both Cw and Cs are small compAred to the

skeleton compressibility C and, therefore, B has a value almost

equal to 1! If, on the other hand, the clay is not fully saturated,

Cw becomes comparable to C in which case B may be considerably

less than 1.

For saturated London clay Bishop (1966a), using C = 48 x 10-6 kg/cm2 Cs = 2 x 10-6 kg/cm2 and C = 3,000 x 10-b kg/cm2 obtained a value for B = 0.99.

278.

Experimental determination of the parameter B is fairly

easy. A specimen of clay is subjected to an all-round pressure in-

crease3

in the triaxial apparatus under undrained conditions

and the excess pore pressure Au is measured. The ratio

A u/60-3 then gives the value of B. This has been done for all

the triaxial tests reported in section 11.1. After determining the

initial suction of a specimen the cell pressure in each test was in-

creased, in steps such as shown in Fig. 9,15, letting the pore

pressure stabilise under each increment. The results are shown in

Fig. 9.16 and 11.33. In Fig. 9.16 the data for tests T-HO-8 and 9

are plotted in detail while those from other tests are compiled in

Fig. 11.33: There is a slight scatter, but quite definitely the

value of B equal to 1 is obtained. This not only verifies the

validity of equation (11.4.3) but also shows that the samples were.

fully saturated.

11.4.2 Parameter A

From equation (11.4.1) for a saturated clay (B = 1) we

have

Au= 60-3 1- A(6.0-1 -A0 3 ) (11.4.4)

which is the expression for the excess pore water pressure due to

Owing to extensive overlapping only a few points are shown.

279.

the combined effect of isotropic and shear stresses. The parameter

A, therefore, can be determined by measuring the pore pressure in

an undrained triaxial test during application of shear stresses.

(a) In the standard consolidated undrained triaxial test, the

minor principal stress (i.e. the cell pressure) is held constant

during shear. We then have tiCr3 0 and

(11.4,5)

(b) If, on the other hand, both the cell pressure and the

axial stress are increased during shear, the pore pressure parameter

A is given by

6u - Am A _ 0-1 - 0" 3

(11.4.6)

In the present investigation both methods have been used and as will

be demonstrated later, there is little difference in the results

obtained.

It is woth recognising at this point that the parameter

A is not a constant soil property. Lambe (1962) and Bishop and

Henkel (1962) have discussed the various factors that influence the

value of A and due consideration must be given to these before

selecting the value(s) of A to be used in design. A few of

these factors will be considered here.

280.

(a) The effect of stress histoty

The influence of stress history on the pore pressure para-

meter A has been the subject of many investigations (Henkel 1958,

Simons 1960b; Lambe 1962, Parry 1968). Depending on whether a clay

is normally-consolidated or over-consolidated the build-up of pore

pressure for the same set of total stresses may be quite different.

Normally consolidated clays usually show high A values while over-

consolidated clays indicate lower values. Also, as shown by Lambe

(1962) and Ladd (1965) anisotropic consolidation, loading/unloading

sequences and consolidation pressure may have important influence

on the value of A.

(b) The influence of stress level

From a practical point of view this is, perhaps, the most

important factor which has to be taken into account in any settle-

ment analysis. When values of A are quoted in the literature

they usually correspond to failure conditions (Bishop and Henkel

1962, Lambe 1962). In the ordinary foundation problem, however,

the shear stresses caused by the applied load are well below the

shear strength of the clay and the use of A at failure (Af) must

be misleading.

In Fig. 11.34 are plotted the values of A against the

effective stress ratio ( (T1 T/0-3 ') for Oxford Circus, obtained from

the stress - strain relationships shown in Figs. 11.1 and 11.27.

The values of Af corresponding to the maximum deviator stresses

281.

for the three specimens that were taken to failure are also shown.

In spite of the scatter it is clear that the value of A drops

from 0.6 at the beginning of a test to approximately 0.45 at failure.

Similar data for Ongar are shown in Fig. 11.35. The

A values obtained from standard tests, i.e. method (a) above - for

stress - strain relationships see Figs. 11.2 and 11.28 - are in-

dicated by open circles while the results obtained from method (b)

- stress - strain relationships shown in Figs. 11.11, 11.12, 11.14

and 11.15 - are indicated by the open triangles. The A values

corresponding to compression subsequent to unloading in the case of

the A3 and D tests, (Figs. 11.4, 11,5, 11.18 - 11.20) are plotted

as the solid dots and those for extension (cell pressure constant,

axial stress decreasing - see Figs. 11.3 - 11.5, and 11.18 - 11.20)

are also given. Af values at failure, corresponding to maximum

stress differences are indicated by the crosses. As usual there

is considerable scatter in the results, although the rate of strain

and overall range of stresses were similar in all tests. It can

be seen that all the values of A for compression lie within a

band and the discrepancies between different types of test are not

significant. Once again the parameter A decreases slowly from

an average 0.6 at low stress levels to 0.4 at failure. The A

values for extension are considerably lower than those for com-

pression but increases as failure is approached.

Perhaps the most surprising result to emerge from the

282.

above data is that the A value does not drop even more drastically

near failure. Experimental work on clays over-consolidated in the

laboratory, has often shown very low and even negative values of

A at failure (Henkel 1958, Simons 1960b, Bishop and Henkel 1962,

Lambe 1962). However, from an analysis of the Ashford Common

data (Webb 1966) the writer finds that Af'

determined from con-

solidated undrained tests, depends to a great extent on the pre-

shear effective consolidation pressure. It can be seen from Fig.

11.36, where Webb's data have been plotted, that Af becomes zero

at low effective stresses but increases as the consolidation pressure

is increased. Similar results have been obtained for other over-

consolidated clay and clay shales e.g. Kings Norton Marl (Chandler

1967) and Edmonton Shale (Sinclair and Brooker 1967).

It appears from Figs. 11.34 and 11.35 that for Oxford

Circus and Ongar the pore pressure parameter A for a typical

foundation problem will be in the range 0.5 - 0.6. For the

appropriate stress paths in the field - following the same argument

as in section 11.3 for elastic modulus - A values should be as

shown by the solid dots in Fig. 11.35. However, it is dear that

in the range of stresses normally encountered in practice, i.e.

within an adequate factor of safety, the parameter A will not

vary appreciably with stress level.

It is difficult to say precisely, from the limited data

presented above, a great deal about the possible variation of A

283.

with depth beneath a foundation. Of course, the level of shear

stresses due to the applied pressure will vary with depth and,

strictly speaking, so will the parameter A. However, if relation-

ships such as shown in Fig. 11.34 and 11.35 hold irrespective of

depth, this variation of A may not be great.

Much of what has been said above applies as well to the

relationship between A and strain. As the axial strain increases

with increasing stress level A will decrease with increasing strain

and the parameter corresponding to the strains in the field will

apply. However, analysis based on considerations of stress and

stress level seems to be more satisfactory.

(c) Other factors affecting A

In addition to stress history and stress level, certain

other factors may also influence the value of the pore pressure

parameter A measured in the laboratory. These include sampling

disturbances and test conditions.

Sampling disturbances may increase the pore pressure of

a soil specimen. This not only results in a smaller initial suction

measured for an "undisturbed" specimen than for a "perfect" sample

(Ladd and Lambe 1963) but, as suggested by Lambe (1962), can also

give rise to a more compressible soil skeleton thus causing more of

the applied shear stresses to be transmitted to the pore water.

This will give higher values of A measured in the laboratory.

The rate of shear, which has been discussed previously,

284.

also plays an important role in the development of pore pressures

under undrained conditions. Bishop, Alpan, Blight and Donald

(1960) have shown that too rapid a rate of loading results in non-

uniformity of pore pressure within a specimen and base measurement

may not, therefore, be representative of the whole specimen. (See

also Blight 1963, Crawford 1963). Bjerrum, Simons and Torblaa

(1958) and Richardson and Whitman (1963) have found that measured

pore pressuring during shear depends significantly on the rate of

loading, and Bishop and Henkel (1953) and Lo (1961) have shown that

for normally consolidated clays, creep at constant stress difference

may also be accompanied by an increase of pore pressure.

Temperature is also known to affect the pore pressure

measured during an undrained triaxial test (Ladd 1961, Paaswell

1967, Narain and Singh 1967, Mitchell and Campanella 1968).

Although this factor has not been investigated in the present work,

all the tests were run in a constant temperature room (19 ±1°C).

The influence of temperature on the measured values of A is,

therefore, unlikely to be great. However, ground temperature may,

sometimes, be different from that in the laboratory and should be

taken account of where this difference is significant. The nature

of pore fluid may also affect the pore pressure set up in a specimen

but little is known on this subject.

285.

11.5 Volume change characteristics

This section deals with the deformation of London clay

under drained conditions i.e. the axial and volumetric strains

associated with the expulsion of water from the clay. The experi-

mental work was conducted almost exclusively on the clay from Ongar,

except for the three teats of type Cl on Oxford Circus specimens.

The basic data have been presented in section 11.1.

11.5'4,1 Volumetric strains

Fig. 11.44 shows the axial and volumetric strains of the

specimens consolidated isotropcially from initially isotropic stress

conditions (BI tests). Since the effective stresses before con-

solidation varied from specimen to specimen, depending on the in-

itial suction (see Table 11.2), the results have been expressed in

terms of the non-dimensional quantity Acr1,/(cr1t)o where

(0-1 ')0 = ( (73' )o is the effective stress before consolidation and

(3-1 ' =6Q'3' is the stress increase due to consolidation

It can be seen that both axial and volumetric strains (6I

and

6 v) are proportional to A0-1 1/(0y)0 which indicates that,

within the range of stresses considered, ( 6 o- ii/( aii)o < 0.7

and 50 lbs/in2 < (CT ')0 <100 lbs/in2 ),

286.

(a) for a particular effective stress before consolidation,

the strains are directly. proportional to the stress increment and

(b) the compressibility is inversely proportional to the

effective stress before consolidation.

Thus, under isotropic stress conditions, the stress - strain

relationships are essentially linear for small stress increments.

It can also be seen from Fig. 11.44 that the ratio E,I4,v

is on

the average equal to 0.45 (range 0.40 - 0.49) which is greater than the value 1/E v = i) for an isotropic, elastic material.

It must be emphasised here that relationships such-as shown

in Fig. 11.44 are not expected to hold for high stresses and large

pressure increments because, in general, the compressibility of

soils is not inversely proportional to pressure and stress - strain

relationships are non-linear.

In Fig. 11.45 are plotted the volumetric strains obtained

from the B2 tests, in which the specimens were consolidated aniso-

tropically - keeping the axial stress constant and decreasing the

cell pressure, thus giving different stress - increment ratios, -

from initially isotropic stress conditions. The details are given

in Fig. 11.39 and Table 11.3. The open circles are the individual

points while the solid dots indicate the cumulative strains for the

first two increments in each test. Also shown are the data for the

E type tests in which the specimens were subjected to slow increases

of both axial and lateral stresses under drained conditions - from

287.

initially isotropic stresses - keeping the stress increment ratio

(K' AO-3t/ C1') constant. (First parts of test nos. T-HO-31,

32 and 33. See Figs. 11.21 - 11.23). Although the E test

results lie a little below the B2 test data it is possible 156

draw a mean curve through all the points which almost coincides

with the isotropic line, also shown in Fig. 11.45 for comparison.

It can, therefore, be said that the influence of the lateral stress

increase on the volumetric strain during consolidation or drained

compression is small. It is worth noticing, however, that the

linear relationship between stress and volumetric strain holds only

in the range LTI'/( Cri')0‹. 0.5, beyond which the curve begins

to tail off. The axial strain in these tests, of course, depends

on the stress - increment ratio (see Table 11.3) - a point which

will be discussed later.

Fig. 11.46 shows the volumetric strains obtained in the

consolidation stages of the Cl, C2 and D tests. It should be

pointed out that consolidation of these specimens followed different

stages of undrained loading, along stress paths shown in Figs. 11.10,

11.13, 11.16, 11.18 - 11.20. Consequently, the major and minor

principal effective stresses* and their ratios at the start of con-

solidation were different for all tests (see Tables 11.4 and 11.5).

In the particular case of triaxial tests described here Cril and 0-3' coincide with the vertical and horizontal stresses and the terminology can be interchanged without loss of generality.

288.

Also the lateral stress increase during consolidation was, in fact,

partly reloading. In the case of the C1 and D tests the stress

changes during consolidation were isotropic (i.e. AT 3l)

while those for the C2 tests were anisotropic. It is also of

interest to note that the stress paths followed by these specimens

are similar to those that an element of soil is likely to undergo

in the field (see Chapter 5).

It can be seen from Fig. 1146 that although there is some

scatter, there appears to be no significantly different pattern of

behaviour from the different types of test, and all the points fall

within the general scatter. Moreover, the mean curve drawn through

the points lies close to the isotropic line, replotted here from

Fig. 11.44. The most important conclusion to be drawn from these

results is that neither the stress path followed during undrained

loading prior to consolidation nor the magnitude of the lateral

effective stress has any significant influence on the volume change

that occurs during consolidation, which is, therefore, primarily a

function of the major principal effective stress.

In Fig. 11.47 the v vs LC71 (r 1/( 11)0 data have been

compiled for all the different types of test mentioned above. Once

again it can be clearly seen, in spite of the scatter, that the

volumetric strain of undisturbed London clay is primarily determined

by the major principal stress (i.e. the axial stress) and is not

significantly influenced by the lateral stresses - at least within

289.

the range of stresses considered.

Similar results have been obtained for a wide variety of

soils by many research workers. Rutledge (1947) in the Triaxial

Shear Report concluded that "curves of the major principal stress

plotted against water content for standard consolidation tests, for

consolidation under hydrostatic pressure and for total axial stresses

in slow drained tests are identical within the accuracy of the test

data". Bjerrum (1954), from tests on remoulded Zurich and Allschwyll

clays found that volume change resulting from isotropic consolidation

was approximately equal to that for standard drained compression.

De Wet (1962) observed little difference in the volume changes of

remoulded Ball clay from isotropic and one-dimensional consolidation

tests. Broms and Ratnam (1963) tested hollow specimens of Kaolin

in an independent stress control apparatus and found that the volume

change was primarily a function of the major principal stress and

approximately independent of the minor and intermediate principal

stresses. Unique relationships between volume change and axial

stress were obtained for undisturbed Leda clay, for all values of

strss ratio, by Raymond (1965). Lee and Farhoomand (1967) found

similar results for a granular material (crushed granite). There

is, therefore, ample experimental evidence to support the conclusion

arrived at above for undisturbed London clay.

There is, however, no reason to believe that such relation-

ships will hold for all types of soils. Tests on sand and a

290.

variety of remoulded clays have shown that volume change may be a

function of the average principal effective stress (Fraser 1957,

Wood 1958, Henkel 1958, Wade 1963, Sowa 1963, Roscoe et al 1963,

Schofield and Wroth 1968). Some of these studies relate conditions

at failure when the influence of shear stress becomes more important.

In the context of the work being described here, where stresses and

strains are relatively small and well below failure conditions, the

relationship between volume change and the major principal effective

stress, obtained above, seems to be approximately valid.

In order to find possible explanations for the above

phenomenon, let us now consider Fig. 11.48. Here the volume

change data have been plotted for the standard drained tests (i.e.

A(r3, = 0) for the range 0 <,,a(7 1 1/( Or1t)0‹ 0.8, ( Tit).

being the pre-shear effective stress (see Fig. 11.24). For com-

parison the isotropic consolidation line has also been shown. It

can be seen that the volumetric strain is essentially linear for the

stress range considered, except for the very small stress increments.

Two major features of Fig. 11.48 are easily observed:

( a)

For the same value of ©G 1'/( 0')o, volumetric strain

increases with the pre-shear effective stress. This means that

under this type of loading, compressibility of the clay is not in-

versely proportional to rr1IN /o - in contrast to the results of

isotropic consolidation tests.

(b) The unique relationship between volume change and major

291.

principal effective stress does not hold for standard drained com-

pression, the volume change being always lower than for isotropic

stress increase, thus contradicting the relationship obtained for

all other tests, although the discrepancy lessens with higher values

of the pre-shear effective stress.

It is well known that for an isotropic material, (*eying

the normal laws of elasticity, the volume change due to increases

of principal effective stresses ' (Y- 1 T ' Acr 2' and 31 is

given by

AO- + a tT 2° + er

Av - C V CV

(11.5.1)

3

where Ccv is the average compressibility of the material over the

range of stresses considered. This means that, for the same in-

itial condition, the volume change caused by an increase of axial

stress AU' in a standard drained test (/1 447-2

= 60-3

= 0)

should be 3 of the volume change caused by an isotropic stress

increase of the same amount (ACT 1 = AG-2 = 3

'). It is

easy to see from Fig. 11.48 that experimental results do not corres-

pond to this condition. In order to take this variation from ideal

elastic theory into account Skempton has proposed an expression for

volume change of soils in terms of the compressibility Ccv and a

(structural parameter" Sd (Skempton and Bishop 1954). For the

case of triaxial test conditions (Acr 1 > 6(T

2' = A07

3') this

is given by

V , c [ 6a- sci. &cy, a- 3' ) I V cv 3 (11.5.2)

292.

The parameter Sd is primarily a function of the dilatancy of the

soil (i.e. the volume change caused by the application of shear

stresses) and is equal to 1 for an isotropic elastic material.

For the case of isotropic consolidation ( 6 1` .Acr 30

equation (11.5.2) reduced to

A Acr CV

')

(11.5.3)

and for axial compression (A0-3 = 0)

= - Ccv(Sd . L1 cr i) 01.5.0

Therefore, for identical stress increase in the two cases, we have

_ Sd

So we

[A, v / i

value of Sd from Fig.

(11.5.5)

11.48

V 11 / V

can determine the

directly by the application of equation (11.5.5). The results are

shown in Fig. 11.49. It can be seen that Sd is a function of the

pre-shear effective stress, as well as the stress level. By extra,.

293.

polating the early parts of each curve it is possible to imagine

that Sd starts with a value of 3- but increases with increasing

stresses. With only small error, it is possible to assign average

values for each (CT1')0, for the range 0.2 < CNC"11/( 0-1' )0< 0.8

as follows,

( C1

I) Sd

56.0 lbs/ir.2 0.55

80-95 lbs/in2 0.65

120 lbs/in2 0.80

Now, to study the influence of A Cr 3' on volume change, we re-

arrange equation (11.5.2) to get

- c 1 K' Sd(1 - K') '

V cv L 1 (11.5.6)

where K' = LSC73,hscr1 . stress increment ratio.

Combining equations (11.5.3) and (11.5.6) we have, for

equal increments of AO''

tavi v - Sd

K'(1 - Sd) v (e)i v

(11.5.7)

Equation (11.5.7), then, gives the relation between the ratio of the

volumetric strains for anisotropic and isotropic stress changes

294.

during consolidation and the corresponding stress increment ratio,

for equal increases of the major principal stress and is plotted

in Fig. 11.50.

There is not enough experimental data to check the validity

of equation (11.5.7). It is possible, however, to compare the

results of the E and C2 tests, the relevant values of Sd

having been obtained from Fig. 11.49. The details are given in

Table 11.6 and a comparison of the predicted and observed values of

61( v)i are shown in Fig. 11.51. It will be seen that the

correlation is not good, the observed values being, save for one

exception, too high. On the other hand, the average of all the

observed values of e v/(6v)i is almost exactly 1 indicating

that there is little difference between the volume changes, caused

by isotropic and anisotropic consolidation (hence the closeness of

the two curves in Fig. 11.45). Much more work is needed on this

subject before the volume change characteristics of London clay can

be properly understood, but one important point of practical sig-

nificance can be made from Fig. 11.50.

It has been explained in Chapter 7 that effective stress

changes during consolidation in the field are anisotropic insofar

as there is a re-distribution of stresses due to the decrease of

Poisson's ratio of the soil from 0.5 to the fully drained value i.

For points beneath the centre of a uniform circular load - it can

be seen from Fig. 7.3 - even with as low as 0.2 and the

295.

pore pressure parameter A equal to 0.5 (see section 11.4) the

stress increment ratio during consolidation will not be less than

0.7. Over much of the depth beneath the foundation it will be

considerably greater. This means that in the field, even if

lateral stresses did influence the volume change of the soil,.

the points will lie within the area shown by the bold lined triangle

in Fig. 11.50. It can clearly be seen from this that the assump-

tion of isotropic effective stress increase during consolidation

will cause a maximum error of about 10% for volumetric strains near

the surface and perhaps not more than 5% for the entire depth.

Since it is difficult to reproduce experimental data an London clay

to this degree of accuracy any consideration of the influence of the

lateral stress on volumetric strains, in practical settlement

analyses, is unwarranted.

It is now possible to compile the compressibility data

obtained from all the different types of test.mentioned above. In

Fig. 11.52 the volumetric compressibility Ccv (defined as

ccv = 6 v/I-(7.11) is plAtted against the effective major principal

stress before consolidation ((3"10 0. As usual, the scatter is

large but the best fit curve is, perhaps, the most representative

of all the points, although it may be possible to draw slightly

different curves for one or two groups of points. It can be seen

that, within the range under consideration, Ccv is approximately

inversely proportional to (cr ,)o - a point which was self-evident

296.

in Fig. 11.47. The compressibility data plotted in Fig. 11.52,

then, are the ones that should be applicable to a field problem.

A comparative study of these results with those from the oedometer

tests will be made in Chapter 12.

11.5.2 Axial strains

So far, discussions of the deformation of London clay

under drained conditions have been concerned mainly with the volu-

metric strains. For settlement studies, however, it is the

vertical component of the deformation that is important - and this

will now be considered.

It has already been shown in Fig. 11.44 that during iso-

tropic consolidation the ratio € /(:v

(axial strain/volumetric

strain) of undisturbed London clay, in the range of stresses under

consideration, is 0.45. If the clay had behaved as an isotropic

elastic material, this ratio would have been 4. This ratio, which

will hereafter be called (2<,_ , (61 /(:-

v for all round consolida-

tion) then gives a% measure, at least qualitatively, of how far a

material deviates from isotropic* elastic behaviour.

The quantity c'( , however, is not a constant material

parameter. Among other things, it depends on the stress level

The word "isotropic" is used in the present discussion to mean either isotropic (i.e. all round) stress increase or identical elastic properties in the three orthotropic directions.

297.

(Cr 1/0-1) - as shown in Fig. 11.53, where the ratio E 1/ v o 1 L-v

for isotropic stress increases has been plotted against the

effective stress ratio before consolidation. The value of

4= 0.45 mentioned above refers, of course, to initially isotropic

stress conditions, (i.e. (Cr11/0"

3)0 = 1.0 ). It can be seen

that the higher the effective stress ratio reached after undrained

loading, the greater will be the value of OK for subsequent all

round consolidation. One reason for this may be that because part

of the lateral stress increase during consolidation in the case of

Cl and D tests is recompression (see stress paths in Fig. 11.10,

11.13, 11.18, 11.20), the clay is stronger in the lateral direction

than when the stresses are initially isotropic, thus cuasing more

of the deformation to take place vertically.

Now in a foundation problem the increase of effective

stresses during consolidation will not be isotropic. Referring

once again to Chapter 7, it can be seen that the stress increment

ratio during consolidation depends on the pore pressure parameter

A and Poisson's ratio s) '. Restricting our consideration to in-

itially isotropic stresses, Fig. 11.54 shows the ratio f i/ev

plotted against the stress increment ratio as obtained from the

C2 and E tests. The ratio corresponding to A a-3,/6crit . 0,

hereafter called p , has been obtained from the standard

drained test data plotted in Fig. 11.241

p , in fact, varies with the pre-shear effective stress. But for 46C71 1/(C1-1 1 )0‹ 0.5 an average value of 1.45 for the stress range 50 ::(0-1')0 100 lbs/in2 has been taken.

298.

It is clear that, as expected, E., 1/6v increases as the

stress increment ratio decreases. So even if the lateral stresses

have only a minor influence on the volumetric strains, (see section

11.%1), the axial strains and hence settlements are significantly

dependent on the stress increment ratio.

The solid curve in Fig. 11.54 is the best fit curve

through the observed points which intersects the ordinate

6 /(--- v =1 at Ac-3'/AcT = 0.29. At this point the lateral

strain is zero and the stress increment ratio, therefore, corres-

ponds to the incremental coefficient of earth pressure at rest,

o as defined in Chapter 10 (see section 10.4). It is interest-

ing to note that the value of Ro 0.29 corresponds almost exactly

to the value 0.28 obtained from similar stress range from the strain

gauge oedometer tests (see Fig. 10.74).

In order, now, to predict the relationship between

E 1/c v and K'( = LA cr3

') analyses will be made on the

basis of (a) isotropic elasticity and (b) anisotropic elasticity

and the respective results compared with the experimental data shown

in Fig. 10.54.

(a) Isotropic elasticity

For conditions of axial symmetry, if E and are

respectively the Young's modulus and Poisson's ratio "of the clay

for drained compression, the axial and radial strains are given by

E 2 =E3= E, E

) (11.5.8)

1 cr

299.

For small strains, the volumetric strain

v + 263 - 1 -2 (A011 + 2 /SU3') (11.5.9) E

and € 1 1 - (11.5.10) e IT (1 - 2 ')(1 + 2K')

where K' = ACr3'/40-1'

Therefore, for an isotropic material the strain ratio G

is determined by only one elastic parameter

Now for zero lateral strain (e 2 = 6 3 = 0), it can

easily be shown that

Ro = %)' 1 -

or (11.5.11)

where Ro = ACT 3 A (71' corresponding to one-dimensional strain.

Combining the last two equations we have

E 1 1 + Ro - 2RoK'

(11.5.12)

(1 - Ko)(1 + 2K')

Equation (11.5.12) then gives the relation between the strain ratio

300.

and the stress increment ratio during consolidation. For the

range of stresses under consideration, Ro 0.29 and we get the

curve shown by the chain-dotted line in Fig. 10.54. It can be

seen that the curve starts at Elk. v = - for Kt = 1 - as all isotropic materials must - and passes through the experimental

point corresponding to Ro 0.29. Lambe (1964) and Seed (1965)

have suggested the use of equation (11.5.12) to predict the strati

ratio for consolidation in the field but it can be seen that over

much of the range the correlation is not good indicating that the

assumptions of isotropy are not strictly valid.

(b) Anisotropic elasticity

It will now be assumed that the clay behaves as a cross-

anisotropic material (Barden 1963) during drained compression,

having different elastic properties in the vertical and horizontal

directions. Five elastic parameters are now required to define

the behaviour of the material - these are:

E:1 = the elastic modulus in the vertical direction

E3' = the elastic modulus in the horizontal direction

and three values of the Poisson's ratio

= effect of one horizontal strain on the other

horizontal strain

= effect of horizontal strain on vertical strain

= effect of vertical strain on horizontal strain

For a triaxial specimen with its axis vertical, the strains are

301.

given by

2 2

-.> ) cr 1 - (a)) I -

L E l , E3,

31) ( Lo- ) )(11.5.13)

.5 ) C- 2 = 6 3 - Ac13: - )1 (Acr.4 ') - --2-• ( &c7"1 ') (13»

E 3, E 3'

- E3'

where Acr1 and ACr3' are increases of axial and radial

effective stresses. For small strains, the volumetric strain

1 - 2

v G 1 +E 2€3 - 3 (AO '

11)

E ,

+ 2(1 - J - 0- ) 3 1 2 E 3

(11.5.14)

So the strain ratio

1 - 2n .N)2K

(11.5.15) e (1 -

3) + 2n(1 ,)1 - .) 2)

where K' = 66 31/210- 1 ' and n = E1 l/E3'

Although there are five elastic constants we need only know three

parameters in the form of 3, n 2 and n(1 - to define

the strain ratio of a cross anisotropic material.

Now under Ko conditions

= 1 and K' = Ro and substituting this in

302.

equation (11.5.15) we get

.;"). R _

3 or n(1 - J ) =

n(1 - J1) 1 1‘*0

Combining equations (11.5.15) and (11.5.16)

(11.5.16)

€ 1 1 - 2n• 2K'

(11.5.17) E. v 1 + 2 .s.)

3 (m/Ro - 1) - 2n 2

Kt

(For the isotropic ase we can recover the classical expression,

1/€ v = i for K' = 1, from equation (11.5.17) by putting

n = 1, -) 2 = 3 = J and Ro = ")/1 ). The parameters

N)3

and n.>2 can be determined from two tests with different

values of K' and measuring E 1/E v in each case. The most

convenient values of K' to choose are 0 and 1.0 i.e. the

standard drained test and the isotropic consolidation test.

For the present work, it has already been shown (Fig.

11.54) that

when = 0, C i/E v = = 1.45

K' = 1, E v = = 0.45 ) (11.5.18)

and it = 0.29

Solving equation (11.5.17) for the values given above

we get

303.

3 o.16 and

2 0.18

For the particular problem under consideration we then have the

general expression

6 1 1 - 0.36K' (11.5.19)

v 1 - 0.32(3.45K 1 - 1) - 0.36K'

This equation has been plotted in Fig. 11.54 as the broken curve

to give the relation between E 1/E, v and K' for the range of

stresses under consideration.

It will be seen that a far better correlation with ob-

served data is obtained by assuming anisotropic behaviour than

by assuming isotropy.

It must be stressed, howevet, that relationships such as

given by equation (11.5.19) are very much stress dependent and should

be determined for a particular problem with due regard to the range

of stresses that are going to be applied.

The influence of stress level on 0( has already been

shown in Fig. 11.53. The variation of i3 with stress level is

even more marked as indicated in Fig. 11.55, where the ratio

dE 1/dEv has been plotted against (071'/0-31)0 from the final

stages of the C2 and E tests (see Figs. 11.17 and 11.21 -

11.23). Although ig is not very sensitive in the range

1.0 <(Cri70-300‹ 1.5 it increases very rapidly for further

304.

increase of the stress ratio and should approach infinity at failure.

In this region elastic theory does not apply and analyses such as

described above will not produce any meaningful result. For

Factors of Safety greater than 2, however,(for Ongar this means

(C71t/C5- t) o less than approximately 1.6) elastic theory can be 3

assumed to be at least approximately valid and values of O( and

13

obtained from figures such as 11.53 and 11.57 coupled with

the appropriate value of Ro should be sufficient to determine the

relationship between 6 1/(-7,v

and K' on the basis of anisotropic

elasticity. From practical problems in London clay, because of

very high in-situ horizontal stresses, the increase of stresses due

to a foundation load will keep the values of ( 7 t/crh')o

sufficiently low and for a greater part of the depth even less than

1 and in such cases relationships shown in Fig. 11.54 will probably

not be too much in error.

The practical implications of the foregoing results are

of the greatest importance. As shown previously the lateral

stresses will not have much influence on the volumetric strain of

an element of soil beneath a foundation. But the vertical strain,

and hence the settlement, will only be a percentage of the total

volumetric strain depending on the stress increment ratio. Thu4

Although the analyses have been done in terms of major and minor principal effective stresses they should be applicable to vertical and horizontal stress increments in the field.

E: 1 and E 3,

in the vertical direction ( Ey, E1') can be determined from the

(and thus n) independently. The elastic modulus

305.

for example, if A 0- 3f/per1, = 0.7 we have, from Fig. 10.54,

el/e, = 0.6. But in a conventional settlement analysis, use

of oedometer test data implies one-dimensional strain i.e.

E:A v = 1.0. So if the volumetric strains are equal, direct use

of the oedometer compressibility will over-estimate the vertical

strain of the element by more than 60%. (This point will be dis-

cussed in greater detail in Chapter 12).

11.5.3 Elastic parameters of London clay

So far we have determined three elastic quantities J 3' ,

n .) 2 and n(1 - J 1) = N) 3/Ro which are sufficient to define the

strain - ratio fo an anisotropic material (equation 11.5.15). In

order to obtain all the five parameters it is necessary to determine

standard drained test data (Fig. 11.24), by applying equation

(11.5.13a) in which is substituted AO-3, = 0. The results are

plotted in Fig. 11.56 in the form of secant modulus vs effective

stress ratio. Since,the modulus increases with consolidation

pressure the data have been plotted in terms of the non-dimensional

quantity E2 v '/crc As in the case with the undrained modulus

(see section 11.3) Ev decreases with increasing shear stress,

indicated by the effective stress ratio, and the best fitted linear

relationship has been obtained by the method of least squares.

- 2 a.

v ver

(11.5.20)

MI6!!

To determine the horizontal modulus (Eht = El31)

standard drained tests were performed on three specimens with their

axes parallel to the direction of bedding at effective consolidation

pressures of 60, 100 and 100 lbs/in2 (Fig. 11.25). During the

consolidation stages of test nos. T-H0-35 and T-H0-36 both the

axial and volumetric strains were measured as in the C1 tests.

The results are shown at the bottom of Table 11.2. It can be seen

that the average ratio e i/E, for the two tests is 0.40.

Now the axial strain €1 for a horizontal sample is,

in fact, equivalent to the lateral strain (C3) of a vertically

oriented sample. This means, from equation (11.5.14) that

€ 1 i

v Hor _ C v .1 ver

Using (61/E ver = 0.45 we have

€ 1 2(1 - 0.45) = 0.28 v Hor

The observed (E1/(.1v)Hor does not correspond well with the ratio

calculated on the basis of the tests on vertical specimens. How-

ever, more tests on horizontal samples are needed before this dis-

crepancy can be properly examined, although one possible reason may

be that the two horizontal specimens, for which data are available,

had their stress/strain modulii very similar to those for the

307.

vertical specimens (compare Figs. 11.24 and 11.25) indicating that

their degree of anisotropy was, perhaps, not great.

The secant modulus for the horizontal specimens (expressed

in non-dimensional terms by dividing it by the pre-shear consolida=te

tion pressure 4-0') has been plotted, as before, against stress

level lir11/cr3, in Fig. 11.57. Although the scatter is large the

best fit line has been obtained, as for the vertical samples, by

the method of least square. Fig. 11.58 shows the comparison

between the horizontal and vertical modulii. It can be seen that

E is greater than El 1 but only by a small amount, the ratio

n = E:vVEEh' being in the range 0.85 - 0.95. This confirms that

the degree of anisotropy, insofar as the elastic modulii are con-

cerned, is small. Of course, to have a clearer idea of the overall

anisotropy one must consider the difference between the three

Poisson's ratios as well.

With the average value of n = 0.87, for the range

<A(711,/(cri t)0A( 0.5, we can..now calculate --)1 and *> 2 in

addition to •••.) 3 which has already been determined. The results

are summarised in Table 11.7. For comparison, the Poissonts ratio

calculated on the assumption of isotropic elasticity

Of = R0/1 + R0) is also shown.

308.

The above parameters can be used to predict the com-

pressibility of undisturbed London clay for drained compression.

From equation (11.5.14)

E'

1 - 2 3 (60-11)+2(1- 2 ) El 3

Substituting &cr ' = Acr3 for isotropic consolidation

1 - 2,)3

2(1 - 1 - )2)

- E 1

E. 3

E (11.5.21)

Compressibility 12cv is defined as

v

ccv f(Cr/)o - (11

The two equations can be combined to give

C _ 1 - 2.)

3 2(1 - 1 2 )

Cv p 3'

(11.5.22)

(11.5.23)

Taking the values of E 1 = El v f = 52((11)0 and

E 3 = E = 60( 1t)o from Fig. 11.56, for (01 1/0-31)0

= 1.0 and the elastic parameters listed in Table 11.7, we get

c _ 0.027 cv (0-1,)0 (11.5.24)

309.

Equation (11.5.24) has been plotted in Fig. 11.59 as the

predicted relationship between volumetric compressibility and

effective stress which shows extremely good agreement with the

average experimental curve, replotted here from Fig. 11.52. How-

ever, this is not a completely independent check because one of the

five elastic parameters has been determined for the all round con-

solidation tests.

Although the above data suggest that the deformation of

London clay under drained conditions can be approximately analysed

in terms of anisotropic elasticity a few inconsistencies still re-

main to be investigated. Most important is the apparent insensi-

tivity of the volumetric compressibility to lateral stress incre-

ment which cannot be explained in terms of elastic theory.

11.6 Rate of consolidation

The time vs volumetric strain and time vs axial strain

data for the consolidation stages of the triaxial tests have been

presented in Figs. 11.37 - 11.43. All the tests were terminated

when there was no more measurable volume change or axial deformation.

Secondary consolidation was not studied in detail.

It can be seen from the isotropic consolidation data

(Figs. 11.37 and 11.38) that the end of primary consolidation is

reached at about the same time with respect to both axial and

volumetric strains. The two Cv

values indicated in Table 11.2

310.

for each consolidation stage refer to the two strains and have been

calculated on the basis of t50. As would be expected, the

difference between them is not great although in six out of eight

cases the Cv

values for axial strain are slightly greater. Figs.

11.60 - 11.63 show the comparisons between the observed rate of

consolidation and the Terzaghi theoretical curves fitted at t50.

It can be seen that the agreement, on the whole, is satisfactory

for degrees of consolidation of up to 70%, although at early stages

there is a tendency for the observed rate to be a little faster.

Towards the end of consolidation, however, the observed rate is

always slower than theoretical, indicating the influence of second-

ary and creep effects as in the case of the oedometer tests (see

Chapter 10, section 10.2). The curves for1 and e v, which

should ideally be identical, show small differences, reflecting

the differences in Cv, although the maximum discrepancy is no

more than 5% at any time.

Similar results are obtained for the Cl and D tests

in which the specimens were subjected to isotropic consolidation

after different stages of undrained loading. The Cv

values are

given in Tables 11.4 and 11.5 and the comparisons between the ob-

served rate of consolidation and the Terzaghi theoretical curves,

fitted at t50, are shown in Figs. 11.64 - 11.69. Here the

discrepancy between the rate of volume change and the rate of

axial deformation is somewhat greater than for the BI tests -

311.

about 10%. Perhaps some difference is to be expected because the

samples in these tests were subjected to shear stresses under which

creep and secondary effects influence the axial and volumetric

strains in different degrees.

The results of anisotropic consolidation (B2 and C2)

tests are plotted in Figs. 10.70 - 10.72, and the Cv values are

summarised in Tables 11.3 and 11.5. Very good agreement is

obtained between the rate of volume change and the theoretical

Terzaghi curve fitted at t50. No comparisons are shown for the

rate of axial deformation, because the rate at which the lateral

pressures were decreased during consolidation (see Chilpter 9,

section 9'4..2) did not exactly conform to the rate of dissipation of

the average pore water pressure. Although this did not affect the

volume change or the total axial strain, the rate of axial deforma-

tion was considerably affected (see consolidation curves in Figs.

11.39 and 11.43).

The permeability data shown in Tables 11.2 - 11.5 have

been calculated from the Cv •values obtained for volume change,

using the expression

C _ v -N(

C w cv

where u w = unit weight of water

C cv = coefficient of volume compressibility and

(11.6.1)

312.

k = coefficient of permeability.

Fig. 11.75 shows the Cv values, obtained for volume

change from the different types of test, plotted against the average

vertical effective stress during consolidation while similar data

for the coefficient of permeability are shown in Fig. 11.76. It

is clear that the scatter is large and no consistent picture emerges

as to the influence of the type of test on Cv and k. Although

the C2 tests seem to indicate slightly higher values, the number

of tests is too small to allow one to be conclusive about this.

Notwithstanding the scatter, however, it is possible to see a

definite decrease of the coefficient of permeability with pressure

(Fig. 11.76), while the mean curve lies fairly close to that ob-

tained from oedometer tests. In the case of Cv, on the other

hand, the data show only a small decrease with pressure, but this

agrees well with the range of Cv values obtained from oedometer

tests - also plotted in Fig. 11.75 for comparison.

The two isotropic consolidation tests on horizontal

specimens indicate that both Cv and it are considerably higher

for horizontal drainage than for vertical drainage (see Table 11.2),

although the data are clearly inadequate to allow any definite com-

parisons to be made. However, similar results have been obtained

from oedometer tests by Ward et al (1959) and Agarwal (1967).

It has been mentioned earlier that there is good agree-

ment between the observed rates of settlement and the predictions

313.

based on the Terzaghi theory fitted at t50. This is more clearly

seen from Figs. 11.73 and 11.74 where the theoretical and observed

rates of consolidation have been plotted against the Terzaghi

time factor Tv for the different types of test reported above.

Thbre can be no doubt about the validity of the Terzaghi theory

in predicting the average rate of consolidation of undisturbed

London clay, at least for degrees of consolidation of up to 70%T

In this respect, too, the behaviour of London clay under triaxial

test conditions is very similar to that under oedometer conditions

(see Chapter 10, section 10.2). It can be said, therefore, that

for one-dimensional drainage, both the volumetric compressibility

and the coefficient of consolidation of undisturbed London clay

are essentially independent of whether the specimen is allowed to

deform in all directions or is restrained laterally. Of course,

the influence of stress level on these quantities will still have

to be considered in any;;settlement analysis.

It is worth recalling, at this point, that field pre-

dictions of the rate of settlement, based on laboratory data, have

ogten been found to be slower than actually measured (see Chapter

3). There are many reasons for this, the most important being

It has been shown in Chapter 10 that there is greater discrepancy between the observed rate of dissipation of the maxi-mum pore water pressure, as opposed to the average degree of con-solidation, and the Terzaghi theory. Pore pressures were not measured in the triaxial tests, since to do this, much longer testing times would have been necessary.

314.

that in addition to vertical drainage, consolidation in the field

takes place horizontally. This three-dimensional flow of water

causes settlement to progress much faster than predicted on the

assumption of one-dimensional flow (see Chapter 13). The general

theory of three-dimensional consolidation in a porous elastic medium

was formulated by Biot (1941). Biot (1955) extended this work to

cover the more general case of the anisotropic medium. The

solution of these mathematical problems are, however, much too

complex for use in engineering analysis and only a few special cases

have so far been solved. Gibson and Lumb (1953) used numerical

methods to solve the Terzaghi equation for combined radial and

vertical flow. Analytical solutions have been obtained for the

consolidation of a semi-infinite, isotropic elastic medium, sub-

jected to uniform boundary loads (Gibson and McNamee 1957, 1963,

de Josselyn de Jong 1957, McNamee and Gibson 1960). These

analyses show that the problem of aonsolidation in the field is

Intimately linked with the problem of stress distribution and a

knowledge of both the Young's modulus and Poisson's ratio are re-

quired for its evaluation. However, as has been shown in the

previous section, the properties of London clay, like most other

clays, are anisotropic. Analytical solutions of the consolidation

problems for such media are not yet available.

There is also considerable difficulty in determining

from laboratory tests the true coefficient of consolidation that

315.

should be applicable in the field. Rowe (1968) made an extensive

study of the calculated and observed rates of settlement of em-

bankment:dams, with particular reference to sand drains, and found

that the Cv

values determined from field rates of settlement

were generally many times higher than those obtained from con-

ventional oedometer tests on 3 in diameter samples. This he

attributed to the presence of thin layers of more pervious materials

which, if spaced closely, may alter the field value of Cv

con-

siderably (see also RO--7 1959, 1964). With usual structural

foundations on London clay, however, this problem is not likely to

be as great because, in the absence of sand drains, horizontal

drainage may only be accentuated insofar as the presence of lamina-

tions causes horizontal permeability to be greater than vertical

(Ward et al 1959).

It may be remarked, finally, that the presence of random

fissures in London clay may also influence the rate of consolidation

in the field. But, of this, very little is known.

316.

TABLE 11.1

SECANT MODULUS AT DIFFERENT FACTORS OF SAFETY

Location

Oxford Circus

Ongar (a) Compression) (b) Extension )

(From Fig.

Initial Stress Condition

0--1

= 03

a-- =

11

11.30)

Secant Modulus F.S. = 5 3

145 132

105 90 122 100

EACTp) 2

115

78 8o

1

72

52 25

TABLE 11.2

RESULTS OF ISOTROPIC

Test No.

Drainage Condition Single (S) or Double (D)

(a- t) 1 o

= ( (7 ')o 3 lbs/in2

Acrit

= La-3 lbs/in2

Acril Axial Strain

oz 1'

Vol. Strain

€ 26 „, (crY) „..,€

Vertical I aaalaa

T-110-8 s 58.5 38.0 0.650 0.66 1.58 T-Ho-9 s 56.0 31.0 0.554 0.59 1.36 T-H0-21/1 D 70.2 29.8 0.425 0.48 0.98 T-H0-21/2 D 100.0 30.0 0.300 0.37 0.78 T-HO-24/1 D 66.0 29.0 0.439 0.44 1.02 T-HO-26/1 D 58.0 12.0 0.207 0.200 0.50 T-HO-26/2 D 70.0 50.0 0.714 0.78 1.78 T-HO-33/1 D 59.0 41.0 0.695 0.87 1.79

Horizontal aaalaa

T-H0-35/1 D 62.0 28.0 0.452 0.60 1.40 T-110-36/1 D 66.5 33.5 0.504 0.62 1.63

_____

CONSOLIDATION (B1) TESTS

E 1 ccl E 1 _

acv ' v ,

c Ir

(From 6 v)

in2/min x 10-3

cv (From E 1)

in2/min x 10-

k .

-.!

cms/sec 10-10

I, ( IN Cr

) 1

0

in2/lb x 10-4

60- , 1

in2/lb x 10-4

0.418 1.74 446 1.32 1.80 11.46 0.435 1.90 4.38 1.20 1.80 12.06 0.490 1.61 3.29 1.40 1.45 7.30 0.473 1.23 2.60 1.28 1.17 4.65 0.432 1.52 3.52 1.89 2.15 11.58 0.400 1.67 4.18 2.68 2.21 14.14 0.438 1.56 3.56 2.22 2.45 13.34 0.483 2.10 4.35 1.09 1.45 9.65

0.428 2.14 5.00 6.92 2.93 22.40 0.380 1.87 4.85 4.23 3.46 25.67

317.

TABLE 11.3

RESULTS OF ANISOTROPIC

Test No.

Drainage Single (S) or Double

( (r.. 0 1 ( cr ') 3 0 cr 1 A ' Acr I ' ' AT 5 ' no- I ,

60- t 1 ( cr 1) 1

(D) lbs/in2 lbs/in2 lbs/in2 lbs/in2

T-HO-20/1 D 84.o 84.o 26.o 22.0 0.85 0.31 T-HO-20/2 D 110.0 106.0 20.0 13.0 0.65 0.18

cum 84.o 84.o 46.o. 35.o 0.76 0.55 T-110-22/1 D 58.0 58.0 27.0 17.0 0.63 0.47 T-HO-22/2 D 85.0 75.0 25.0 10.0 0.40 0.29 T-HO-22/3 D 110.0 85.0 30.0 6.o 0.20 0.27

cum(1&2) 58.0 58.0 52.0 27.0 0.52 0.90 cum(1,2,

3) 58.0 58.0 82.0 33.0 .0.42 1.41

318.

CONSOLIDATION (B2) TESTS

Axial Strain 61%

Vol. Strain C. v%

E 1 Cv ccl 1161

=

c E v =

k c

V (d am 1 ) ( n0- Z ) 1 1

in2/min x 10-3

in2/lb x 10-4

in2/1p x 10-9-

cm/sec x 10-10

0.47 0.94 0.50 1.98 1.81 3.62 10.92 0.33 o.48 0.69 1.58 1.65 2.4o 5.78 o.8o 1.42 0.54 - 1.74 3.19 - 0.74 1.20 0.63 3.10 2.78 4.44 20.98 0.58 0.70 0.83 1.84 2.32 2.80 7.85 0.68 0.64 1.06 2.3o 2.27 2.13 7.47 1.33 1.90 0.70 2.56 3.65 - 2.01 2.54 0.79 - 2.45 3.10 -

.1.

TABLE 11.4

RESULTS OF ISOTROPIC

Test No.

Drainage Single (S) or Double

(Tit), (0-31)0 ( orif)0 Acy =

air , 3

i16r1 f Axial

Strain E l%

Vol. Strain

E v% ( 63,)0

(<T1 ' )o (D) lbs/in2 lbs/in2 lbs/in2

ilasaE

T-110-10 S 77.7 46.o 1.69 29.0 0.373 0.59 1.00 T-H0-11 S 70.3 43.5 1.62 24.5 0.348 0.70 1.16 T-HO-13 S 67.7 53.3 1.27 26.5 0.497 0.62 1.06

T-H0-17 s 59.o 46.8 1.26 27.2 0.461 0.61 1.05

T-Ho-19 s 49.6 43.6 1.14 36.4 0.734 0.71 1.60

T-HO-27/a D 60.3 28.0 2.15 27.0 0.448 0.76 1.20 T-H0-27/b D 87.3 55.0 1.59 30.0 0.344 0.67 1.09 T-HO-28/a D 76.8 52.0 1.48 26.0 0.339 0.71 1.32 T-H0-28/b D 102.8 78.0 1.32 25.0 0.243 0.34 0.71

Oxford Circus

T-OC-4 s 108.9 31.0 3.51 35.5 0.325 0.47 0.78

T-OC-5 S 121.9 41.5 2.93 28.7 0.235 0.38 o.6o

T-0C-6 S 157.5 70.5 2.23 47.5 0.301 o.54 0.82

CONSOLIDATION: Cl TESTS

ccl 6 0 _

, c

E• v cv (From C-,- i

in2/min x 10-3

cv (From € v)

in2/min x 10-3 cms/seox 10 1°

€ v Ao- I 1 in2/lb x 10-44. - 2

=

AT , 1

in2ilb x 10-4

0.590 2.03 3.45 1.80 2.76 14.57 0.628 2.86 4.69 2.22 3.09 22.17 0.592 2.36 3.98 1.11 0.96 5.85 0.581 2.24 3.86 4.22 3.17 18.72 0.444 1.95 4.39 1.27 1.01 6.78 0.633 2.81 4.44 1.95 2.64 17.93 0.615 2.23 3.63 1.33 1.71 9.43 0.538 2.73 5.08 2.89 3.21 24.94 0.478 1.36 2.84 1.85 2.31 10.04

0.606 1.34 2.21 - - - 0.633 1.32 2.09 - - - 0.653 1.14 1.74 - - -

319.

TABLE 11.5

RESULTS OF CONSOLIDATION

Test No.

Drainage Single (S) or Double (D)

( 0-11)0

lbs/in2

( Cr3 ' )o

lbs/in2

( cri T)Aci I

lbs/in2

4 310

lbs/in2

A a-.1

4 v-i '

( Cr3') o A 0-1 f ( Cr1 ' )o

C2 tests

T-H0-29/a D 73.0 40.0 1.825 22.0 16.0 0,-73 0.301 T-H0-29/b D 95.0 56.0 1.696 30.0 15.0 0.50 0.316 cum 73.0 40.0 1.825 52.0 31.0 0.596 0.712 T-HO-30/a D 69.3 48.0 1.444 22.0 15.7 0.71 0.318 T-H0-30/b D 91.3 63.7 1.432 20.0 11.0 0.50 0.219 cum 69.3 48.0 1.444 42.0 26.7 0.636 0.606

D tests

T-HO-14 52.8 35.0 1.509 22.0 22.0 1.00 0.417 T-HO-15 71.9 47.0 1.530 20.0 20.0 1.00 0.278 T-H0-16 59.5 40.2 1.480 24.8 24.8 1.00 0.417

(C2 AND D) TESTS

Axial Strain El%

Vol. Strain

C v%

C1

ccl € 1 _

ccv = c v

Cv (From E) v

in2/min x 10-3

k

cm/sec x 10-10

v A °-1 t

in2/lb x 10-4

6c71' in2/lb x 10-4

0.68 0.78 0.873 3.09 3.54 3.3o 17.87 1.17 1.06 1.104 3.90 3.53 2.11 11.39 1.85 1.84 1.005 3.56 3.54 0.57 0.89 0.640 2.59 4.05 3.70 22.93 0.46 0.64 0.710 2.3o 3.25 2.41 11.98 1.03 .1.53 0.673 2.45 3.64

0.68 1.20 0.562 3.07 5.45 1.34 11.70 0.30 0.51 0.588 1.50 2.60 2.00 7.96 0.61 0.93 0.656 2.46 3.75 1.46 8.38

320.

TABLE 11.6

Test no.( (r1 ?)o

Au' .31

(Fig. 11.49

po---I

I K' = ----'- d d 6- 1' ((T71,)o

E Tests

T-H0-31/1 70.0 0.68 0.60 0.336 31/2 93.5 0.43 0.62 0.305 31/3 122.0 0.25 o.8o 0.412

cum 1 & 2 70 0.54 0.60 0.74 T-H0-32/1 56.0 0.27 0.60 0.70o T-110-33/1 100.0 0.42 0.70 0.600

Ca Tests

T-1162201 84.o 0.85 o.65 0.310 2o/2 110 0.65 0.70 0.18

cum 1 & 2 84.o 0.63 0.68 0.55 T-H0-22/1 58 0.63 0.58 0.466

22/2 85 0.4o 0.65 0.294 22/3 110 0.20 0.75 0.273

321.

( e ). % v 3. (Fig. 11.44)

G v % Observed

€1( E ) . IT 3. Observed

eic e ). 1r I Predicted (Fig. 11.50)

0.83 0.78 0.94 0.88

0.76 0.79 1.04 0.79

1.02 0.95 0.93 0.85 1.57

1.72 1.10 0.64 0.71 1.48 1.30 o.88 0.83

0.76 0.94 1.24 0.96

0.44 0.48 1.09 0.90

1.38 1.42 1.03 0.89

1.15 1.20 1.04 0.85 0.73 0.70 0.96 0.79 0.67 0.64 0.96 o.8

322.

TABLE 11.7

ELASTIC PARAMETERS OF UNDISTURBED LONDON CLAY

1 -

Assumption Ro la= '' 1 ...> 2 N.) 3 Ek '

Anisotropy 0.29 0.87 0.37 0.21 0.16

Isotropy 0.29 1.0 ..,) 1 = ..)2 = „:),.5 = 0.21

323.

CHAPTER 12

A COMPARATIVE STUDY OF THE TRIAXIAL AND OEDOMETER TEST DATA

In Chapter 11 comparisons have been made between some

oedometer and triaxial test results. In this chapter a more

detailed study will be made of the various quantities that can be

determined from both triaxial and oedometer tests, in relation to

their practical applications.

12.1 Volumetric compressibility

For the oedometer test, the coefficient of volume com-

pressibility my is defined as

1 d e my = 1 + eo

(12.1.1)

where eo is the void ratio at pressure po, and Ae is the

change of void ratio for increase of pressure from po to po + L p.

The relationships between my and the effective vertical

stress for the different oedometer tests have been presented in

Chapter 10 - the data are summarised in Fig. 10.54. The various

factors that affect the compressibility of clays have also been

discussed in Chapter 10.

The volumetric compressibility for triaxial compression,

C cv, defined as

324.

C — . 1

(12.1.2)

cv V CT1

(where AT1' is the change of the major principal stress from

(a-11)o

to (Cr!)o + tscr ,) has been plotted against (Cr.1 1)o 1

in Fig. 11.52. It should be noted that in the range of stresses

under consideration the effective vertical stress is the major

principal stress for both oedometer and triaxial tests.

The two sets of results have been replotted in Fig. 12.1

for comparison. The oedometer data consist of four compressibility

vs vertical effective stress relationships for the following tests

which have been reported in detail in Chapter 10,

(a) Standard oedometer test - first loading

(b) Standard oedometer test - second loading

(c) Tests in the high pressure (hydraulic) oedometer

(d) Controlled rate of strain test.

(The data do not include the results of the standard

oedometer tests in which the specimens were allowed to rest for 90

days at the in-situ effective overburden pressure - they are shown

in Figs. 10.21 and 10.22).

The triaxial test data, on the other hand, are represented

by a single Ccv vs co-1 1)0 curve for all the consolidation tests

reported in Chapter 11.

The first point to note from Fig. 12.1 is that the com-

pressibilities determined from triaxial tests lie within the range

325.

of those obtained from oedometer tests. The triaxail compressibi-

lities fall in between the first and second loading standard tests,

the former always giving higher values. Over most of the stress

range, the hydraulic oedometer data lie close to the triaxial

results while the controlled rate of strain test gives compressi- .

bilities- which are smaller than the triaxial values at low stresses

but greater at high stresses. It is difficult to say how far these

differences are significant, because the stress conditions in all

the oedometer tests - except, perhaps, the second loading standard

tests - are similar, yet there appears to be a marked difference

in the compressibilities. That initial conditions in the oedo-

meter affect the compressibility has been discussed in detail in

Chapter 10. It is possible, however, that the controlled rate of

strain test under-estimates the compressibility at low stresses -

giving smaller values than even the second loading standard tests

(see Fig. 10.54).

Nevertheless, the above study reveals that the difference

in compressibility obtained from triaxial and oedometer tests is

only small - the first loading oedometer test giving slightly higher

values. But it has been described already in Chapter 10 that these

latter tests were affected by initial swelling and the close corres-

pondence between the triaxial and hydraulic oedometer data suggests

that a properly conducted oedometer test, where the initial swelling

is really prevented, gives a reliable estimate of the volumetric

326.

compressibility of undistrubed London clay. The comparison shown

above also supports the general conclusion, arrived at in the pre-

vious chapter, that the lateral stress has only a small effect on

the compressibility. It should be remembered that the triaxial

tests were conducted, following many different stress paths, prior

to and during consolidation, while deformation was always one-

dimensional in the oedometer, the corresponding stress paths being

as shown in Fig. 10.74. Yet the data plotted in Fig. 12.1 do not

suggest any strong influence of stress path on the volumetric com-

pressibility.

The effect of small pressure increments, following long

periods of rest was not studied in the triaxial tests, It has

been found from the oedometer tests, however, that, after a rest

period of 90 days, an increase of effective stress of less than

10% does not cause any appreciable volume change. Applying this

to a problem in the field, this means that a "threshold" value of

at least 10% of the in-situ vertical effective stress must be ex-

ceeded before any significant volume change would occur.

12.2 Axial strain

It follows from the above discussion that volumetric com-

pressibility of undisturbed London clay can be correctly determined

from oddometer tests, (provided the initial swelling is truly pre-

vented) as well as from triaxial consolidation tests. Similar

327.

results will be obtained by the two methods of testing because the

lateral stresses have no significant influence on the compressibility.

On the other hand - it has been shown in Chapter 11 - the lateral

stress increment does have an important influence on the vertical

strain associated with the process of consolidation. For the

same increase of the vertical effective stress, two specimens will

undergo different amounts of settlement depending upon the magnitude

of the lateral stress increment. This can best be seen with re-

ference to the stress paths shown in Fig. 12.2.

Point A represents the in-situ vertical and horizontal

effective stresses Of the Ongar specimens, their magnitudes being

Crv' = 32.0 lbs/in

2

CS h = 76.0 lbs/in2

After sampling the total stresses are reduced to zero and the

specimens are acted upon by an all round effective stress, Pk =

58.5 lbs/in represented by the point B (see chapter 8). Such a

specimen is now mounted in the oedometer and, in fact, a "swelling"

test is initially carried out, i.e. nominally no change in height

is allowed, at the end of which the effective stresses are given by

the point C (see Table 10.6). For subsequent one-dimensional

consolidation the stress path CDD' is obtained. (Curve 1 of Fig.

10.74). It can be seen that the portion CD of the path ODD'

328.

is a straight line, giving Ro = A 0-ht/Arry' = 0.28. At higher

pressures the stress path CDD' curves down and finally becomes

parallel to the normally consolidated Ko line. It has also been

found from the triaxial tests that, following initial isotropic

stress conditions, Ro = 0.29 in the range of effective stresses

50 - 100 lbs/i0 The corresponding stress paths are shown by

BE and FG. It can be seen that all the Ko lines CD, BE and

FG are essentially parallel and to produce one-dimensional con-

solidation in the field, therefore, the effective stress path must

be parallel to them.

Let us now consider, as an example, that the element in

the field is subjected, under undrained conditions, to axi-symmetric

total stress increases, Acr vi . 30 lbs/in2 and AO- hl = 8 lbs/in2.

Taking the pore preasure parameter A = 0.5 (see Chapter 11, section

11.4), the excess pore pressure A u is given by

6 u = 8 + 0.5(30 - 8) = 19 lbs/in2

So, at the end of undrained loading, the new values of effective

stresses are

v = 32 + (30 - 19) = 43 lbs/in2

ht = 76 + (8 - 19) = 65 lbs/in2

329.

This will cause the effective stress point to move from A to X

and the immediate settlement will be a function of the effective

stress path AX.

The excess pore pressure will now begin to dissipate

and if the total stresses remain unchanged, the specimen will have

followed, during consolidation, the stress path XY" which is

parallel to the isotropic line 00". In practice, however, there

will be a reduction of the horizontal stress due to the decrease of

Poisson's ratio during consolidation. For points beneath the centre

of a uniform circular load the magnitude of this reduction, which

depends on the pore pressure parameter A and Poisson's ratio

has been analysed in Chapter 7. Assuming that the point is located

at a depth z/b = 0.5, Fig. 7.3 gives, for A = 0.5 and 4 = 0.22,

a stress increment ratio (K1 ) during consolidation of 0.75. The

corresponding stress path is shown by the line XY. Now the stress

path required for one-dimensional (oedometer) consolidation is XY1

(K' = 0.29), the line XZ being drawn parallel to the Ko lines

for the stress range under consideration. That the axial strains

for the two stress paths will be quite different can clearly be

seen from Fig. 11.54, which gives the following results:

Stress path K'= crh s/A (ry' E 1/E v

XY' 0.29 1.0 XY 0.75 0.56

330.

Since the volume change in the two cases will be essentially

the same the vertical deformation of the element in the field will

be only 56% of that for one-dimensional consolidation. In other

words direct use of the oedometer data will over-estimate the

vertical strain by as much as 79% (i.e. 0.44/0.56)

The above example emphasises the importance of taking

into consideration the relevant stress path in determining the

vertical strain of an element of soil during consolidation in the

field, because it is the integration of all such strains beneath a

foundation that gives the total settlement of a structure. A com-

plete procedure for such a settlement analysis is developed in

Chapter 13.

The preceding discussion implies that volume changes

during consolidation in the field are acoompanied by significant

changes in lateral strain. The customary assumption made in the

usual settlement analysis that the strain during consolidation is

one-dimensional (i.e. e 1

= e v) may be approximately correct in

certain problems (see Skempton and Bjerrum 1957), but, for an

ordinary building foundation, where the loaded area is of limited

extent compared to the thickness of the clay layer, this is not even

remotely so.

Only liMited data are available of field measurements of

lateral strain beneath foundations. Wilson and Hancock (1960)

measured horizontal movements of up to 3.4 in in the foundation clay

331.

beneath the North Ridge Dam in Western Canada. An overall short-

ening towards the point of maximum settlement of 6 in in a length

of 1200 ft has been observed by Hardy and Ripley (1961) in the

foundation of an aluminium smelter plant in Kitimat, British

Columbia. Measurements beneath a test embankment at Ska-Edeby,

Sweden, showed large horizontal movements which varied with depth

(Osterman and Lindskog 1963) and Eggestad(1963) measured vertical

and horizontal strains beneath a model footing on dry sand and

found that the maximum strains occurred at a depth of about 3/4

width of the loading plate. Cappleman (1967) presented data from

an extensive field study of horizontal movements of pipe conduits

under earth dams. Although not all horizontal deformations re-

ported in the above references are due to consolidation - there

are other factors, e.g. shear deformation and general ground sub-

sidence which also cause lateral strains - the data serve to in-

dicate that settlement of structures in the field are often accom-

panied by lateral deformation of the subsoil and the assumption of

one-dimensional strain is not generally valid. It is worth

emphasising, however, that the magnitude and kind of lateral move-

ment (inward or outward) will depend on the soil type and the re-

lative magnitude of the vertical and horizontal stress increments

and every case has to be considered on its merit.

332.

12.3 Rate of consolidation

A comparative discussion of the rate of consolidation for

oedometer and triaxial tests has been made in Chapter 11 (section

11.6). It has been shown that, for uni-axial drainage, the

Terzaghi theory predicts the average rate of consolidation of un-

disturbed London clay quite accurately for degrees of consolidation

of up to 70%, beyond which secondary and creep effects make the

observed rate slower than predicted. The dissipation of the

maximum pore water pressure (studied only in the high pressure oedo-

meter), on the other hand, proceeds faster than predicted at early

stages of consolidation, but slower towards the end (see Chapter 10,

section 10.2).

Similar values of the coefficient of consolidation (Cv)

and the coefficient of permeability (k) are obtained from both

triaxial and oedometer tests (see Figs. 11.75 and 11.76). Either

method can, therefore, be adopted to determine the value of Cv

required to calculate the rate of settlement of a structure in the

field. Much work still needs to be done, however, on the subject

of the rate of consolidation, particularly under three-dimensional

conditions, permitting both axial and radial flow of water. Both

the oedometer and the triaxial apparatus can be used for the purpose

provided, of course, that a representative sample is tested. (see

Escario and Uriel 1961, Mackinley 1961, Shields and Rawe 1965).

333.

CHAPTER 13

THE STRESS PATH METHOD OF SETTLEMENT ANALYSIS

13.1 Introduction

In the foregoing sections of the thesis theoretical and

experimental considerations have, been given to the study of stress

path and its influence on the deformation and consolidation of un-

disturbed London clay. The experimental work has shown with clarity

the importance of taking proper account of the stress path in ex-

amining the stress - deformation characteristics of London clay and

its relevance to settlement of structures in the field. In this

chapter the use of stress path in settlement analysis will be

demonstrated by considering a typical foundation in London clay.

It is emphasised that the problem does not refer to any real

structure and should be taken solely as an illustration.

13;2 Formulation of the problem

The problem to be studied consists of a circular, flexible,

smooth footing, 40 ft in diameter and founded at a depth of 20 ft

at a site for which the soil profile is taken to be the same as for

Bradwell - to which reference has already been made in the thesis.

The settlement analysis will be made for the centre of this circular

foundation. The Bradwell data have been chosen simply because the

effective stress - depth relationships have been established to

334.

considerable depth at this site, (Skempton 1961), and these are

necessary for the stress path method of analysis. It will further-

more be assumed that the experimental data presented in Chapters 10

and 11 (for the Ongar clay) will apply to the clay beneath the

foundation.

From considerations of bearing capacity of the soil (see

Skempton 1951) a net foundation pressure of 20 lbs/in2 has been

adopted. This gives a factor of safety against failure of more

than 3 and corresponds to a gross pressure of 28.8 lbs/in2 (1.85

T/ft2). The "buried" footing effect will not be considered in the

analyses that follow, Reference can be made to Fox (1948) for the

necessary corrections.

Distribution of stresses

Fig. 13.1(a) shows the distribution of vertical and

horizontal effective stresses in the soil before the commencement

of excavation. The data are taken directly from Skempton (1961).

It may be noted that a total depth of 160 ft (i.e. 8 x radius of the footing) beneath the foundation has been considered. This is in

accordance with Fig. 6.22 where it has been shown that almost 95%

of the total elastic displacement of a non-homogeneous medium takes

place within a depth equal to 8 x radius of the loaded area. It

will subsequently be demonstrated, however, that a considerably

smaller depth is adequate for calculation of the consolidation

335.

settlement. In order to obtain the in-situ effective stresses for

points beneath the depth of 90 ft - in the paper by Skempton the

data are given for this depth only - the Cr y' line has been ex-

tended linearly while the curve for Cr h' has been extended to a

point at a depth of 160 ft for which the effective horizontal

stress is given by Ko corresponding to the over-consolidation ratio

at this point! The distribution of the average effective stress,

(Tim' = (0-v' 20r 1/3) with depth, is also shown in Fig. 13.1(a)

while the variation of the effective stress ratio 07v1/0-h with

depth is indicated by curve 1 in Fig. 13.1(c).

The increases of vertical and horizontal total stresses

due to the net foundation pressure have been calculated for the

Boussinesq problem (Poisson's ratio s) = 1) and plotted in Fig.

1351(b). Although the elastic modulus of the clay almost certainly

varies with depth, it has been shown in Chapter 6 that such variation

does not influence the stress distribution in a semi-infinite medium

to any great extent - for which the Boussinesq problem gives a close

approximation. The settlement will, of course, be obtained by

numerical integration of the strains beneath the foundation, in

calculating which due account will be taken of the variation of the

elastic modulus with depth.

For Bradwell erosion has reduced the effective over-burden pressure by 210 lbs/in? This gives, at a depth of 160 ft ( Cry' = 74 lbs/in2), an over-consolidation ratio of 3.9, for which Ko = 1.2 (Fig. 10.76).

336.

The soil is now divided into twenty layers, as shown in

Fig. 13.1(a). The first eight layers are all 5 ft thick while the

lower twelve layers have thicknesses of 10 ft each. For an ideal

settlement analysis, samples from a number of different depths should

be obtained and be first brought to the stress conditions indicated

by Fig. 13.1(a) and then subjected to stress increments given by

Fig. 13.1(b). In the present illustration, however, the data for

the Ongar clay, presented earlier in the thesis, will be considered

to be applicable to all depths.

13.4 "Immediate" settlement

Table 13.1 summarises the data required for the calculation

of the "immediate" (elastic) settlement. It will be seen that the

effective stress ratio before construction varies throughout the

depth from 0.35 to 0.82 - hence the necessity of testing samples

from different depths. As indicated by Fig. 11.35, however, the

pore pressure parameter varies only slightly in this range of stress

ratio and an average value of 0.55 can be taken without great loss of

accuracy. The excess pore pressures set up under undrained con-

ditions have been calculated using this value of A [A u =A crh +

ACA Crlr - 411r1)] and shown by the dotted curve in Fig. 13.1(b)

- also tabulated in column 11 of Table 13.1. The vertical and

horizontal effective stresses as well as the stress ratios at the

end of construction, calculated on the basis of no volume change,

337.

are shown in columns 12, 13 and 14.

The Young's modulus, appropriate for each layer, can now

be obtained from Fig. 11.30, for the corresponding effective stress

ratio at the end of construction, (0-1)o referring to the

average effective stress prior to construction. The data are

shown in columns 15 and 16 and the variation of E with depth is

plotted in Fig. 13.1(d). The immediate settlement can now be

easily calculated by integrating the strains of all the layers

beneath the foundation, according to the equation

"cry — a- h dz i - r E (z)

(13.4.1)

The settlement thus obtained is 0.68 in (column 17 of

Table 13.1). It is interesting to note that, if a depth of 80 ft

(4 x Radius) only is considered, integration of the strains for

the first twelve layers gives a settlement of 0.61 in. Thus,

ignoring the lower 80 ft under-estimates the "immediate" settle-

ment by not more than 10%.

The above results can be compared with the "immediate"

settlement obtained by the conventional method of calculation.

Here, of course, the effect of stress path is not taken into con-

sideration and the average Young's modulus determined from stress

- strain relationships of standard undrained tests is 5,600 lbs/in

The "immediate" settlement is, then, (Terzaghi 1943),

338.

f) .212 (1 0 ) E (13.4.2)

where q = net foundation pressure (20 lbs/in2)

b = Radius of the circle (20 ft)

Ir= Dimensionless influence factor (1.5 for settlement of

the centre)

Substituting the appropriate values in equation (13.4.2) we have

= 1.29", compared to 0.68 in obtained by the stress path method.

In order to illustrate the use of Gibson's theoretical

analysis for non-homogeneous soil medium, numerical results of

which have been presented in Chapter 6, the elastic settlements have

been computed for the two cases shown by the broken lines in Fig.

13.1(d). The details of calculation are given in Table 13.2.

It will be seen that the two cases give settlements of 0,90 in and

0.58 in respectively while the actual settlement for the correct

variation of E with depth is 0.68 in.

13.5 Consolidation settlement

Let us consider an element of soil, beneath a foundation,

which undergoes both axial and lateral deformation during consolida-

tion. The volumetric strain of such an element due to the dissipa-

tion of excess pore water pressure 6, u, is given by

d E = (mv)3 . Q u (13.5.1)

339.

where (mv)3 is the coefficient of volume compressibility for

three-dimensional strain. The corresponding vertical strain is

..,_ X. (mv)3 . (13.5.2)

where X is the ratio of the vertical strain to the volumetric

strain. The vertical compression of the element during consolida-

tion can, then, be expressed as

d c X (mv)3 . b u . dz (13.5.3)

where dz is the thickness of the element. The consolidation

settlement of the foundation resting on a bed of clay of thickness

z is, therefore

pc = I X (mv)3 . A u . dz (13.5.4)

In the case of one-dimensional consolidation, in which

lateral strains are zero during load application as well as during

subsequent consolidation, the settlement is given by

z oed = f (m v)1 • 6. Cr • v 1

0 dz (13.5.5)

where Licr21"

is the increase of vortiea/( stress, and

340.

(mv)1 is the coefficient of compressibility for one-

dimensional strain.

Equation (13.5.5) gives the "conventional" method of calculating

final settlements and can be obtained by the straightforward applica-

tion of the oedometer test results (Skempton and McDonald 1955,

Skempton, Peck and McDonald 1955).

Skempton and Bjerrum (1957) modified equation (13.5.5),

taking account of the lateral strain during load application, but

still assuming one-dimensional strain during dissipation of the

excess pore pressures, and gave the following expression for the

consolidation settlement

Pctc1 (m v)1 )1 ts u . d (13.5.6) jr

Combining equations (13.5.5) and (13.5.6), Skempton and Bjerrum

obtained the following simplified relationship between the settle-

ment of a structure in the field and that obtained by the straight-

forward application of the oedometer test results

Pct =/1 fD oed (13.5.7 )

where the factor /1-t is a function of the soil type and the

geometry of the foundation.

Now it has been shown in Chapters 11 and 12 that volu-

341.

metric compressibility of undisturbed London clay, in the range

of stresses considered, is not significantly influenced by the

stress path and is primarily a function of the vertical effective

stress. So, for the same set of vertical stress increases the

oedometer and triaxial compressibilities are approximately equal,

i.e.

(mv)1 = (mv)3 = my (13.5.8)

Equations (13.5.4) and (13.5.6) can, therefore, be re-written as

) f a = Jx (mv) . dz )(13.5.4a)

o ) ) )

rci =( (my ) . u • dz )(13.5.6a) ) 0

Thus the settlement of a structure, where the foundation

soil undergoes three-dimensional strain (equation 13.5.4a) is

different from that given by the Skempton and Bjerrum method (equa-

tion 13.5.6a), even if the compressibilities may be identical. It

is only when A = 1 (i.e. one-dimensional strain) that the two co-

incidtt

In fact, there will still be a slight difference, because, in the Skempton and Bjerrum method my corresponds to the in-situ vertical effective stresses while in the stress path method my should correspond to the vertical effective stresses at the end of construction.

342.

The parameter X will, in general, depend on the soil

type, stress level and the stress increment ratio and should be

determined experimentally (see Chapter 11).

For the problem under consideration, the consolidation

settlement will be calculated by both equations (13.5.4a) and

(13.5.6a), the two methods being referred to as the stress path

method and the Skempton and Bjerrum method respectively. The com-

pressibility data obtained from the high pressure oedometer tests

(see Fig. 10.54) will be used throughout - the relevant data for

the stress range of the problem are shown in Fig. 13.3.

(a) The stress path method

The complete procedure for calculating the consolidation

settlement by the stress path method is summarised in Table 13.3.

It can be seen from column 6 and Fig."13.2(a) that the increase of

vertical effective stress during consolidation, expressed as a ratio

of the effective stress before consolidation, decreases sharply

with depth. At 45 ft the ratio is only 9%. It has been shown

from oedometer tests (see Chapter 10) that after long rest periods

at the in-situ overburden pressure, stress increases of less than

10% do not cause any appreciable volume change. So, the consolida-.

tion settlement has been calculated for this depth only, the clay

underneath being assumed to undergo no volume change.

The compressibility data, tabulated in column 7 are taken

directly from Fig. 13.3 - corresponding, for each layer, to the

343.

effective vertical stress before consolidation. The stress in-

crement ratio K' (column 9 and Fig. 13.2(c)) depends on the

Poisson's ratio of the material as well as on the pore pressure

parameter A and has been obtained for values of A = 0.55 and

2,7 = 0.21; from the theoretical work presented in Chapter 7 (see

Fig. 7.3). The corresponding values of -X , tabulated in column

9, have been interpolated from the experimental curve in Fig. 11.54.

The last column of Table 13.3 gives the vertical compression of

the individual layers which are added up to obtain the consolida-

tion settleinent (1.61 in).

(b) Skempton and Bjerrum's method

The calculation of consolidation settlement by Skempton

and Bjerrum's method is shown in Table 13.4. Once again the

effective depth is taken to be the depth - in this case 60 ft -

within which the vertical effective stress increments are greater

than 10% of the in-situ stresses. A settlement of 4.30 in is

obtained - which is the settlement given by the "conventional"

method (equation 13.5.5). The factor ft of Skempton and Bjerrum's

method is 0.69 (for A = 0.55 and z/b = 1.5) and the corresponding

settlement is, therefore, 0.69 x 4.30 = 2.96 in.

Poisson's ratio = 0.21, corresponding to 170 = 0.29 - for an isotropic material - has been used in the analysis (see Table 11.7).

3/14.

13.6 Comparison of different methods of settlement analysis

Table 13.5 summarises the "immediate" and consolidation

settlements obtained by the different methods of analysis described

above. It can be seen that the stress path method gives settle-

ments - which are only 54% of those given by Skempton and Bjerrumts

method. The almost identical ratio obtained for both "immediate"

and consolidation settlements is, perhaps, fortuitious, but the

effect of taking proper account of the stress path in settlement

analysis is amply clarified by the data in Table 13.5.

One final point is of interest. Column 9 of Table 13.2

indicates that A varies only slightly within the depth considered

so that a weighted average of 0.53 can be assigned for the entire

problem. Referring, now, to equations (13.5.4a) and (13.5.6a)

it can be seen that for constant A, ec = h pct*, where p cl

is the settlement given by Skempton and Bjerrum's method. Sub-

stituting A = 0.53 .and f)c1 = 2.96 in, we get the settlement

under three-dimensional conditions, = 1.57 in which is slightly

less than the settlement obtained by numerical integration in Table

13.2.

This statement is only approximately correct because the compressibilities for the two methods of analysis are slightly different (see Footnote on P-341)

345-

13.7 Rate of settlement

The final step in a settlement analysis is to calculate

the rate of consolidation settlement and thus to determine the

complete time - settlement relationship for the structure concerned.

In the analysis that follows it will be assumed that the excess

pore pressures do not begin to dissipate until construction is

complete and that secondary consolidation is absent.

The settlement of a structure at any time t after the

end of construction is given by

t- (13.7.1)

f)

where P. is the "immediate" (elastic) settlement (0.68")

is the total consolidation settlement (1.61") and c

U is the degree of consolidation settlement at time t

as evaluated from the theory of consolida-

tion

The rate of consolidation settlement of structures founded

on clay is usually calculated on the assumption that flow of pore

water occurs in the vertical direction only, for which Terzaghi's

theory of one-dimensional consolidation applies. Although it is

widely recognised that this is too simplifying an assumption - the

flow of water in the field being generally three-dimensional - the

mathematical solutions to problems involving three-dimensional con-

346.

solidation have not been easily available to engineers. As a con-

sequence Terzaghi's theory is almost universally used to predict

the rate of consolidation although it is generally known that such

predictions are usually slower than actually observed in the field,

particularly for structures on over-consolidated clays. In the

following paragraphs, the rate. of settlement of the structure under

consideration will be predicted by both Terzaghi's theory and some

approximate three-dimensional solutions for which numerical results

are available.

(a) One-dimensional consolidation

The process of one-dimensional consolidation, when the

distribution of total stresses remains unchanged, is governed by the

well known Terzaghi equation

2_ c " = " vi z2 t (13.7.2)

where Cv1 is the coefficient of consolidation for one-dimensional

strain and

u(z, t) is the excess pore pressure, which is a function

of the co-ordinate z and time t.

The degree of consolidation at any time t is given by

U = 1 fut dz f uo dz

(13.7.3)

347.

where ut is the excess pore pressure at time t and

uo is the initial excess pore pressure.

The solution of the above equation for a number of cases with

different boundary conditions have been given by Terzaghi and

Frbhlick (1936) (see also Janbu, Bjerrum and Kjaernsli 1964). For

the problem under consideration the distribution of the initial

excess pore pressure with depth is shown in Fig. 13.2(b). Con-

tinuing with the assumption that only the first 45 ft contributes

towards the consolidation settlement the problem is defined by

an impermeable boundary at a depth of 45 ft dnd a fully permeable

upper boundary (it is assumed that the base of the footing is fully

permeable). Although an exact solution for the initial condition

shown in Fig. 13.2(b) is not available* the problem can be simpli-

fied to the case of a triangular distribution of the initial ex-

cess pore pressure as indicated in the figure. The corresponding

relation between the degree of consolidation U and time factor

Tv (defined as Tv

= C vt/H2 where H is the maximum drainage

length) has been obtained from Janbu, Bjerrum and Kjaernsli (1964)

and the complete calculation of the time - settlement relationship

is shown in column A of Table 13.6. A value of Cv = 0.003 in2/

min (11 ft2/year) for the average stress range within the effective

depth has been taken from the oedometer test data (Chapter 10).

woorcs...seurevikamem.....m.mr

It can, of course, be obtained numerically (Gibson and Lumb 1953).

348.

The time - settlement relationship thus obtained is plotted in

Fig. 13.4 (curve A). It can be seen from both Table 13.6 and Fig.

13.4 that the rate of settlement is very slow taking more than 50

years to reach 80% of the total settlement.

(b) Three-dimensional consolidation

It has been explained earlier that the process of con-

solidation in the field is interconnected with the problem of stress

distribution and any rigorous analysis must consider the two together.

The general theory of three-dimensional consolidation for an iso-

tropic elastic medium was developed by Biot (1941), who later ex-

tended the theory to the more general case of the anisotropic

medium (Biot 1955, 1957). Rigorous analytical treatment of

practical problems, however, leads to very complex mathematics and

solutions have so far been obtained for a few special cases (Gibson

and McNamee 1957, 196o, 1963, de Josselyn de Jong 1957). Numerical

data are not yet available for ready use in practical settlement

analysis.

In what follows, approximate numerical solutions of

Gibson and Lumb (1953) and Davis and Poulos (1966) will be used to

study the influence of three-dimensional consolidation on the rate

of settlement of the structure under consideration.

The general Biot equation for three-dimensional consolida-

tion is

349.

2u -6 2u -a 2u 1 0 -611 C + + (13.7.4) v3 I) x2 -6 y2 -6 z2 3 -Ot

where Cv3 is the three-dimensional coefficient of consolidation

and G is the sum of the total normal stresses

( Cr „ cr cr ) xx yy zz

In the particular case where the total stresses remain constant

o (i.e. at = 0) the above equation reduces to its simplified form

"a 2u )2u -6 2u Cv3 x2 y2 ± z2

(13.7.5)

which is an extension of the Terzaghi one-dimensional equation

(Terzaghi (1943)).

(i) Gibson and Lumbe solution

Gibson and Lumbe (1953) obtained a numerical solution of

equation (13.7.5) by using the finite difference method. The

particular problem they considered was the rate of consolidation of

the centre of a uniformly loaded circular footing founded on a

stratum of clay whose thickness was 3.2 times the radius of the

footing and which was impermeable at its base. The initial excess

pore pressures were assumed to be equal to the increase of vertical

total stresses (i.e. pore pressure parameter A = 1.0). Although

the problem under consideration is somewhat different from the one

described above, Gibson and Lumbe suggested that their solution may

hold, at least for short times, for other thicknesses of the com-

cr Z), for Poisson's ratio

z *.

350.

pressible stratum. In order to get a rough estimate of the

difference between one and three-dimensional methods of analysis

Gibson and Lumbe's solution can, therefore, be used as a first

approximation. The time - settlement relationship thus obtained

is shown by curve 2 in Fig. 13.4. The details of calculation are

given in columns B of Table 13.6. Values of Cv = 0.005 in2/min

(11 ft2/year) and H = 45 ft - the same as for the one-dimensional

analysis - have been used. Comparing curves 1 and 2 it can be

seen that the rate of settlement for three- dimensional consolida-

tion is decidedly faster - taking 15 years to reach 80% of the

total settlement as compared to 50 years for one-dimensional con-

solidation.

(ii) Davis and Poulos solution

Davis and Poulos (1966) adopted an approximate method - 2

for t#e ntmerical solution of the three-dimensional consolidation

equation for circular and strip footings founded on layers of finite

depth. Their method which is called "the crossover procedure" is

described below.

The rate of dissipation of the average pore water pressure

is first obtained by solving equation (13.7.4) in which 1)Q/7t is

assumed to be zero and the distribution of the initial pore pressure

is given by ui = 0/3 where 0 is the bulk stress (Cr cryy xx

Suchla dissipation curve

The stress distribution is obtained from Poulos (1967), referred to in Chapter 6.

351.

for the geometry of the problem under consideration (H = 45 ft,

R = 20 ft) is shown as curve 1 in Fig. 13.5. Next, a stress dis-

tribution is obtained for Poisson's ratio ::,)= 0 and equation

(13.7.4) is solved a second time for the new values of the bulk

stress, again assumed to remain constant during consolidation.

The corresponding curve for our problem is shown as curve 2 in

Fig. 13.5. (Curves 1 and 2 are called the Z0.5 and Z0 curves

respectively). Now, for a medium having = z during load

application and ";)= 0 at the end of consolidation, it is postu-

lated that the true time - settlement relationship, alloying for

internal stress redistribution, follows a curve that starts as the

0.5 curve and gradually crosses over to the Zo curve..;..., It is

further assumed that this crossover takes place in proportion to

the degree of settlement, i.e. the corrected degree of settlement

is

uo z0.5 + u o (z0 - z0.5 )

(13.7.6)

Davis and Poulos found that the numerical solution for

the semi-infinite medium obtained by the approximate crossover

procedure shows very good agreement with the rigorous analytical

solution of Gibson and McNamee (1960).

To facilitate calculation of the degree of consolidation

for a soil having a Poisson's ratio •‘)= )' a further simplification

352.

is achieved by assuming that the Z 4 / curve is a linear inter-

polation between the curve for = 3 (the Z0.5 curve) and the

curve for ;,)=- 0 (the Z0 curve). The corrected degree of

settlement is, then, given by

U 0.5' - (13.7.7) 1 (1 2'')(Z0 - Z0.5)

For our problem, the corrected U vs Tv relationship is obtained

by substituting ' = 0.21 in equation (13.7.7) (see curve 3 of

Fig. 13.5).

The field time settlement relationship for the problem

under consideration has been calculated by the approximate Davis

and Poulos method (see columns C of Table 13.6) and plotted as

curve 3 in Fig. 16.4. It can be seen that the rate of settlement

so obtained is even faster than the Gibson and Lumb solution, 80%

of the total settlement, in this case, occurs after only' 11 years.

It is too early to judge the reliability of Davis and

Poulos' method of estimating the rate of settlement under.bhree-

dimensional conditions. The author is not aware of the method

being applied to any full scale structure at the present time. It

appears from a comparison of the three time ,= settlement curves in

Fig. 14.4 that the method is at least a good approximation.

The above analysis leaves no doubt whatever that, where

flow of water is three-dimensional, the assumption of one-dimensional

353•

consolidation grossly under-estimates the rate of settlement. Much

of the discrepancy between calculated and observed rates of settle-

ment of structures can, therefore, be attributed to this inaccuracy.

Referring to the case records mentioned in Chapter 3, it can be seen

that, for structures on over-consolidated clays, settlement virtually

ceases after only 4 - 5 years of construction. This does not

appear unreasonable. If settlement at the end of construction is

as high as 50% of the total, three-dimensional consolidation will

cause most of the settlement to occur within a few years after cones

struction. Furthermore, horizontal permeability in the field, if

greater than the vertical, will hasten the process of consolidation,

leading to an even faster rate of settlement. This factor, which

has not been considered in the present work, is still to be in-

vestigated. Non-homogeniety of the soil with respect to com-

pressibility and permeability also affects the rate of settlement in

the field (Schiffman and Gibson 1964). Finally, geological details

such as small layers of silt or sand which may not be easily detected

from ordinary borings may radically idfluence the rate of settlement

of full scale structures (Rowe (1968)).

TABLE 13.1

CALCULATION OF IMMEDIATE

Layer No.

(1)

Thick- ness (ft)

(2)

Depth* below founda- tion(3-1 (ft) (3)

Effective stresses construction (lbs/in2)

07-h ' 0 ' v , m

(4) (5) (6)

before

0-1

Stresses at end (lbs/in2)

60-v ACII.

(8) (9)

-I"- h

(7)

1 2 3 4 5 6 7 8

9 10 11 12 13 14 15 16 17 18 19 20

5.o 11

II II

I, il " II

10.0 n ?I

tt ti it " " " " " "

2.5 7.5 12.5 17.5

22.5 27.5 32.5 37.5

45.o 55.0 65.o 75.0 85.0 95.0 105.0 115.0 125.0 135.0 145.0 155.0

9.5o 11.60 13.70 15.80

17.90 20.00 22.10 24.20

27.30 31.40 35.50 39.60 43.70 47.80 51.90 55.0o 59.10 63.2o 67.30 71.40

27.o 32.2 36.70 40.50 44.o 46.8 49.4 51.5

54.4 58.0 61.2 63.6 65.9 69.5 72.5 75.0 78.0 81.0 84.o 87.0

21.17 25.33 29.03 32.27 35.3o 37.87 40.30 42.4

45.37 49.13 52.63 55.60 58.50 62.26 65.63 68.33 71.70 75.06 78.43 81.80

0.35 0.36 0.38 0.39 0.41 0.43 0.45 0.47

0.50 0.54 0.58 0.62 0.66 0.69 0.72 0.73 0.76 0.78 0.80 0.82

19.90 19.00 17.00 14.00 11.50 9.2o 7.44 6.2o

4.6o 3.44 2.5o 2.00 1.60 1.28 1.05 0.88 0.75 0.65 0.57 0.50

16.50 10.24 5.86 3.20 1.86 1.10 0.68 0.52

0.25 0.11 0.07 0.04

Depth to centre of layer

SETTLEMENT BY THE STRESS PATH METHOD

of construction

&Tv - A u 0- , OM

(ii) (12)

Cr h i

(13)

lbs/in

(1o)

0'v' E E 2

(16)

Si Pi

(17)

( irm' )o

(15)

Grh'

(14)

3.4o 18.37 11.03 25.13 0.44 284 6,010 .034 8.76 15.06 15.54 27.42 0.57 247 6,260 .084 11.14 11.99 18.71 30.57 0.61 235 6,890 .097 10.80 9.14 20.66 34.56 0.60 238 7,680 .084 9.64 7.16 22.24 38.70 0.57 242 8,540 .068 8.10 5.56 23.64 42.34 0.56 25o 9,470 .051 6.76 4.4o 25.14 45.68 0.55 253 10,200 .040 5.68 3.64 26.76 48.38 0.55 253 10,730 .032

4.35 2.64 29.26 52.01 0.56 25o 11,340 .046 3.33 1.94 32.90 56.17 0.59 241 11,84o .034 2.43 1.41 36.59 59.86 0.61 235 12,370 .024 1.96 1.12 40.48 62.52 0.65 222 12,340 .020 1.60 0.88 44.42 64.30 0.69 213 12,460 .016 1.28 0.70 48.38 68.22 0.71 207 12,890 .012 1.05 0.58 52.37 71.45 0.73 204 13,400 .010 0.88 0.48 55.4o 74.12 0.75 197 13,460 .008 0.75 0.41 59.44 77.25 0.77 192 13,800 .006 0.65 0.38 63.47 80.35 0.79 187 14,040 .005 0.57 0.32 67.55 83.43 0.81 182 14,300 .004 0.50 0.28 71.62 86.50 0.83 178 14,560 .002

E. .677"

354.

355.

TABLE 13.2

CALCULATION OF ETIMIATE SETTLEMENT IN

NON-HOMOGENEOUS MEDIUM

Circular Footing : Diameter (2b) = 40 ft. Net Foundation pressure (q) = 20 lbs/in2

Immediate settlement,

. —213— I I I G(0) f

where G(0) = E(0) = shear modulus at the surface.

r

Case Fig. 13.1(d)

1 2

E(0) 2 lbs/in

6,000 11,000

G(0) 2 lbs/in

2,000 3,700

m Fig. 6.9

54 22

G(0) p /b

5.55 25

1 I? Fig. 6.21

0.375 0.45

P. . (in)

0.90 0.58

(3- m

111 500

TABLE 13.3

CONSOLIDATION SETTLEMENT

Layer

(i)

Depth to Bottom of Layer (ft)

(2)

Thickness dz

(ft)

(3)

( 0- ' )o

lbs/in2

(4)

v &O- '

lbs/in2

(5)

r" C1.1 \

tr \v) C. - 00 ON

5.o 5.o 11.03 18.37 10.0 II 15.54 15.06 15.0 n 18.71 11.99 20.0 I? 20.66 9.14 25.0 1, 22.24 7.16 30.0 n 23.64 5.56 35.0 n 25.14 4.4o 40.0 IT 26.76 3.64 50.0 10.0 29.26 2.64

Note: ( 6lr')0 = Effective vertical stress at the end of consolidation colisiruclioln.

= Effective vertical stress increase during consolidation ( = 6u)

(STRESS PATH METHOD)

AC V I M X 10-4 g2/110,

ACht

K' - -\. C1 J . ---

V S r in

c = x, . M .

6 cr :I. . dz

( 0- ') V 0 6 cry '

(6) (7) (8) (9) (10)

1.67 • 6.33 0.72 0.58 0.404 0.97 6.01 0.75 0.56 0.304 0.64 5.81 0.78 0.54 0.226 0.44 5.70 0.79 0.53 0.166 0.32 5.6o 0.80 0.52 0.126 0.24 5.51 0.81 0.51 0.094 0.18 5.43 0.81 0.51 0.073 0.14 5.35 0.82 0.51 0.059 0.09 5.22 0.83 0.50 0.160

sum. 1.610

356.

TABLE 13.4

CONSOLIDATION SETTLEMENT

Layer Depth to Bottom of Layer (ft)

Thickness dz

(ft)

6- v t

lbs/in2 Acr v i

lbs/in2

1 5.0 5.0 9.50 19.90 2 10.0 II 11.60 19.00 3 15.0 It 13.70 17.00 4 20.0 It 15.50 14.00 5 25.0 I/ 17.90 11.50 6 30.0 II 20.00 9.20 7 35.0 It 21.10 7.44 8 4o.o it 24.2o 6.20 9 50.0' 10.0 27.30 4.6o 10 60.0 10.0 3.14o 3.44

"Conventional" settlement f)o = 4.30 in "Skempton and Bjerrum" settlement p

01 = r- oed a = 0.69 x 4.30

= 2.96 in

Notes: = In-situ vertical effective stress = Total vertical stress increase Cry'

(SKENPTON AND WERRUM METHOD)

4o-- vpoed ( m x 10 4

iv2 n /lb in )

7... mv . Acrv ' . dz ( 0- ' )o

2.09 6.45 0.770 1.64 6.28 0.716 1.24 6.13 0.625 0.90 6.00 0.504 0.64 5.86 0.404 0.46 5.73 0.316 0.35 5.6o 0.250 0.26 5.48 0.204 0.17 5.32 0.294 0.11 5.12 0.212

Sum 4.30 in

357.

358.

TABLE 13.5

COMPARISON OF SETTLEMENTS CALCULATED BY DI1FbRENT METHODS

Type of Stress Path Skempton and Conventional Settlement Method

(A) Bjerrum's Method

Method (i.e. Standard

Ratio TET

(B) Oedometer)

"Immediate" Settlement 0.68 in 1.29 in 0.53

Consolidation Settlement 1.61 in 2.96 in 0.54

Total Settlement 2.29 in 4.25 in 4.30 in 0.54

TABLE 13.6

RATE OF SETTLEMENT

Time (t) (years)

C t T =-2--r

U%

(A)

Ur (in)c

Terzaghi

r.+Ii. 10 (in) °

P +UP t i l c

(B)

u%

Gibson

U p (in,

v H2 Ci + (%

1 0.0054 14.5 0.23 0,91 0.40 17.5 0.28 2 0.0108 21.2 0.34 1.02 0.45 26.0 0.42 4 0.0216 29.o 0.47 1.15 0.50 40.0 0.64 6 0.0324 34.5 0.56 1.24 0.54 49.5 0.80 8 0.0432 38.5 0.94 1.30 0.57 58.0 0.93 10 0.0540 41.5 0.67 1.35 0.59 63.o 1.01 15 0.0810 48.o 0.77 1.45 0.63 73.0 1.17 20 0.1080 53.0 0.85 1.53 0.67 78.0 1.25 25 0.1350 57.0 0.92 1.60 0.70 82.5 1.33 50 0.2700 70.0 1.13 1.81 0.79 100 0.5400 90.0 1.45 2.13 0.93

Notes: H = Length of drainage (45 ft) = C = Coefficient of consolidation = 0.003 iX2/min

1i = "Immediate" settlement (0.68 in) Consolidation settlement (1.61 in) pi+.1)pec = Total settlement (2.29 in)

359.

and Lumb

+ U IC. 01)

ri +II PC

ri pc

(0) Davis and Poulos

P + UP 1 (in) c

P.

ri

0.42

28.o 0.45 0.48

37.0 0.60 0.58

50.0 0.81 0.65

58.5 0.94 0.70

65.0 1.04 0.74

69.5 1.12 0.81

78.0 1.26 0.84

82.0 1.32 0.88

86.0 1.38

0.96 1.10 1.32 1.48 1.61 1.69 1.86 1.94 2.01

1.13 1.28 1.49 1.62 1.72 1.8o 1.93 2.00 2.06

0.49 0.56 0.65 0.71 0.75 0.79 0.85 0.87 0.90

360.

CHAPTER 14

CONCLUSIONS AND SUGGESTIONS FOR FURTHER RESEARCH

14.1 Conclusions

The settlement of structures on over-consolidated clays

has been studied in the light of the influence of stress path on

the deformation characteristics of such clays, with particular

emphasis on London clay. It has been shown that the stress path

of an element of soil beneath a foundation in the field differs in

many important respects from those implied in the existing methods

of settlement analysis and predictions of settlement are, therefore,

not always accurate. Deformation of soil being essentially path

dependent, it is necessary, for successful application of laboratory

data to field problems, that a soil be tested under the same set

of stresses that it is likely to undergo in the field and a method

of analysis has been proposed that takes this into consideration.

The stresses and displacements in non-homogeneous elastic

media, whose modulus of elasticity varies linearly with depth, have

been calculated from Gibson's analytical solution (Gibson 1967).

The results show that the distribution of stresses in such a medium

(incompressible) differs only slightly from the Boussinesq stress

distribution for identical boundary load. It is, therefore,

possible to calculate the "immediate" (elastic) settlement for a

non-homogeneous medium by numerkblintegration of the strains

361.

beneath a foundation, assuming Boussinesq stress distribution but

taking proper account of the variation of the Young's modulus with

depth.

In a foundation problem, the modulus of elasticity of the

soil generally varies with depth as a consequence of increasing

effective stresses before construction as well as due to the

different stress levels that are imposed by the applied foundation

pressure. To obtain the true variation of E with depth, there-

fore, it is essential that an undistrubed sample be first brought

back to the stresses prevailing in the ground before sampling and

then subjected to stress increments that it is likely to undergo

in the field. It is found that the Young's modulus of London clay

so obtained differs considerably from that determined from standard

undrained tests. The same is true of the pore pressure parameter

A, although to a lesser extent.

In calculating the consolidation settlement, the influence

of lateral stresses cannot be ignored. The experimental data show

that although the volumetric compressibility of undisturbed London

clay is primarily a function of the vertical effective stress and

is largely independent of lateral stresses - at least within the

range of stresses considered - the vertical strain is greatly in-

fluenced by the relative magnitude of the vertical and lateral

stress increments during consolidation. Direct use of the oedo-

meter test results will give accurate prediction of the consolida-

362.

tion settlement only when the stress increment ratio K' ( = 110-h t/

vl ) during consolidation is equal to the Ro value of the

material (i.e. when the strain is one-dimensional). The ratio K'

would, of course, be equal to 1 but for the redistribution of

stresses in the foundation due to the decrease of Poisson's ratio

of the soil from 2 during load application to the drained value

after full consolidation. An approximate method of calculating

K' for points beneath the centre of a circular foundation has been

suggested, -but it is necessary to determine experimentally the re-

lationship between K' and the vertical strain for the problem

concerned.

The volumetric compressibility of London clay can be

determined from both triaxial and oddometer tests, the latter giving

satisfactory results only when the initial swelling is positively

prevented. For this it is necessary to eliminate all bedding

errors as well as any deformation of the apparatus which may be

important when the clay tested in the oedometer undergoes only

small deformations. It is, then, possible to use the oedometer

compressibility (my) data coupled with the relationship between

K' and k ( = Vey) to predict the consolidation settlement

of a structure. It is also important to note that the pressure

increment ratio (Ap/p) beneath a formation decreases rapidly with

depth and it may be necessary to exceed a threshold value before

apprecaible volume change can occur. This can only be detected,

363.

however, from °odometer tests where specimens are allowed long rest

periods and then subjected to small load increments. Based on

tests with 90 days' rest followed by small load increments, a

threshold value of 10% of the in-situ overburden pressure has been •

suggested for London clay. This consequently reduced the effective

depth beneath a foundation which contributes towards the consolida-

tion settlement of a structure. A method of settlement analysis

taking all these factors into consideration has been proposed.

The example of a circular foundation shows that the settlement

calculated by the stress path method is considerably smaller than

that given by either the conventional method or the Skempton and

Bjerrum method of analysis.

The incremental Ro value of undisturbed London clay is

found to depend on stress level. For the Ongar clay it is as

small as 0.29 at the low pressure range (<100 lbs/in2), but in..

creases with increasing pressure until it becomes equal to the

Ko value (0.64) for remoulded London clay. The relationship between K' and X during anisotropic consolidation can be

analysed in terms of anisotropic elasticity. A method of determ-

ining the necessary parameters (E1"

3' ,)1" ,)2" .)3') from

triaxial tests is demonstrated.

The pre-consolidation pressures of London clay at Ongar

and Wraysbury have been determined from tests on specimens loaded

2 to pressures of up to 7,000 lbs/in in the controlled rate of strain

364.

oedometer. It has been assumed that, on the e vs log p plot,

the geologic rebound is parallel to the laboratory rebound, pro-

vided the latter is begun from pressures greater than the pre-

consolidation pressure. Limited field evidence available suggests

that this assumption is at least approximately valid for the cases

considered. It is found that the pre-consolidation pressure at

Wraysbury, to the west of the London Basin, was considerably greater

than that at Ongar, to the east, thus supporting the estimates of

Skempton (1961) for Bradwell and Bishop et al (1965) for Ashford

Common.

The consolidation tests in the triaxial apparatus as well

as in the hydraulic oedometer, allowing vertical drainage only,

show that Terzaghi's theory of one-dimensional consolidation pre-

dicts the rate of volume change and the rate of settlement of un-

disturbed London clay extremely well although the prediction of the

rate of dissipation of the maximum pore pressure is less accurate.

In the field, however, most settlements take place under three-

dimensional conditions. In such cases the one-dimensional theory

grossly under-estimates the rate of settlement. Only a few approx-

imate numerical solutions to the three-dimensional problem are at

present available, the application of which are shown to improve

considerably the prediction of the rate of settlement of structures

in the field.

365.

14.2 Suggestions for further research

It emerges from the work presented in this thesis that

further research may fruitfully be carried out on the following

subjects.

(a) More work is required to study the influence of stress

paths on the stress - strain modulus of undisturbed clays. Samples

should be brought back to the stresses prevailing in the ground and

then subjected to further load increments. , It is desirable that

samples be obtained from different depths to see if the data for one

depth can be applied to an entire foundation problem.

(b) The influence of lateral stress on the deformation of clay

during drained compression needs to be studied more thoroughly,

again with samples from different depths. Two distinct points

should be considered - the influence of the lateral stress on

(i) volume change and (ii) the ratio of the axial strain to volu-

metric strain.

(c) Reliable estimates are required of the in-situ vertical

and horizontal stresses in soil media. As direct measurements

seem difficult, if not impossible, model studies to simulate de-

position and subsequent erosion may be considered.

(d) The question of threshold value and the influence of small

load increments following long rest periods should be studied, using

for example, an hydraulic oedometer of the type described in the

thesis.

366.

(e) Theoretical and experimental work on three-dimensional

consolidation with numerical results therefrom are urgently required

to predict more accurately the rate of settlement in the field.

a

Sin oe CO5 (t «) e

(4) s;r) o< cos ( + D<) FN ,x) -(11a) FC/60(1 c a (6.22)

.4 oc

( p/b)

co SInc<Cos(%,00

(f...., 4.4 )0(IF (q -4)cx) +.44 +21o9 ((-,, -4 )0() '

6,4 (6.23) - [Fo +Alpo - r(-,2,,,x)] +2 4- 0/7i„ + it e

A (p/6)

( (.+.) 1..4)o ) + F (0)

(p/b)

-1-21090 44,)°‘ 1 _ z/b

b (6.24) 0

.00

ot. ,c) e—

A 03/0 94.

(rxqz (6.a5)

Stresses

0

APPENDIX A

Put b = c‹, = , d = V-

1. Plain Strain

= cif)

2nG(0)

Displacement:

do( (6.26)

do< (6.27)

where to ( Pc 091

2. Axi—symmetric case

Displacements:

u. (f; y (lb 4G(0)

z q6

Stresses

crzz -7f- T 2.

00 el°( FIto() + FP6 41-)49 Jo (rod J , (a) I 2 (PAM to ((l )0(

A1 A 04)

oo

(f,- 0) Ji 60 f 1+ (4°) F (CA F (4). + )°( oC

00 Jo (f o) J (-0

(:t [F((b+ 4).4) + F ( 0e).1 -[Fq 4.0) - F(to)]

2 +4)0( 1o9 ((**- 4)°0 + 2

(6.28) do(

A (134)

_cry, b• • 2

t)c( [ 1 tia:(jribo (1)0,0 I [F*Lt + 4 12) + F 4, c‹) + filiri; 1 (6.29) Jo (f00 Ji 60 (lac A (i3/0 -''' ..,- 2 to9 ( 72,- +Ceti + LF a +4 ) 0( - F q0()) + -171.--; t I ao-

op i ...,

0

d of (6.30)

.o0 z J 1 ( r/b k) -1 i (X) e- -6

A ( 3/4) 416_ _ (-f;+.1 )0< [F -f, +t)0‘)— F (%o< /6.

r where F ( Po oo + Lo 9 (-ic- 0)-1 + 1 + 2--i ca,

APPENDIX B

Limiting values

These are obtained by taking the right hard sides

P and 7--> O.

1. Vertical displacements

(a)

The expression

oc)

of equations (6.22) - (6.30 to the limits

1 4. (43;"°() F (4: °) r1.7)°(

can be written as S

b kIF (1b 6-4 log( 140'() + + 1

1 f(t34,04) e2-64)°c (-21cx) —tog (0() -e2(g

2 4-0c

2.0 04 2 404.)] 4.1 +1 21340(..

s =

- e + 0

2/44)

\ - ),k —

,) and taking

for large

limits

Now E.( —`X) -

Substituting for

x z W b

1.3)4Y., Sin 0( Cos (0) e

oe...2

Z

and for the axi-symmetric loading

s a (1 ±

For strip loading: o0

(6.31)

e- 1) °c 1̀0(f'`) J1(°‹ ) + °() doIL (6.32)

(b) ---->o The expression

+ (tx) F - F ((t- t )00 5 -

w

= qb

b b 2G(0)

040( F (kot) + log (P4,0)j + 1 +

÷ (P/b,g)[e24°' E (- 2 ia9 (.0.f., ex) ) I (-2 (16 ) la' VC; 4 )

-I- 2 Wv>c

Now E ( X) = + log )\ + 0( ) for small )‘ .

Substituting for Ei - 2 40-0C) and taking limits

s = 2

so, for strip loading

/b [c E ; (-2 Pi,- 04)1 +-1

00

b 09 Sin o< Cos (l X) e \

c4,

x z w (b b = —2—

m

dc< ( 6. 33 )

0

For axi-symmetric loading

r z = b b 2m ° Jo f c4) j1(Q) d°( (6.34) N

Sin(DC) Cos ( 04-) z oc

b (I- 3- 0C 2

q It

2. Stresses : following similar argument

For both limits 14-t--°0 and 0 the stresses are identical:

Strip loading

co C

Sin(O() Cos (-x-:°()

Crzz 2 =

, q

e - Toc)((1

dot-

b /

o0

x- z 2 q fl u

0

Sin( a() Sin (25.0()

oc

_ • e

Axi-symmetric loading

cx J o b () J1 (0K) e b Eck

o_zz

q

z do<

q

cr rr = j I0.1.),S (01, ) e ° b

s.ac) j 1 (c) e 1( b

Z oc b 1 + 1

.. z zA. b (z ')c,i., . o.

f 0,, (b (4) j 0 \ A

d c)c-

q

cr. rz

1

(6.39)

375.

APPENDIX C

The following relationships have been used throughout the

computations:

Exponential Integral Function

?() ti

(i) (Tabulated in Janke & Emde 1943 . Selby and Girling 1965) For small \

x + log +

where '6 = Euler's constant = 0.577215

For large ,

1 ->\

E.(- = - e 1 0 L

Function

FO‘) = e2 ‘ E. (-2.\ ) - logy

(iv)

1. Displacements

1 + 2 --b c)c

+ 1 + 1

2 ct (24

376.

In the limits:

A= 0° ) )

For all r--- A

Fig. C.1 shows the distribution of the function

Jo(0)J1(-'7%,) 1 . —

for the settlement w(0, 0) beneath an axi-symmetric loading.

All integrations were performed by the traphezoidal rule and the

results for were checked with the known values of the

surface settlements beneath a circular load on a homogeneous medium

AID:yin and Ulery (1962).

IP

r _ b

Ala vin and Ulery (1962)

Calculated as above

= w(131. , 01

.5

0.4652

0.465

1.0

0.3183

0.318

2.0

0.1292

0.129

4.0

0.063

0.0625

G(o)

0

0.500

0.500

2. Stresses

For vertical stresses, the fiction

377.

(lz ) F 13 + 17 04, F (z + )-I- 2log +

- I F + -1T),X F (4.3),4- L

' -1* ZZ = — 1 b + 2 + + ab b

behaves like a series of curves shown in Fig. C.2 (for = 1)

aidcl- confirm the pattern of approaching identical values in the

limits —> 0 and 00 . Even though the values of ZZ for

the intermediate range of may be much higher than the limiting

values, particularly with increasing 04. , the quantity Z

in the integral governs the convergence of the series and

the effect of ZZ is no more than marginal. Therefore the

stresses for any value of are not significantly different

from their limiting values.

(3/6 oo Circular Load : Surface settlement _.

co 4 G(0) Jo (4)) -1 1 (c() (0,0) dor, w • -,-,-1 io.o

5.0

qb 0( A

j

0

\ 1.0

0.5

o.1 I , . 1

/..---------- 7

c< to

1 , 1 i 1 1 I I 1 1

1

0-8

06

-I< 0-4

0-2 ZS

0

0

-0.2

-0-

FIG C•1 CALCULATION OF SETTLEMENT IN NON-HOMOGENOUS SOIL MEDIA.

..IIIIi-MI 4

Limiting P/ =o values !M..° z/b =1.0

,

3

2

i

35

30

25

20 c N N

15

10

•01 1 1 10 A/a

F I G C•2 CALCULATION OE S TRESSES IN-NON HOMOGENEOUS SOIL MEDIA : DISTRIBUTION OF FUNCTION ZZ (0() --

378.

APPENDIX D

LEAKAGE IN TRIAXIAL TESTS

D.1 Effect of leakage into saturated undrained specimens

Let us take a saturated specimen which is enclosed in a

rubber membrane and consolidated in the triaxial cell to an

effective stress (C7- ')0. Assuming that all drainage valves are 3

now closed, water leakage'into the specimen will cause the total

volume and the pore water pressure to increase and the effective

stress to decrease. Strictly speaking the net increase of volume

of the specimen will be equal to

(1) the volume of leakage, minus

(2) the volume change of the pore water, minus

(3) the volume change of the soil grains, plus

(4) volume change inside the membrane due to expansion.

Poulos (1964) has shown that for clays the quantities (2), (3) and

(4) are negligible in comparison with the volume of leakage.

Fig. D.1 shows the effect of this leakage on the effective

stresses of the specimen. The volume of leakage tsv causes the

effective stresses to decrease by an amount A u. Assuming the

initial swelling curve, for small volume changes, to be a straight

line, the swelling ratio S is defined as:

v A u S. ( 07 )

3 0

(D.1)

379.

S, is, therefore, the slope of the initial swelling curve plotted

on a non-dimensional basis.

The initial swelling ratio of the Ongar clay has been

determined (Test no. T-H0-18) by the method proposed by Poulos.

A specimen was set up in the triaxial apparatus under a

cell pressure of 60 lbs/in2 and a deviator stress of about 50% of

that at failure applied under undrained conditions. The pore

pressure (i.e. the back pressure) was then increased in steps of

5%, 10% and 20% (cumulative) of the effective lateral stress

allowing full swelling to take place under each increment. The

results are shown in Fig. D.2. Assuming the swelling curve to be

Au a straight line in the range 0 '\ < 5 the initial swelling (

ratio is 0.090. This is much higher-tan the values for other

clays obtained by Poulos, such as Bearpaw clay shale (0.03) and

Boston Blue clay (0.015).

So the maximum permissible leakage which will cause a

change in effective stress of not more than 2% is given by

= 0.09 x .02 = .0018

for a sample 1-i" dia. x 3" high v = 88 c.c.

Therefore Llv = 0.0018 x 88 = .158 c.c. = 158 m.m.3

For a test lasting 30 days this means a leakage rate of 5.3

m.m.3/day.

Dv

380.

The above analysis strictly applies to undrained tests

for the particular stress level used in the pilot test. These

stresses were, however, similar to those for most of the tests

reported in this thesis.

D.2 Amount of leakage through membranes

(a) Due to hydraulic pressure gradient

Assuming the validity of Darcy's Law (Poulos 1964), the

rate of flow through a porous medium,

q =KiA

where

K = coefficient of permeability (cm/sec)

i = hydraulic gradient

A = area of flow (cm2)

Taking the pressure difference between the inside and out-

side of the membrane (0.016 in. thick in the present series of tests)

as 100 lbs/in2, the area of a sample (li" dia. x 3" high) equal to

91.2 cm2 and K = 5 x 10-1 6cm/sec for natural rubber, as determined

by Poulos (1964),

—16 100 x 70.3 q = 5 x 10 x x 91.2 0.016 x 2.54

= 0.79 x 10-8 cm3/sec

381.

= 0.68 mm3/day

(b) Due to osmotic pressure difference

Assuming the pore space was filled with sea water (i.e.

high salt concentration) Poulos (1964) determined the rate of flow

through a membrane 0.006 cm thick as equal to

0.19 mm3/day/cm2

Using Poulos' data for the relationship between flow and

membrane thickness the rate of flow through a membrane 0.016" thick

into a sample 12" dia. x 3" high

= 0.19 x 20 x 91.2 mm3/day 350

= 0.99 mm3/day

Therefore, the maximum possible leakage for the above

conditions, due to the combined effect of hydraulic and osmotic

pressure differences

= 0.68 + 0.99 = 1.67 mm3/day

382.

D.3 Leakage through connections and creep of saran tubing

A special test was run with the set-up shown in Fig. 9.13

- except that, instead of the cell assembly and the sample, the end

of the top drainage cap was sealed by clamping it against an

aluminium disc with a Dowty seal. A pressure of 60 lbs/in2 was

applied and maintained for 30 days by the self-compensating mercury

control. After the air bubble had stabilised under the initial

compression the displacement was measured regularly. The results

are shown in Fig. D.3.

It will be seen that the maximum volume change in 30

days for a back pressure of 60 lbs/in2 would be (@ 21.6 in/c.c.)

1.6 x 1 - 0.075 c.c. 21.6

Assuming that leakage is directly proportional to pressure,

this would mean, for a typical test with a back pressure of 20 lbs/in2

a total leakage of 0.023 c.c.

Initial Swelling Ratio's' is given by

i

AV AU =5 V (q)0

.\ Compression

Swelling .

V f AV

----1.4 A U 1-4---

( ai) 0

w

n -J 0 >

Effective Stress

FIG. D1 EFFECT OF WATER LEAKAGE ON EFFECTIVE STRESSES IN SATURATED SAMPLES DURING TRIAXIAL TESTS.

(After Poulos1964)

5 0 15 10 20

4. au 0/0 ( (5Do

FIG. D•2 DETERMINATION OF INITIAL SWELLING RATIO OF LONDON CLAY FROM ONGAR.

I I I I II I I I

-•

- 11 Initial

(0-3).=

for

Swelling Ratio

0<A U

<510

s=0.09

T- HO-18 1 1

( 15-3')0 41 psi

I I 1 I I 1

I I I I I I .

1.0

0.8

0.2

0

0.6

Ay / , + •••-- 10 v

0.4

0 10 20 30

2.0

1.0

0

Time (Days)

FIG. D3 LEAKAGE THROUGH CONNECTIONS AND CREEP OF SARAN TUBING ( Pressure :60 p.s.i)

1 1 i i 1 1

383.

APPENDIX E

TIME - SETTLEMENT DIAGRAMS FOR STANDARD OEDOMETER TESTS

A few typical cases are shown Ion semi-log plots in Figs.

E.1 - E.10. The following notations have been used to denote the

three values of the zero reading:

1 Initial Reading on the dial gauge before load application

2 - Zero reading correcting for apparatus deformation only

3 - Zero reading obtained by Casagrande construction from the

shape of the early portion of each curve (see Taylor 1948).

Time ( Minutes )

FIG. El : 0-H0-2 1 1st LOADING.

5 000 100 10 1000

Time ( Minutes )

FIG. E2 : 0-H0-21 2nd Loading.

1 100 1000 5000 10

0

•-•••-__....____._.

- 4

0 8 • 4-

(1) • 8-16

---...

16-32

Tift2

230

220

210 0

1600

1500

1400

,•••••••,,,

c -4.

190

1800

a! a

1700

200

0-

8-16

16-32

0.1 3300

1

Time ( Minutes )

10 100 1000 5000

o 3200

0 ® --•

1 - 2 3100

• • 2 -4

2 900 .,,••••••

c .1.

t o 2 800

rn c .....

1:7 m 0 cc 2700

0

—co

• 4 -8

0

2600

25 00

2400

2300

2200

2100

T / ft 2

FIG. E3 : 0 -H0-8 / 1st LOADING.

3000

Time ( Minutes )

FIG. E4 : 0 -H0-8 / 2nd LOADING.

5000 100 1000 10 1

0 °Awe

---4------.---.7„.

- 2

2 - 4

1

0

4 - 8

-16 8

o0

16 -32

T/ft2

0.1 2900

2800

2700

2600

a c -0 m 0 2400 rx -

fa

c)

2300

2200

2100

2000

........

C -4. 'cp 2500 ,

Time ( Minutes )

0.1 3400

10

100

1000 5000

6

3300

32 00

3100

3000

-4

4 - 8

2900 o

2800

2700

8 -16

2600

2500

2400

2300

®

16 -32

Tift 2

2200

FIG. ES : 0-H0-9/ 1st LOADING.

Time (Minutes)

FIG. E6: 0-H0-9 / 2nd LOADING.

C

10

2700

310 0

300

290

280

2400

2300

2200

rd

a 2 500

cn C

2600

cc

J•1 1 10 100 1000 5001

oa

2 -4

-1- 1

0 ._ 4- 8

0 _

8-16

0 0

t

16-32

T/ft 2

0.1 3500

1 Time (Minutes)

10 100 1000 5000

3450

—.. 3400 c

..t l o ..„...

cnc 3350 17 MI C9 tx

rti 330 0

0

3

3 -6

3250

T/ft 2

6-12 •

3200

FIG. E7 : 0-0C- 4 / 1st LOADING.

3350

3300

3250

.t s o 3200

cn c -0 n3 u cc 3150

al

0

3100

3050

3000

34000.1

o

Time ( Minutes )

10 100 1000 5000

sci ®

4 -12

12 -24

M a

2950

41.---........

T/ft 2

24-32

2900

FIG. E 8 : 0-0C-4 / 2nd LOADING.

FIG. E 9 : 0-0C-9 /1st LOADING.

100

Time ( Minutes )

10 0-1 1 190

180

170

0:9 160 CC

150

140 0

130 0

0

3 - 6

0

)

6 r --L'a--------„_. -12

o

i

0 12-24

O

24 -32

T/ft 2

SnO0 ___

rn 1500

tx

1400

1300

Time (Minutes)

0.1

1

10

100

1000

5000

U

--...il ---

. 3-6 •

-----T

1 6-12

.--

0

7

_1 12-24

0

24-32 T/ft2

FIG. E10: 0-0C-9/ 2nd LOADING.

1700

1600

384.

APPENDIX F

EFIWCT 0- SIDE FRICTION ON THE MEASURED

LATERAL STRESSES IN STRAIN GAUGE OEDOMETERS

An upper limit of the errors caused by side friction on

the measured lateral stresses can be found by considering an element

of the oedometer ring of unit length and replacing the total

frictional force above the level of the strain gauges by an eccentric

force trii/2 acting on top of the ring as shown in Fig. F.1.

where .77 = side friction per unit area of the surface of the

ring - assumed uniformly distributed and

H = height of the ring.

(a) If there is no side friction the ring is acted upon by the

lateral stress 0rr only causing a hoop stress 0-9®_ Crrr( r/t)

where r = radius of the ring

t = thickness of the ring

(From the theory of thin circular rings - see Timoshenko 1955).

This is the condition for which the ring is calibrated

(with water pressure), the measured quantity being the hoop tension

Qg given by

C 99 = 009/E

(F.1)

(b) When friction is present and is replaced by the axial stress

385.

and a bending movement as described above, the hoop tension

is given by

I t\ -t Q9 ecH -/ + gg _

1 2 E 2t E z . t

e QQ1

(IGO H) + -- -

CrGG 1:11 I)

E E t

E E 2t E \,2 t

(F.2)

where and E are respectively the Poisson's ratio and Young's

modulus of the material of the ring - in this case brass.

(The above assumes that the side friction does not alter the hoop

stress). So, the error caused by the assumption of no friction,

from equations (F.1) and (F.2), is,

c H Q9 GQ t

GG 0- QG

(F.3)

Now, Cr GG rr t

r-C =

/' 0- rr

where JOL is the coefficient of friction between the ring and the

soil specimen and will be taken to be 0.2, a figure certainly on the

high side.

386.

Substituting in equation (F.3),

98 - 09

C90

rr t

(rrr(

•/L. H (F.4)

Putting the values for the test rings:

=

_ 09

/4-=

G9

0.2, H = 1 in and r = 1.5 in

- 0.044 3(0.2) -- 1 .15 99

So the maximum possible error in the lateral stress measurement, due

to the effect of side friction, is less than 5%. It must be

emphasized that this error constitutes an upper limit because the

above analysis assumes a free standing column and does not take

into account the restraining effect of the circular ring, a rigorous

analysis of which is complicated.

H

FIG. Fl SIDE FRICTION IN OEDOMETER TESTS

387.

REFERENCES

Abbreviations

Trans. ASCE - Transactions of the American Society of Civil Engineers.

J. ASCE SMFD - Journal of the American Society of Civil Engineers, Soil Mechanics and Foundation Divisions

ASTM. STP - American Society for Testing and Materials, Special Technical Publication

C.G.J. - Canadian Geotechnical Journal

Proc. ICE - Proceedings of the Institution of Civil Engineers, London

ICSMFE - International Conference on Soil Mechanics and Foundation Engineering

IUTAM - International Union of Theroetical and Applied Mechanics

H.R.B. - Highway Research Board, Washington, D.C.

H.R.R. - Highway Research Record, published by the Highway Research Board.

N.G.I. - Norwegian Geotechnical Institute

388•

Abbott, A. M. (1960) - "One-Dimensional Consolidation of Multi-. Layered Soils"rGeotechnique, Vol. 10, No. 4,13. 151.

Acura, W. and Fox, L. (1951) - "Computation of Load Stresses in a Three Layer Elastic System".-Geotechnique, Vol. 2, p. 293.

Agarwal, K. B. (1967) - "The Influence of Size and Orientation of Sample oil the Undrained Strength of London Clay" - Ph.D. Thesis, University of London.

Ahlvin, R. G. and Ulery, H. H. (1962) - "Tabulated Values for Determining the Complete Pattern of Stresses, Strains and Deflections Beneath a Uniform Circular Load on a Homogeneous Half Space" - H.R.B. Bulletin No. 342, p. 1.

Akai, K. and Adachi, T. (1965) - "Study of One-Dimensional Consolida-tion and the Shear Strength Characteristics of Fully Saturated Clay in Terms of Effective Stress" - Proc. 6th ICSMFE, Montreal, Vol. 1, p. 146.

Alderman, J. K. (1956) - Discussion of Ref: Skempton and McDonald (1955), Proc. ICE, Vol. 5, Part 1, p. 169.

Bara, J. P. and Hill, R. R. (1967) - "Foundation Rebound at Dos Amigin Pumping Plant" - J. ASCE, SMFD, Vol. 93, SM 5, September 1967.

Barber, (1940) - see De Barros (1966).

Barden, L. (1963) - "Stresses and Displacements in Cross-Anisotropic Soil" - Geotechnique, Vol. 13, No. 3.

Barden, L. (1965) - "Consolidation of Normally Consolidated Clays" -J. ASCE SMFD, Vol. 91, SM 5, September 1965.

Barron, R. A. (1948) - "Consolidation of Fine Grained Soils by Drain Wells" - Trans. ASCE, Vol. 113, 1948.

act, M. A. (1935) - "The Effect of Certain Discontinuities on the Pressure Distribution in a Loaded Soil" - Physics, Vol. 6, No. 12, p. 367.

Biot, M. A. (1941) - "General Theory of Three-Dimensional Consolidation" - J. of Applied Physics, Vol. 12, p. 1557164.

Biot, M. A. (1941a) - "Condolidation Settlement Under a Rectangular Load Distribution" - J. of Applied Physics, Vol. 12, p. 426.'430.

389.

Biot, M. A. (1955) - "Theory of Elasticity and Consolidation for a Porous Anisotropic Solid" - J. of Applied Physics, Vol. 26, p. 182-185.

Biot, M. A. and Elingan, F. M. (1941) - "Consolidation Settlement of a Soil with an Impervious Top Surface" - J. of Applied Physics, Vol. 12, p. 578-581.

Biot, M. A. and Willis, D. G. (1957) - "The Elastic Coefficients of the Theory of Consolidation" - J. of Applied Mechanics, Vol. 24, No. 4, p. 594-601.

Binnie, G. M. and Price, J. A. (1941) - "An Apparatus for Measuring the Lateral Pressure of Clay Samples under Vertical Load" - J. ICE, Vol. 15, p. 297.

Bishop, A. W. (1947) - "The London Clay, Part III - Strength Variations in London Clay" - Journees Internationales Consacrees a l'Etude ... Silicates Industriels, Vol. 13, No. 9.

Bishop, A. W. (1952) - "The Stability of Earth Dams" - Ph.D. Thesis, University of London.

Bishop, A. W. (1954) - "The Use of Pore Pressure Coefficients in Practice" - Geotechnique, Vol. 4, p. 148-152.

Bishop, A. W. (1957) - "Some Factors Controlling the Pore Pressures 43eitt 145 during Construction of Earth Dams" - Proc. 4th ICSMFE, London, Vol. 2, p. 294-300.

Bishop, A. W. (1958) - "Test Requirements for Measuring the Coefficient of Earth Pressure at Rest" - Proc. Brussels Conf. on Earth Pressure Problems, Vol. 1, p. 2.

Bishop, A. W. (1958a) - Contribution to Discussions, Brussels Conf. on Earth Pressure Problems, Vol. 3, p. 38.

Bishop, A. W. (1959) - "The Principle of Effective Stress" - N.G.I. Publication No. 32.

Bishop, A. W. (1966) - "Strength of Soils as Engineering Materials" - 6th Rankine Lecture, Geotechnique, Vol. 16, June 1966.

Bishop, A. W. (1966a) - "Soils and Soft Rocks as Engineering Materials" - Inaugural Lecture, Imperial College, May 1966.

390.

Bishop, A. W. and Henkel, D. J. (1953) - "Pore Pressure Changes during Shear in two Undisturbed Clays" - Proc. 3rd ICSMFE, Zurich, Vol. 1, p. 94.

Bishop, A. W. and Donald, I. B. (1961) - "The Experimental Study of Partly Saturated Soils in the Triaxial Apparatus" - Proc. 5th ICSMFE, Paris, Vol. 1, p. 13-21.

Bishop, A. W., Alpan, I., Blight, G. E. and Donald, I. B. (1960) -"Factors Controlling the Strength of Partly Saturated Cohesive Soils" - Proc. Res. Conf. Shear Strength of Cohesive Soils, Boulder, p. 503.

Bishop, A. W. and Henkel, D. J. (1962) - "The Measurement of Soil Properties in the Triaxial Test" . 2nd Ed. Edward Arnold, London.

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, Vol. 15, No. 1, p. 1-31.

Bishop, A. W., Webb, D. L. and Skinner, A. E. (1965) - "Triaxial Tests on Soil at Elevated Cell Pressures" - Proc. 6th ICSMFE, Montreal, Vol. 1, p. 170-174.

Bishop, A. W. and Little, A. L. (1967) - "The Influence of the Size and the Orientation of the Sample on the Apparent Strength of the London Clay at Malden, Essex" - Proc. Geo. Conf. Oslo, Vol. 1, p. 89-96.

Bishop, A. W. and Vaughan, P. R. (1962) - "Selset Reservoir - Design and Performance of the Embankment" - Proc. ICE, Vol. 21, p. 305.

Bjerrum, L. (1954) - "Theoretical and Experimental Investigations on the Shear Strength of Soils" - N.G.I. Publication No. 5.

Bjerrum, L. (1964a) - "Relation Between Observed and Calculated Settle-ments of Structures in Clay and Sand" - (In Norwegian) - N.G.I. Oslo, October 1964.

Bjerrum, L. (1964b) - "Secondary Settlement of Structures" - Proc. IUTAM Symposium, GrenOble, 1964. Also N.G.I. Publication No. 73.

Bjerrum, L. (1965) - "Mechanism of Progressive Failure in Slopes of Over-consolidated Plastic Clay and Clay Shales" - 3rd Terzaghi Lecture, J. ASCE SMFD, Vol. 93, SM 5.

391.

Bjerrum, L. (1967) - "Engineering Geology of Normally Consolidated Marine Clays as related to Settlement of Buildings" - 7th Rankine Lecture, Geotechnique, Vol. 17, dune 1966.

Bjerrum, L., Simons, N. and Torblia, (1958) - "The Effecto of Time on the Shear Strength of Soft Marine Clay"- Proc. Brussels Conf. on Earth Pressure Problems, Vol. 1, p. 148-158.

Blight, G. E. (1963) - "The Effect of Non-uniform Pore Pressures on Laboratory Measurements of Shear Strength of Soils" - ASTM/NRC Symposium on Laboratory Shear Testing of Soils, Ottawa, ASTM STP No. 361.

Bozuzuk, M. (1963) - "The Modulus of Elasticity of Leda Clay from Field Measurements" - CGJ, Vol. 1, No. 1.

Brinch Hansen, J. (1961) - See Hansen and Misi (1961).

Broms, B. B. and Ratnam, M. V. (1963) - "Shear Strength of Aniso-tropically Consolidated Clay" - J. ASCE SMFD, Vol. 89, No. SM 6, p. 1.

Brooker, E. W. and Ireland, H. 0. (1965) - "Earth Pressure at Rest Related to Stress History" - CGJ, Vol. 2, No. 1, Feb. 1965.

Buckton, E. J. and Currel, J. (1943) - "The New Waterloo Bridge" - J. ICE, Vol. 20, No. 7.

Buckton, E. J. and Fereday, H. J. (1938) - "The Reconstruction of Chelsea Bridge" - J. ICE, Vol. 7, p. 383.

Buchan, S. (1938) - "The Water Supply of the County of London" -Memoirs of the Geol. Survey, London, HMSO.

Burmister, D. M. (1943) - "Theory of Stresses and Displacements in Layered Systems" - Proc. HRB, Vol. 22, p. 127-148.

Burmister, D. M. (19W - "The General Theory of Stresses and Dis- placements in Layered Soil Systems" - J. of Applied Physics, Vol. 16, No. 2, p. 89-96, No. 3, p. 126-127, No. 5, p. 296-302.

Burmister, D. M. (1952) - "The Application of Controlled Test Methods in Consolidation Testing" - ASTM, Spl. Tech. Pub. No. 126.

Burmister, D. M. (1956) - "Stress and Displacement Characteristics of a Two Layer Rigid Base Soil System" - Proc. HRB, Vol. 35.

392.

Burn, K. N. (1965) - Discussion on Klohn (1965) - CGJ, Vol. 2, p. 341.

Cappleman, H. L. (1967) - "Movements in Pipe Conduits under Earth Dams" - J. ASCE SMFD, Vol. 93, SM 6.

Carothers, S. D. (1920) - "Plane Strain: Direct Determination of Stresses" - Proc. Royal Soc. Series A : Vol. 97, p. 110-123.

Casagrande, A. (1936) - "The Determination of Pre-consolidation Load and its Practical Significance" - Proc. 1st ICSMFE, Cambridge, Mass., Vol. 3.

Casagrande, A. and Wildon, S. D. (1951) - "Effect of Rate of Loading on the Strength of Clays and Shales at Constant Water Content"-Geotechnique, Vol. 2, p. 251-263.

Cebertowicz and Wedzinski, (1958) - "Relation advenant entre les Tensions Internes Horizontales et Verticales du Sol" - Brussels Conf. on Earth Pressure Problems, Vol. 1, p. 28.

Chandler, R. J. (1967) - "The Strength of a Stiff Silty Clay" -Proc. Geo. Conf., Oslo, Vol. 1,R.103.

Christie, I. F. (1963) - "The Consolidation of Soils" - Ph.D. Thesis, Edinburgh University.

Christie, I. F. (1965) - "Secondary Compression Effects During One- Dimensional Compression Tests" - Proc. 6th ICSMA, Montreal, Vol. 1.

Clayton, K. M. (1964) - "Geology and Geomorphology - Guide to London Excursions" - 20th Int. Geographical Congress, London.

Cooling, L. F. and Skempton, A. W. (1942) - "A Laboratory Study of London Clay" - J. ICE, Vol. 17, No. 3, p. 251.

Cooling, L. F. and Gibson, R. E. (1955) - "Settlement Studies of Structures in England" - Conf. Corr. Cale. and Obs. Stresses and Displacements of Structures, Vol. 1, p. 295.

Creasey, L. R., Adams, H. C. and Lampitt, N. (1965) - "Museum Radio Tower" - Proc. ICE, Vol. 30, p. 33-78.

Cornforth, D. H. (1964) - "Some Experiments on the Influence of Strain Conditions on the Strength of Sand" - Geotechnique, Vol. 14, No. 2.

393.

Crawford, C. B. (1963) - "Pore Pressures within Soil Specimens in Triaxial Compression" - ASTM, Spl. Tech. Pub. 361.

Crawford, C.B. (1964) - "Interpretation of Consolidation Tests" - J. ASCE SMFD, Vol. 90, SM 5.

Crawford, C. B. (9165) - "Resistance of Soil Structures to Consolida-tion" - CGJ, Vol. 2, No. 2.

Crisp, J. (1953) - Discussion on Schmertmann (1953) - Trans. ASCE, Vol. 120, p. 1228.

Davis, E. H. and Poulos, H. G. (1963) - "Triaxial Testing and Three-Dimensional Settlement Analysis" - Proc. 4th Aust.-N.Z. Conf. on Soil Mechanics, p. 233.

Davis, E. H. and Poulos, H. G. (1966) - "The Analysis of Settlement under Three-Dimensional Conditions" - Research Report No. R.65, Univ. of Sydney.

Davis, E. H. and Poulos, H. G. (1968) - "The Use of Elastic Theory for Settlement Prediction under Three-Dimensional Conditions" Geotechnique, Vol. 18, No. 1, p. 67.

Davis, E. H. and Raymond, G. P. (1965) - "A Non-linear Theory of Consolidation" - Geotechnique, Vol. 15, No. 2, p. 161.

Dawson, R. F. (1940) - "A Comparison between Results of Loading Tests, Settlement Analysis based on Consolidation Tests and Observed Settlements" - Proc. Purdue Conf. on Soil Mechanics and its Application.

Dawson, R. F. and Simpson, W. E. (1948) - "Settlement Records of Structures in Texas Gulf Coast Area" - Proc. 2nd ICSMFE, Rotterdam, Vol. 4, p. 125.

De Barros, S. T. (1966) - "Deflection Factor Charts for Two and Three Layer Systems" - HRR, No. 145, p. 83.

De Beer, E. E. (1963) - "The Scale Effects in the Transposition of the Results of Deep Sounding Tests on the Ultimate Bearing Capacity of Piles and Caisson Foundations" - Geotechnique, Vol. 13, No. 1.

Delory, F. A., Gass, A. A. and Wong, W. W. (1965) - "Measured Pore Pressures used for the Control of Two Stage Construction of an Embankment" - GCJ, Vol. 2, p. 216.

394.

De Wet, J. A. (1962) - "Three Dimensional Consolidation" - HRB. Bulletin No. 342, p. 152.

Eggestad,A.(1963) - "Deformation Measurements below a Model Footing on the Surface of a Dry Sand" - Proc. European Corif. on Settle-ment Problems, Wiesbaden, Vol. 1, p. 233.

Escario, V. and Uriel, S. (1961) - "Determining the Coefficient of Consolidation and Horizontal Permeability of Radial Drainage" - Proc. 5th ICSMFE, Vol. 1.

Fadum, R. E. (1948) - "Influence Values for Estimating Stresses in Elastic Foundations" - Proc. 2nd ICSMFE, Vol. 3, p. 77.

Fraser, A. M. (1957) - "The Influence of Stress Ratio on Compressi-bility and Pore Pressure Coefficients in Compacted Soils" -Ph.D. Thesis, University of London.

FrostiA.D.and Mason, J. (1963) - "Stag Place Development" - The Structural Engineer, Vol. 41, p. 347.

Fookes, P. (1966) - Letter to Geotechnique, Vol. 16, p. 261.

Fox, L. (1948) - "The Mean Elastic Settlement of a Uniformly Loaded Area at a Depth below Ground Surface" - Proc. 2nd ICSMFE, Rotterdam, Vol. 1, p. 129.

Geresvanoff, N. (1936) - "Improved Methods of Consolidation Tests and Determination of the Capillary Pressure in Soils" - Proc. 1st ICSMFE, Vol. 1, p. 47-50.

Gibson, R. E. (1958) - "The Progress of Consolidation in a Clay Layer increasing in Thickness with Time" - Geotechnique, vol. 8, No. 4.

Gibson, R. E. (1963) - "An Analysis of System Flexibility and its Effect on Time Lag in Pore Water Pressure Measurements" -Geotechnique, Vol. 13, No. 1:

Gibson, R. E. (1967) - "Some Results Concerning Displacements and Stresses in Non-homogeneous Elastic Half Space" - Geotechnique, Vol. 17, No. 1.

Gibson, R. E. (1968) - Better to Geotechnique, Vol. 18, No. 2.

Gibson, R. E. and Lumb, P. E. (1953) - "Numerical Solutions to Some Problems in the Consolidation of Clay" - Proc. ICE, Vol. 2, pt. 1, p. 182.

395.

Gibson, R. E. and McNamee, J. (1957) - "The Consolidation Settlement of a Load Uniformly Distributed over a Rectangular Area" - Proc. 4th ICSMFE, Vol. 1, p. 297.

Gibson, R. E. and McNamee, J. (1960)'- "Plane Strain and Axi-symmetric Problems of the Consolidation of a Semi-Infinite Clay Stratum" - Qtly. J. of Mech. and App. Maths, Vol. 13, p. 210.

Gibson, R. E. and McNamee, J. (1963) - "A Three-Dimensional Problem of the Consolidation of a Sea-Infinite Clay Stratum" - Qtly. J. of Mech.-and Appl. Maths, Vol. 16, p. 115.

Gibson, R. E., Hussey, G. L. and England, M. J. L. (1967) - "The Theory of One-Dimensional Consolidation of Saturated Clays" -Geotechnique, Vol. 17, No. 3.

Gibson, R. E. and Marsland, A. (1960) - "Pore Water Pressure Observa-tions in a Saturated Alluvial Deposit beneath a Loaded Oil Rink" - Conf. on Pore Pressure and Suction in Soils, Butterworths, London.

Golecki, J. (1959) - "On the Foundation of the Theory of Elasticity of Plane Incompressible Non-homogeneous Bodies" - Proc. IUTAM Symposium, Pergamon Press.

Gray, H. (1945) - "Simultaneous Consolidation of Contiguous Layers of unlike Compressible Soils" - Trans. ASCE, Vol. 110, p. 1327.

Hantal, W. S. (1950) - "Settlement Studies in Egypt" - Geotechnique, Vol. 2, p. 33.

Hanna, W. S. (1953) - "Settlement of Buildings on Pre-Consolidated Clay Layers" - Proc. 3rd ICSMFE, Zurich, Vol. 1, p. 366.

Hansbo, S. (1960) - "Consolidation of Clay Layers with Special Reference to the Influence of Vertical Sand Drains" - Royal Swedish Geo. Inst., Proc. No. 18.

Hansen, J. B. and Misi, T. (1961) - "An Empirical Evaluation of Consolidation Tests with Little Belt Clay" - Danish Geo. Institute, Bull. No. 17.

Hardy, R. M. and Ripley, C. F. (1961) - "Horizontal Movements Assoc-iated with Vertical Settlements" - Proc. 5th ICSNFE, Paris, Vol. 1, p. 665.

396.

Harr, M. (1966) - "Foundations of Theoretical Soil Mechanics" -McGraw-Hill, New York.

Haythornthwaite, R. M. (1960) - "The Shear Strength of Saturated Remoulded Clays" - "Plasticity", Pergamon Press, New York.

Helendlund, K. V. (1953) - "Settlement Observations in Finland" -Proc. 3rd ICSMFE, Zurich, Vol. 1, p. 370.

Hendron, A. J. (9163) - "The Behaviour of Sand in One-Dimensional Compression," - Tech. Report No. RTD-TDR-63-3089: USAF.

Henkel, D. J. (1958) - "The Correlation between Deformation, Pore Water Pressure and Strength Characteristics of Saturated Clay" - Ph.D. Thesis, University of London.

Henkel, D. J. (9160) - "The Shear Strength of Saturated Remoulded Clays" - Proc. Res. Conf. on Shear Strength of Cohesive Soils, Boulder, p. 533.

Henkel, D. J. (1960a) - "The Relationships between Effectit.e Stresses and Water Content in Saturated Clays" - Geotechnique, vol. 10, No. 1.

Henkel, D. J. and Sowa, V. A. (1963) - "The Influence of Stress History on Stress Paths in Undrained Triaxial Teats on Clay" - ASTM, STP No. 361.

Henkel, D. J. and Wade, N. H. (1966) - "Plane Strain Tests on Saturated Remoulded Clay" - J. ASCE SMFD, Vol. 92, SM 6, p. 67.

Hooper, J. A. and Butler (1966) - "Some Numerical Results Concerning the Shear Strength of London Clay" - Geotechnique, Vol. 16, No. 4.

Hruban, K. (1948) - "Sources of Error in Settlement Estimates" - Proc. And ICSMFE, Rotterdam, Vol. 1, p. 127.

Hruban, K. (1959) - "The Basic Problem of a Non-Linear and a Non-Homogeneous Half Space" - Non-Homogeneity in Elasticity and Plasticity, Pergamon Press.

Hvorslev, M. J. (1937) - "Uber die festigkeitsei-genschaften gesttirter bindiger &Men" - Ingeniorvidenskabelige Skrifter A.No. 45, Copenhagen.

Hvorslev, M. J. (1960) - "Physical Components of the Shear Strength of Saturated Clays" - Res. Conf. on Shear Strength of Cohesive Soils, Boulder, p. 169.

397.

Huang, Y. H. (1968) - "Stresses and Displacements in Non-Linear Soil Medium" - J. ASCE SMFD, Vol. 94, SM 1.

Jackson, W. Y. (1964) - "Stress Paths and Strains in Saturated Clay" - M.Sc. Thesis, M.I.T., unpublished.

Jakobson, B. (1957) - "Some Fundamental Properties of Sand" - Proc. 4th ICSMFE, London, Vol. 1, p. 167.

Janbu, N. Bjerrum, L. and Kjaernsli, B. (1964) - "Soil Mechanics Applied to Some Engineering Problems" - (in Norwegian) : N.G.I. Pub. No. 16.

Jahnke and Emde (1943) - "Tables of Functions".

Jones, A. (1962) - "Table of Stresses in Three-Layer Elastic Systems" H.R.B. Bulletin, No. 342.

de Josselyn de Jong, J. (1957) - "Application of Stress Functions to Consolidation Problems" - Proc. 4th ICSMFE, London, Vol. 1, p. 320.

Jurgenson, L. (1934) - "Application of Theories of Elasticity and Plasticity to Foundation Problems" - J. Boston Soc. of Civil Eng., Vol. 21, No. 3.

Kenney, T. C. (1967) - "Field Measurements of in-situ Stresses in Quick Clays" - Proc. Geo. Conf., Oslo, Vol. 1, p. 49.

Kerisel, J. and Quatre, M. (1968) - "Settlement Under Foundations - Calculations using Triaxial Apparatus" - Civil Eng. and Public Works Review : Part I (May 1968) and Part II (June 1968).

Kirk, J. M. (1966) - "Tables of Radial Stresses in Top Layer of Three Layer Elastic Systems at Distance from Load Axis" - H.R.R. No. 145, (H.R.B.)

Kirkpatrick, W. M. (1957) - "The Condition of Failure of Sands" - Proc. 4th ICSMFE, London, Vol. 1, p. 172.

Kjellman, W. (1936) - "Report on an Apparatus for Consummate Investiga-tions of the Mechanical Properties of Soils" - Proc. 4th ICSMFE, London, Vol. 2, p. 16.

Kjellman, W. and Jakobson, B. (1955) - "Some Relations between Stress and Strain in Coarse Grained Cohesionless Materials" - Proc. Roy. Swedish Geo. Inst. No. 9.

398.

KlBhn, E. J. (1965) - "The Elastic Properties of a Dense Glacial Fill Deposit" - CGJ., May 1965.

Ko, H. Y. and Scott, R. F. (1967) - "A New Soil Testing Apparatus" - Geotechnique, Vol. 17, p. 210.

Korenov,B. G. (1957) - "A Die Resting on an Elastic Half Space, the Modulus of Elasticity of which is an Exponential Function of Depth" - Dokl. Akad. Nank S.S.R., Vol. 112.

Ladd, C. C. (1964) - "Stress-strain Modulus of Clay in Undrained Shear" - J. ASCE SMFD, Vol. 90, SM 5.

Ladd, C. C. (996) - "Stress Strain Behaviour of Anisotropically Consolidated Soil during Undrained Shear" - Proc. 6th ICSMFE, Montreal, Vol. 1, p. 282.

Ladd, C. C. and Lambe, T. W. (1963) - "The Strength of 'Undisturbed' Clay Determined from Undrained Tests" - ASTM STP 361.

Lambe, T. W. (1961) - "Residual Pore Pressures in a Compacted Clay" - Proc. 5th ICSMFE, Paris, Vol. 1, p. 207.

Lambe, T. W. (1962) - "Pore Pressures in a Foundation Clay" - J. ASCE SMFD, Vol. 88, No. SM 2.

Lambe, T. W. (1964) - "Methods of Estimating Settlement" - J. ASCE SMFD, Vol. 90, SM 5.

Lambe, T. W. and Horn, H. M. (1965) - "The Influence on an Adjacent Building of Pile Driving for the M.I.T. Materials Centre" - Proc. 6th ICSMFE, Vol. 2, p. 280.

Lambe, T. W. (1965) - "Contribution to Discussions" - Proc. 6th ICSMFE, Montreal, Vol. 3, p. 355.

Lambe, T. W. (1967) - "The Stress Path Method" - J. ASCE SMFD, Vol. 93, SM 6.

Langer, K. (1936) - "The Influence of the Speed of Loading on the Pressure Void Ratio Diagram of Undisturbed Samples" - Proc. 1st ICSMFE, Cambridge, Mass., Vol. 2, p. 116.

Lee, K. L. and Farhoomand, T. (1967) - "Compressibility and Crushing of Granular Soil in Anisotropic Triaxial Compression" - CGJ., Vol. 4, No. 1.

399•

Lekhnitskii, S. G. (1962) - "Radial Distribution of Stresses in a hedge and in a Half Plane with Variable Modulus of Elasticity" - PMM, Vol. 26, No. 1.

Lemcoe, M. M. (1961) - "Stresses in Layered Elastic Solids" - Trans. ASCE, Vol. 126, p. 194.

Leonards, G. and Altschaeffl, A. G. (1964) - "Compressibility of Clays" - J. ASCE SMFD, Vol. 90, SM 5.

LeoKlards, G. and Ramiah, B. K. (1959) - "Time Effects in the Consolida-tion of Clays" - ASTM, Spl. Tech. Pub. No. 254.

Lewis, W. A. (1950) - "An Investigation of the Consolidation Test for Soils" - DSIR, Road Research Lab. Note No. RN/1349/WAL.

Lewis, W. A. (1963) - "The Effect of Vertical Sand Drains on the Settlement of a Road Embankment on the Thames Estuary" - Proc. European Conf. on Settlement, Wiesbaden, Vol. 1, p. 359.

Lo, K. Y. (1961) - "Stress Strain Relationships and Pore Water Pressure Characteristics of a Normally Consolidated Clay" - Proc. 5th ICSMFE, Paris, Vol. 1, p. 223.

Love, A. E. H. (1928) - "The Stress Produced in a Seipi-Infinite Solid by Pressure on part of the Boundary" - Phil. Trans. of the Roy. Soc., Series A, Vol. 228, p. 377.

Lowe, J., Zaccheo, P. F. and Feldman, H. S. (1964) - "Consolidation Testing with Back Pressure" - J. ASCE SMFD, Vol. 90, SM 4.

Lovenbury, H. T. (1968) - "Creep Characteristics of London Clay" -Ph.D. Thesis, University of London.

McKinley, D. G. (1961) - "A Laboratory Study of Rates of Consolidation in Clays with Special Reference to Conditions of Radial Pore Water Drainage" - Proc. 5th ICSMFE, Paris, Vol. 1.

Measor, E. 0. and Williams, G. M. J. (1962) - "Features in the Design and Construction of the Shell Centre, London" - Proc. ICE, Vol. 21, p. 475.

Meyerhoff, G. G. (1956) - Discussion of Ref. Skempton and McDonald, 1955 - Proc. ICE, Part 1, Vol. 5, P. 170.

Miller, R. J. and Low, P. F. (1963) - "Threshold Gradient for Water Flow in Clay Systems" - Proc. Soil Science Soc. of America;. Madison, Wis., Vol. 27, No. 6, p. 605.

400.

Mitchell, J. K. (1964) - "Shearing Resistance of Soil as a Rate Process" - J. ASCE SMFD, Vol. 90, SM 1.

Mitchell, J. K. and Campanella, R. G. (1968) - "Influence of Tempera-ture Variations on. Soil Behaviour" - J. ASCE SMFD, Vol. 94, SM 3.

Moretto, 0. (1965) - General Report on "Shear Strength and Consolida-tion" - Proc. 6th ICSMFE, Montreal, Vol. 3, p. 220.

Nakase, A. (1963) - "Side Friction in Conventional Consolidation Tests" - Report No. 3 : Port and Harbour Tech. Res. Inst., Japan.

Narain, J. and Singh, B. (1967) - "Role of Temperature in Triaxial Testing" - Proc. S.E. Asian Regional Conf. on Soil Eng., Bangkok, P. 103.

Newmark, N. M. (1942) - "Influence Charts for Computation of Stresses in Elastic Foundations" - Circular No. 24, Eng. Exp. Stn. Univ. of Illinois.

Nitchiporovitch, A. A. (1957) - "Results of Field Observation of Settlements of Large Hydraulic Struttures" - Proc. 4th ICSMIE, London, Vol. 1, p. 387.

Nooramy, I. and See, H. B. (1965) - "In-Situ Strength Characteristics of Soft Clays" - J. ASCE SMFD, Vol. 91, SM 2.

Northey, R. D. (1956) - "Rapid Consolidation Tests for Routine In-vestigation" - Proc. 2nd Aust. N.Z. Conf. on SMFE, Christchurch, p. 20.

Northey, R. D. and Thomas, R. F. (1965) - "Consolidation Test Pore Pressures" - Proc. 6th ICSMFE, Montreal, Vol. 1, p. 323.

Odenstad,S.(1949)- Geotechnique, Vol. 1, p. 242.

Osterman,J.and Lindskog,G.(1963) - "Influence of Lateral Movement in Clay upon Settlement in Some Test Areas" - Proc. Europ. Conf. on Settlement, Wiesbaden, Vol. 1, p. 137.

Paaswell, R. E. (1967) - "Temperature Effect on Clay Consolidation" -J. ASCE SMFD, Vol. 93, SM 3.

Parry, R. H. G. (1956) - "Strength and Deformation of Clay" T. Ph.D. Thesis, University of London.

401.

Parry, R. H. G. (196'0 - "Field and Laboratory Behaviour of Lightly Over-Consolidated Clay" - Geotechnique, Vol. 18, No. 2.

Peattie, K. R. (1962) - "Stress and Strain Factors for Three Layer Systems" - HRB Bulletin No. 342.

Peck, R. B. and Uyanik, M. E. (1955) - "Observed and Computed Settle-ment of Structures in Chicago" - Univ. of Illinois, Eng. Expt. Stn., Bulletin No. 24.

Peterson, R. (1954) - "Studies of Bearpaw Shale at a Dam Site in Saskatchewan" - Proc. ASCE SMFD, Vol. 80, Separate No. 476.

Peterson, R. (1958) - "Rebound in Bearpaw Shale in Western Canada" -Bull. Geol. Soc. of America, Vol. 69, p. 1113.

Phukan, A. L. T. (1968) - "Non-Linear Deformation of Rocks" - Ph.D. Thesis, University of London.

Pickett, G. (1938) - "Stress Distribution in a Layered Soil with Some Rigid Boundaries" - Proc. HRB, Vol. 18, Part 2, 1938.

Plantelpa. . (1953) - "Soil Pressure Measurements during Loading Tests on a Runway" - Proc. 3rd ICSMFE, Zurich, Vol. 1, p. 289.

Poulos, S. G. (1964) - "Control of Leakage in the Triaxial Test" - Harvard Soil Mechanics Series No. 71, Cambridge, Mass.

Poulos, H. G. (1967) - "Stresses and Displacements in an Elastic Layer Underlain by Rough Rigid Base" - Geotechnique, Vol. 17, No. 4.

Raymond, G. P. (1965) - "Contribution to Discussions" - CGJ, Vol. 2, p. 101.

Rendulic, L. (193) - "Der Hydrodynamische Spannungsansgleich in Zentral entwasserten Tozylindern" Wasserwitschaft und Teknik, Vienna, Vol. 2.

Rendulic, L. (1936) - 'Discussion of "Relation between Void Ratio and Effective Principal Stresses for Remoulded Silty Clay" - Proc. leb ICSMFE, Cambridge, Mass., Vol. 3, p.

Rendulic, L. (1937) - "Ein grundgesetz der tonmechanik und sein experimentaller beweis" - Bauingenieur, Vol. 18, p. 459.

1+02.

Resendiz, D., Nieto, J. A. and Figueroa, J. (1967) - "The Elastic Properties of Saturated Clays from Field and Laboratory Measure-ments" - Proc. 3rd PANAM. Conf. on S.M., Mexico, Vol. 1, p. 443.

Richart, F. E. (1957) - "A Review of Theories for Sand Drains" -J. ASCE SMFD.

Richardson, A. M. and Whitman, P. (1963) - "Effect of Strain Rate upon undrained Shear Resistance of a Saturated Remoulded Fat Clay" - Geotechnique, Vol. 13, No. 2.

Rivard, P. J. and Kohuska, A. (1965) - "Shehmouth Dam Test Fill" ON

CGJ., Vol. 2, p. 198.

Roberts, J. E. and De Souza, J. M. (1958) - "The Compressibility of Sand" - Proc. ASTM, Vol. 58, p. 1269.

Roscoe, K. A., Seoffield, A. N. and Wroth, C. P. (1958) - "On the Yielding of Soils" - Geotechnique, Vol. 8, No. 1.

Rowe, P. W. (1958) - "Contributions to Discussion" - Brussels Conf. on Earth Pressure Problems, Vol. 3.

Rowe, P. W. (1959) - "The Measurement of Coefficients of Consolidation of Lacustrine Clay" - Geotechnique, Vol. 9, No. 3.

Rowe, P. W. (1964) - "The Calculation of the Consolidation Rates of Laminated, Varved or Layered Clays, with particular Reference to Sand Drains" - Geotechnique, Vol. 14, No. 4.

Rowe, P. W. (1968) - "The Influence of Geological features of Clay Deposits on the Design and Performance of Sand Drains" - Proc. ICE, Paper No. 7058 S (Supp. Vol. 1968).

Rowe, P. W. and Barden, L. (1966) - "A New Consolidation Cell" -Geotechnique, Vol. 16, p. 162.

Rutledge, P. L. (1947) - "Relation of Undisturbed Sampling to Laboratory Testing" - Trans. ASCE, Vol. 109, p. 1155.

Schiffman, R. L. (1958) - "Consolidation of Soil under Time Dependent Loading and Varying Permeability" - Proc. HRB, Vol. 37, p. 584.

Schiffman, R. L. and Gibson, R. E. (1964) - "The Consolidation of Non-Homogeneous Clay Layers" - J. ASCE SMFD, Vol. 90, SM 5.

403.

Schmertmann, J. M. (1953) - "Estimating the Time-Consolidation Be- haviour-of Clay from Laboratory Test Results" - Proc. ASCE, Separate No. 311.

Schmidt, B. (1967) - "Lateral Stresses in Uniaxial Strain" - Danish Geo. Institute, Bulletin, No. 23.

Schmidt, B. (1966) - Discussion on Brooker and Ireland, 1965 CGJ, Vol. 3, p. 239.

Scott, R. F. (1963) - "Principles of Soil Mechanics" - Addison-Wesley, Boston.

Seed, H. B. (1965) - "Settlement Analysis - A Review of Theory and Testing Procedures" - J. ASCE SMFD, Vol. 91, SM 2.

Seed, H. B. and Chan (1959) - "Structure and Strength Character- istics of Compacted Clays" - Proc. ASCE SMFD, Vol. 85, SM 5.

Selby, S. M. and Girling, B. (1965) (Ed.) - "Standard Mathematical Tables" - The Chemical Rubber Co., Cleveland.

Siorota, S. and Jennings, R. A. J. (1959) - "The Elastic Heave of the Bottom of Excavation" - Geotechnique, Vol. 9, p. 62.

Sherlock, R. L. (1962) - "British Regional Geology - London and Thames Valley" - Geol. Survey, 3rd Ed., HMSO.

Sherman, D. I. (1959) - "On the Problem of Plane Strain in Non-Homogeneous Media" - Non-Homogeneity in Elasticity and Plasticity, Pergamon Press.

Sheppard,G. and A3rien,L. (1957) - "The Usk Scheme for the Water Supply of Swansea" - Proc. ICE, Vol. 7, p. 240.

Shields,D.and Rowe, P. W. (1965) - "Radial Drainage Oedometer for Laminated Clays" - J. ASCE SMFD, Vol. 91, SM 1.

Shibata, T. and Karbbe, D. (1965) - "Influence of the Variation of the Intermediate Principal Stress on the Mechanical Properties of Normally Consolidated Clays" - Proc. 6th ICSMFE, Montreal, Vol. 1, P. 359.

Simons, N. E. (1957) - "Settlement Studies of Two Structures in Norway" - Proc. 4th ICSMFE, London, Vol. 2, p. 13.

4o4.

Simons, N. E. (1958) - "Contribution to Discussion" - Brussels Conf. on Earth Pressure Problems, Vol. 3, P. 50.

Simons, N. E. (1960a) - "Comprehensive Investigation of the Shear Strength of Undisturbed Drammen Clay" - Proc. Res. Conf. on Shear Strength of Cohesive Soils, Boulder, p. 727.

Simons, N. E. (1960b) - "The Effect of Over-Consolidation on the Shear Strength Characteristics of Undisturbed Oslo Clay" - Ibid, p. 747.

Simons, N. E. (1963) - "Settlement Studies of an Apartment Building in Oslo" - Proc. European Conf. on Settlement Problems, Vol. 1, p. 179.

Simons, N. E. (1965) - "Consolidation Investigation of Undisturbed Fornebu Clay" - N.G.I. Publication No. 62.

Sinclaire, S. R. and Brooker, E. W. (1967) - "The Shear Strength of Edmonton Shale" - Proc. Geo. Conf. Oslo, Vol. 1, p. 295.

Skempton, A. W. (1948) - "The Rate of Softening in Stiff Fissured Clays with Special Reference to London Clay" - Proc. 2nd ICSMFE, Rotterdam, Vol. 2, p. 50.

Skempton, A. W. (1948a) - "The Effective Stresses in Saturated Clays Strained at Constant Volume" - Proc. 7th Int. Cong. on Appl. Mech., Vol. 1, p. 378.

Skempton, A. W. (1951) - "The Bearing Capacity of Clays" - Building Res. Congress.

Skempton, A. W. (1954) - "The Pore Pressure Coefficients A and B" - Geotechnique, Vol. 4, p. 143.

Skempton, A. W. (1959) - Letter to Geotechnique, Vol. 9, p. 145.

Skempton, A. W. (1960a) - Letter to Geotechnique, Vol. 10, No. 4.

Skempton, A.. W. (1960b) - "Terzaghi's Discovery of Effective Stress" - From Theory to Practice in Soil Mechanics, John Wiley & Sons.

Skempton, A. W. (1961) - "Horizontal Stresses in an Over-Consolidated Eocene Clay" - Proc. 5th ICSMFE, Paris, Vol. 1, p. 351.

Skempton, A. W. (1964) - "Long Term Stability of Clay Slopes" - 4th Rankine Lecture, Geotechnique, Vol. 14, No. 2.

405;

Skempton, A. W. and Bishop, A. W. (1954) - "Soils" - Chapter 10 of "Building Materials" - North Holland Publishing Co., Amsterdam.

Skempton, A. W. and Bjerrum, L. (1957) - "A Contribution to the Settle- ment Analysis of Foundations on Clay" - Geotechnique, Vol. 7, No. 4.

Skempton, A. W. and Henkel, D. J. (1957) - "Tests on London Clay from Deep Borings at Paddington, Victoria and South Bank" - Proc. 4th ICSMFE, London, Vol. 1, p. 100.

Skempton, A. W. and La Rochelle, P. (1965) - "The Bradwell Slip - A Short Term Failure in London Clay" - Geotechnique, Vol. 15, No.3.

Skempton, A. W. and McDonald, D. H. (1955) - "A Survey of Comparisons Between Calculated and Observed Settlements of Structures in Clay" - Proc. Conf. Corr. Calc. and Obs. Stresses and Displ. of Structures, ICE, Vol. 1, p. 318.

Skempton, A. W., Peck, R. B. and McDonald, D. H. (1955) - "Settlement Analysis of Six Structures in London and Chicago" - Proc. ICE, Vol. 4, Pt. 1, p. 525.

Skempton, A.14 and Sowa, V.(1963) during Sampling and Testing"

Skinner, A. E. (1967) - "Cobbins Dissipation Tests on London Partners, Nov. 1967.

- "Behaviour of Saturated Clays - Geotechnique, Vol. 13, No. 4.

Brook Reservoir - Pore Pressure Clay" Unpub. Report to Binnie and

Sowa, VS1963) - "A Comparison of Isotropic and Ani'sotropic Consolida-tion on the Shear Behaviour of a Clay" - Ph.D. Thesis, University of London.

Smith, C. K. and Redlinger, J. F. (1953) - "Soil Properties of Fort Union Clay Shale" - Proc. 3rd ICSMFE, Zurich, Vol. 1, p. 62.

Steffanof, G., 7latarev, K., Grantcharov, M. and Miley, G. (1965) -"Anticipated and Observed Settlement of a High Chimney" - Proc. 6th ICSMFE, Montreal, Vol. 2, p. 202.

Steinbrenner, W. (19 4) - "Tafeln zur Setzungbenechnung" Strasse, Vol. 1, p. 121.

Tan, S. B. (1968) - "Consolidation of Soft Clays with Special Refer-ence to Sand Drains" - Ph.D. Thesis, University of London.

406.

Taylor, D. W. (1942) - "Research on the Consolidation of Clays" -M.I.T. Report 82, Civil and Sanitary Eng. Department.

Taylor, D. W. (1948) - "Fundamentals of Soil Mechanics" - John Wiley and Sons.

Taylor, D. W. and Clough, R. H. (1951) - "Report on Research on Shearing Characteristics of Clay" - M.I.T. Report.

Terzaghi, K. (1925) - "Erdbaumechanik" - Deutiche, Vienna.

Terzaghi, K. (1929) - "Settlement of Buildings due to Progressive Consolidation of Individual Strata" - J. of Math. Physics, Vol. 8, p.26. See also Terzaghi, K., "Settlement Analysis" - World Eng. Congress (Tokyo, 1929), Vol. 8, Pt. 2, p. 1.

Terzaghi, K. (1933) - "Earths and Foundations'►-Progress Report of Special Committee - Proc. ASCE, Vol. 59, p. 1360, Discussion by Terzaghi, K., p. 1363.

Terzaghi, K. (1936) - "Stability of Slopes of Natural Clays" Proc. 1st ICSMFE, Vol. 1, p. 161.

Terzaghi, K. (1936a) - "Opening Discussion on 'Settlement of Structures'," - Proc. 1st ICSMFE, Cambridge, Mass., Vol. 3, P. 79.

Terzaghi, K. (1943) - "Theoretical Soil Mechanics" - John Wiley and Sons Inc.

Terzaghi, K. (1941) - "Undisturbed Clay Samples and Undisturbed Clays" - J. Boston Soc. of Civil Eng., Vol. 28, .No. 3.

Terzaghi, K. (1955) - "Evaluation of Coefficients of Subgrade Reaction" - Geotechnique, Vol. 5, p. 297.

Terzaghi, K. (1961) - "Discussion on Skempton 1961" - Proc. 5th ICSMFE, Paris, Vol. 3, p. 144.

Terzaghi, K. and Fv8lich (1936) - "Theorie der Setzung von Ton-schichten" Denticke, Vienna.

Terzaghi, I. and Peck, R. B. (1948) - "Soil Mechanics in Engineering Practice" - John Wiley and Sons.

Texeira, A. H. (1960) - "Case History of Building Underlain by Un-usual Conditions of Pre-Consolidated Clay Layers" - Proc. 1st Pan Am Conf. on S.M., Vol. 1, p. 201.

407.

Timoshenko, S. and Goodier, J.(1954) - "Theory of Elasticity" -McGraw Hill PubIiShing Co.

Timoshenko, S. (1955) - "Strength of Materials" - D. Van Norstand Co. Inc.

Toms, A. H. (1966) - "Errors in Soil Consolidation Presses" - Eng-ineering, 22nd July 1966.

Triaxial Shear Report (1947) - Waterways ExperimentSAation, Vicksburg, Miss., April 1947.

Tscheborarioff, G. P. (1952) - "Soil Mechanics, Foundations and Earth Structures" - McGraw Hill Publishing Co.

Turnbull, W. J., Maxwell, A. and Ahlvin, R. G. (1961) - "Stresses and Deflections in Homogeneous Soil Mass" - Proc. 5th ICSMFE, Paris, Vol. 2, p. 337.

Uyeshita, K. and Meyerhof, G. G. (1967) - "Deflections of Multi-Layered Soil System" - J. ASCE SMFD, Vol. 93, SM 5.

Vesic, A.B. (1963) - "The Validity of Layered Solid Theories for Flexible Pavements" - Proc. Int. Conf. on Str. Design of Asphalt Pavements, Univ. of Michigan, p. 283.

Vesic, A. B. and Barksdale, R. D. (1963) - "On Shear Strength of Sand at Very Hi Pressures" - ASTM STP No. 361, p. 301.

Vesic, A. B. and Clough, G. W. (1968) - "Behaviour of Granular Materials under High Pressures" - J. ASCE SMFD, Vol. 94, OM 3.

Wade, N. H. (1953) - "Plane Strain Failure Characteristics of Saturated Clay" - Ph.D. Thesis, University of London.

Ward, W. H., Saatels, S. G. and Butler, M. E. (1959) - "Further Studies of the Properties of London Clay" - Geotechnique, Vol. 9, No. 2.

Ward, W. H., Marsland, A. and Samuels, S. G. (1965) - "Studies of the Properties of Undisturbed London Clay at the Ashford Common Shaft" - In-situ and Immediate Strength Tests" - Geotechnique, Vol. 15, No. 3.

Waterways Expt. Station, Vicksburg (1953, 1954) - "Investigations of Pressures and Deflections for Flexible Pavements" - Report 3 (Sept. 1953), Report 4 (Dec. 1954).

408.

Webb, D. L. (1966) - "The Mechanical Properties of Undisturbed Samples of London Clay and Pierre Shale" - Ph.D. Thesis, Univ-ersity of London.

Whitman, R. V. and Miller, E. T. (1965) - "Yielding and Locking of Confined Sand" - J. ASCE SMFD, Vol. 91, SM 4.

Whitman, R. V., Richardson, A. M., and Healey, K. A. (1961) . "Time Lags in Pore Pressure Measurements" - Proc. 5th ICSMFE, Faris, Vol. 1, p. 407.

Williams, G. M. J. and Rutter, P. A. (1967) - "The Design of Two Buildings with Suspended Structures in High Yield Steel" - The Structural Engineer, Vol. 45, No. 4.

Wilson, G. H. and Grace, H. (1942) - "The Settlement of London due to Undordrainage of the London Clay" - J. ICE, Vol. 19, No. 2, p. 100.

Wilson, S. D. and Hancock, C.(1960) - "Horizontal Displacements in Clay Foundations" - Proc. 1st Pan Am, Conf. on SMFE., Vol. 1, p. 41.

Wood, C. C. (1958) - "Shear Strength and Volume Change Characteristics of Compacted Soil Under Conditions of Plane Strain" - Ph.D. Thesis, University of London.

Woodward, H. B. (1922) - "The Geology of the London Basin" - Memoirs of the Geol. Survey, 2nd Ed., HMSO.

Wooldridge,S.(1926) - "Structural Evolution of London Basin" - Proc. Geol. Ass., Vol. 37, p. 162.

Wooldridge, S. W. and Linton, D. L. (1955) - "Structure, Surface and Drainage in S.E. England" - George Philip and Son Ltd.

Winkler, E. (1867) - "Die Lehre von Elastizitat and Fostigteit" Prague, p. 182.

Wu, T. H. (1966) - "Soil Mechanics" - Allyn and Bacon Inc.

Wu, T. H., Loh, A. K. and Malvern, L. E. (1963) - "Study of the Failure Envleopocf Soils" - J. ASCE SMFD, Vol. 89, SM 1.

Yong, R. N. and Osier, J. C. (1965) - "Contribution to Discussion" - CGJ., Vol. 2, p. 229.

409.

Yu, T. M., Shu, W. Y. and Tong, Y. K. (1965) - "Settlement Analysis of Pile Foundations of Structures in Shanghai" - Proc. 5th ICSMFE., Montreal, Vol. 2, p. 356.

Zeevaert, L. (1953) - "Contribution to Discussions" - Proc. 3rd ICSMFE, Zurich, Vol. 3, p. 113.

Zeevaert, L. (1958) - "Consolidation of Mexico City Volcanic Clay" - Conf. on Soils for Eng. Purposes, ASTM STP 232.