Sunteți pe pagina 1din 226

09031-02[1].

qxd 18/Oct/02 12:09 AM Page 29

Characterisation and Engineering Properties of Natural Soils – Tan et al. (eds.)


© 2003 Swets & Zeitlinger, Lisse, ISBN 90 5809 537 1

Behaviour and properties of natural soils and soft rocks


S. Leroueil
Department of Civil Engineering, Université Laval, Ste-Foy, Québec, Canada, G1K 7P4

D.W. Hight
Geotechnical Consulting Group, 1a Queensberry Place, London, United Kingdom, SW7 2DL

ABSTRACT: This paper is an overview of the mechanical and hydraulic behaviour and properties of
natural soils and soft rocks, from soft clays to chalk. It is based on Critical State Soil Mechanics, but the
associated concepts are extended to take into account the influence of the major features of natural soils
that are possible: anisotropy, microstructure, viscosity, partial saturation and discontinuities. The first
part of the paper deals with soil behaviour (14 sections); the second part is more oriented towards soil
properties (6 sections).

1 INTRODUCTION

Classical soil mechanics developed mainly in the first half of the twentieth century. In particular, the fol-
lowing developments can be mentioned: consolidation theory (Terzaghi, 1923), concept of effective stress
(Terzaghi, 1936), concept of preconsolidation pressure (Casagrande, 1936b), understanding of strength
components (Hvorslev, 1937), description of soil compaction and consistency (Proctor, 1933; Atterberg,
1911), and understanding of drained vs undrained soil behaviour (Bishop; Skempton). These works still
cover about 70% of our soil mechanics textbooks. What can be called Modern Soil Mechanics was born at
the turn of the nineteen-sixties, in particular with Roscoe, Schofield, Wroth and co-workers who integrated
shear stress, mean effective stress, void ratio and shear strain in the same framework, often referred to as
Critical State Soil Mechanics (CSSM). This work was reinforced with the stress path method (Lambe, 1967)
and the concept of normalized soil behaviour (Ladd, 1969). It is also at that time that the influence of dila-
tancy was fully understood (Rowe, 1962; Lee & Seed, 1967). It was however in the nineteen-eighties that
some major features of natural soils have been fully recognised: small strain stiffness (Jardine, 1985;
Burland, 1989) and its measurement by geophysical methods (Stokoe & Nazarian, 1985; Stokoe et al.,
1991); concept of collapse surface for explaining static liquefaction of sand (Sladen et al., 1985); importance
of effect of strain rate (Graham et al., 1983a); influence of anisotropy and stress axis rotation (Hight and
co-workers); extension of CSSM concepts to unsaturated soils (Alonso et al., 1987, 1990); the fact that most
natural geomaterials are microstructured (Burland, 1990; Leroueil & Vaughan, 1990). In the last fifteen
years or so: (a) these ideas have essentially been verified for a variety of geomaterials; (b) equipment, in par-
ticular for small strain and shear wave measurement, has been improved; and (c) more and more powerful
numerical models integrating recent findings have really started being applied to analyses in major projects.
The aim of this paper is to synthesize this Modern Soil Mechanics into a general framework that
could be applied to a variety of geomaterials, from soft clays to soft rocks, considered homogeneous.
The basic concepts of CSSM are used as the foundation of this framework. Then, major features of nat-
ural soil behaviour, such as anisotropy, microstructure, viscosity and partial saturation are included in
the framework as extensions of the basic CSSM. The preparation of the paper has been greatly facili-
tated by the reading of contributions to this Workshop.
The paper contains 19 sections in addition to the introduction and a brief conclusion. The sections are
divided in two parts: Part I deals mainly with soil behaviour; and Part II is more oriented towards soil
properties. The general outline is as follows:
Part I: Soil behaviour
– Basic elasto-plastic model and failure criteria for soils (Section 2)
– Limit state curves of natural soils and soft rocks (Section 3)

29
09031-02[1].qxd 18/Oct/02 12:09 AM Page 30

– Behaviour inside the limit state curve (Section 4)


– Behaviour of cohesionless soils (Section 5)
– Influence of anisotropy (Section 6)
– Behaviour of intermediate soils (Section 7)
– Post-failure behaviour (Section 8)
– Influence of strain rate and time (Section 9)
– Influence of temperature (Section 10)
– Influence of microstructure (Section 11)
– Influence of discontinuities (Section 12)
– Influence of soil water chemistry (Section 13)
– Influence of partial saturation (Section 14)
Part II: Soil properties
– Generalities on soil properties (Section 15)
– Properties of soils in remoulded conditions (Section 16)
– Mechanical properties of intact soil (Section 17)
– Properties associated with the passage from intact to destructured, possibly remoulded, conditions
(Section 18)
– Hydraulic properties (Section 19)
– Characteristics of consolidation (Section 20)
Some parts of the paper will appear trivial to many readers. This has (partly) been done on purpose,
the paper aiming at being both close to a state-of-the-art paper and a tool that could be used for pedago-
logical purposes. It is, indeed, an extended and updated version of the “Advanced Soil Mechanics”
course available at Université Laval, Québec.

2 BASIC ELASTO-PLASTIC MODEL (CRITICAL STATE SOIL MECHANICS) AND


FAILURE CRITERIA FOR SOILS

2.1 Elasto-plastic behaviour of ideal soils (Critical state soil mechanics)


Basic concepts of Critical State Soil Mechanics (CSSM) were established in the late nineteen-fifties
and -sixties in England (Roscoe et al., 1958; Schofield & Wroth, 1968), mostly on the basis of tests per-
formed on reconstituted and isotropically consolidated samples of clay. It is representative of saturated
soils which are assumed to be isotropic and not influenced by factors such as strain rate and microstruc-
ture. CSSM thus describes the behaviour of an “ideal soil” and can be summarized as below.
When a clayey soil, initially at a water content close to its liquid limit, is loaded isotropically, its void
ratio decreases along the normal compression line (NCL), a straight line in an e–log p (or v–ln p) dia-
gram1 (Figure 2.1a). If after being loaded to pA, at point A, the soil is progressively unloaded isotropi-
cally, there is a slight rebound along AR, also a straight line in the e–log p diagram. If reloaded from
R, the soil first behaves elastically from R to A, and then follows the normal compression line along AB,
where deformations are mostly plastic. For the soil at R, A is a yield point or a limit state, at which it
starts developing plastic strains. The NCL and the recompression line are characterized by their slopes,
 e/ln p (or the compression index Cc  e/log p) and
 e/ln p.
In a p–q diagram, a soil loaded at point A, under an isotropic effective stress pA, generates a stress
domain in which soil behaviour is elastic. This domain is limited by the yield or limit state curve pass-
ing through point A (Figure 2.1b). For a stress path such as LM, inside the limit state curve, the behav-
iour is elastic; whereas for a stress path such as LT, it is elastic from L to S, and plastic strain will
develop from S to T. For the soil loaded to point B (Figure 2.1a), the limit state curve becomes the
dashed line passing through B in Figure 2.1b. The new limit state curve is associated with the accumu-
lated volumetric plastic strain, and its shape is the same (isotropic hardening), regardless of the stress
path along which it has been created. In an (e–p–q) diagram (Figure 2.2b) the normal compression line
and the limit state curves generate the limit state (or state boundary) surface. When stress conditions on
the limit state surface are normalized with respect to the mean effective stress on the NCL
at the same void ratio, pe, then they lie on the same line in a normalized p/pe–q/pe diagram
(Figure 2.3).

1
Definition of symbols is given in Section 22, at the end of the text.

30
09031-02[1].qxd 18/Oct/02 12:09 AM Page 31

e Log p´

L
CS
Normal
compression
line (NCL) Yield or limit
state curve

R
A T
M S

CSL A B P´
(a) B (b)

Figure 2.1. Limit state or yield or state boundary surface.

q p´
M
q=
C1
T1

T2 C2
T3 C3
O N N N P´
3 2 1
(a)

q
CSL
C1

Hvorslev T1
surface


N1
T2 C2
NCL
N2
T3 C3
e N3
(b)
Figure 2.2. Limit state or yield or state boundary surface.

When yielding occurs at a stress state such as S in Figures 2.1b and 2.4, it is associated with the
development of plastic strains. The plastic strain increments can be described by vp and sp which can
be plotted at point S to form a plastic strain increment vector SV. v (also termed vol) is the volumetric
strain (V/Vo  1  2  3) and s is the shear strain (2( 1  1)/3) in axisymmetic conditions),

31
09031-02[1].qxd 18/Oct/02 12:09 AM Page 32

q/pe´
Mc
C

Compression
Hvorslev Limit state
surface surface
Critical state

A
O
0.5 1.0
p´/p´e
Extension

Me

Figure 2.3. Limit state surface and Hvorslev surface in both compression and extension, in the normalized
p/pe–q/pe diagram.
q

L
CS
p
s
δε

T
V
δε p
S s
δε p
v

L
p
δε v A P´

Figure 2.4. Limit state curve and plastic strain increment vector when normality applies.

with 1, 2 and 3 being the major, the intermediate and minor principal strains. It is generally assumed
that vector SV is perpendicular to the yield or limit state curve; it is then said that there is normality or
associated flow.
At large strains in shear tests, the soil tends towards an ultimate condition at which there is no change
in void ratio and effective stresses. This condition is called critical state or fully softened state. Critical
state can be obtained by drained and undrained shear tests performed on normally or slightly overcon-
solidated clay samples; it is then reached at large strain at the same time as the maximum deviatoric,
( 1 – 3), or shear stress. For a heavily overconsolidated soil, critical state is also reached at large strain,
but after a peak and strain softening in the case of drained shear tests.
Critical state conditions are on a curve, (CSL), in the e–p–q diagram (Figure 2.2). Projected onto the
q–p space, the CSL becomes straight and is defined as follows:

qcs  Mpcs (2.1)

32
09031-02[1].qxd 18/Oct/02 12:09 AM Page 33

where: M  Mc  6 sin /(3  sin ) for triaxial compression conditions and M  Me  6 sin /
(3  sin ) for triaxial extension conditions.
Equation 2.1 is also the Mohr-Coulomb criterion defined for normally consolidated materials and  
is the corresponding friction angle, nc, or critical state friction angle cs. Typical values of friction
angles and relationships with other soil parameters are presented in Section 16.5.
In the e–log p diagram, the projection of CSL is parallel to the normal compression line
(Figure 2.1a). Consequently, in a p/pe–q/pe diagram (Figure 2.3), CSL is represented by a unique
point, C.
On the basis of tests performed with the direct shear box on normally and overconsolidated speci-
mens of clay, Hvorslev (1937) came to the conclusion that the peak shear strength f under an effective
stress  can be described as follows:

f  ce   tan e (2.2)

where e is the Hvorslev angle of shearing resistance and ce is the Hvorslev cohesion, which increases
when the void ratio decreases.
As shown by Schofield & Wroth (1968) and by Wood (1990),

ce  cpe pe (2.3)

cpe is the Hvorslev cohesion parameter. Thus the strength envelopes can be described as follows in the
q/pe–p/pe diagram:
– In triaxial compression

q 6 cos e′ p′
 c′pe  Mce (2.4)
pe′ 3  sin e′ pe′

with

6 sin e′
Mce  (2.5)
3  sin e′

– In triaxial extension

q 6 cos e′ p′
 c′pe  Mee (2.6)
pe′ 3  sin e′ pe′

with

6 sin e′
Mee  (2.7)
3  sin e′

These normalized strength envelopes are shown in Figure 2.3. In the e–p–q diagram, they generate
the upper (and lower) part of the limit state surface, also called the Hvorslev surface (Figure 2.2b).
The previously mentioned elements constitute the basics of CSSM that apply well to “ideal” soils.
They are essential to the understanding of soil behaviour and constitute the foundation on which it is
possible to build a modern soil mechanics to take into account the features of “real” soils. These
features will be examined in the following sections.
A simple application of CSSM to triaxial compression tests is illustrated in Figure 2.5 for soil spec-
imens normally consolidated at point M and then isotropically unloaded at different overconsolidation
ratios, OCRs, before shearing. OCR is defined as the ratio of the maximum axial or mean stress applied
in the past over the currently applied stress. The soil, which initially is in the overconsolidated range
behaves elastically until it reaches the limit state curve (at points A1, B1, C1 and D1) and then progres-
sively moves towards the critical state corresponding either to the effective stress path followed in
drained tests (AA1A2 and BB1B2) or its void ratio in undrained tests (AC1C2 and BD1D2). It can be

33
09031-02[1].qxd 18/Oct/02 12:09 AM Page 34

q
A2
L A2
CS

C2 B1
C1 C2
B1 C1
D1 A1 D2
D2 D1
A1
B2 B2

A
B A M BC ε
P' 1
(a) (b)
e or w

e or w
P'

B2
B2
NC
L D1
D1
D2 B D2
B
B1 A A1 C 1 B1
C2 M A C2
C1 A1
CS
L
A2
A2
ε1
(c) (d)

Figure 2.5. Drained and undrained compression triaxial tests (schematic behaviour in CSSM).

noted, in particular, that the drained test on the slightly overconsolidated soil specimen (stress path
AA1A2) shows ductile behaviour whereas the drained test on the strongly overconsolidated soil speci-
men (stress path BB1B2) shows a strain softening behaviour.

2.2 Failure criteria


It is generally accepted that failure of soils is controlled by the Mohr-Coulomb criterion
f  c   tan  (2.8)

In an octahedral (p  cst) plane, this criterion is an irregular hexagon, as shown in Figure 2.6.
Experience however shows that the failure surface is more rounded than indicated by the Mohr-
Coulomb criterion.
Matsuoka & Nakai (1974) and Lade & Duncan (1975) proposed slightly different criteria that are
described as follows:
1) Matsuoka & Nakai criterion:
I1 ⋅ I2
 cst (2.9)
I3

34
09031-02[1].qxd 18/Oct/02 12:09 AM Page 35

Figure 2.6. Failure criteria in octahedral (p  cst) plane.

2) Lade & Duncan criterion:

I13
 cst (2.10)
I3

where I1, I2 and I3 are the first, second and third stress invariants:

I1  1  2  3 (2.11a)
I2  1 2  2 3  3 1 (2.11b)
I3  1 2 3 (2.11c)

Aubertin & Simon (1996, 2000) also proposed a similar strength criterion for rocks.
The criteria proposed by Matsuoka & Nakai (1974) and by Lade & Duncan (1975) are also shown in
Figure 2.6. It can be seen that the Matsuoka & Nakai’s criterion gives the same axisymmetric extension
(b  ( 2 – 3)/( 1 – 3)  1 in Figure 2.6) strength as Mohr-Coulomb whereas the Lade and Duncan’s
criterion indicates a strength in axisymmetric extension that is slightly larger. Figure 2.7 shows results
obtained on loose and dense Monterey No 0 sand. In both cases, the agreement with Lade & Duncan’s
model is very good.
From the compilation of numerous experimental results, Kulhawy & Mayne (1990) obtained the
relative values of effective friction angles (as defined assuming the Mohr-Coulomb criterion) given in
Table 2.1. It can be seen that the friction angle deduced from triaxial extension tests, te, is larger than
friction angle deduced from triaxial compression tests, tc, typically by 10 to 25%. Lade (2002) indicates
that Equation 2.10 gives a te/tc ratio varying between 1.10 at low friction angles to 1.18 at high fric-
tion angles. It thus seems that 1.15 can be considered as being a reasonable practical number for te/tc.

3 LIMIT STATE CURVES OF NATURAL SOILS AND SOFT ROCKS


(INFLUENCE OF ANISOTROPY)

3.1 Influence of anisotropy


The effective stress paths followed in undrained shear tests after consolidation in the normally consoli-
dated range lie on the limit state surface. Thus, these stress paths are very close to the specific limit state

35
09031-02[1].qxd 18/Oct/02 12:09 AM Page 36

Figure 2.7. Failure surfaces for loose and dense Monterey No 0 sand in octahedral plane (after Lade & Duncan,
1975).

Table 2.1. Relative values of effective stress friction angles (after Kulhawy &
Mayne, 1990).

Friction angle

Test type Cohesionless soil Normally consol. clay

Triaxial compression 1.00 t c 1.00 t c


Triaxial extension 1.12 t c 1.22 t c
Plane strain compression 1.12 t c 1.10 t c
Plane strain extension 1.25 t c 1.34 t c

curves generated by the stress conditions applied during the consolidation stage. An idea of the influ-
ence of consolidation stress conditions on the shape of the limit state surface can thus be obtained from
such tests. Ladd & Varallyay (1965) consolidated specimens of Boston Blue clay in the normally con-
solidated range, under ratios of the radial effective stress r to the axial effective stress a equal to
Konc  0.54, 1.00 and 1/Konc  1.85. The stress paths followed during the subsequent undrained com-
pression and extension tests are shown in Figure 3.1. The following can be seen: (a) the shape of the
limit state surface is strongly influenced by the effective stress ratio applied in the normally consoli-
dated range; (b) at large strains, the stress paths go to the same strength envelope; (c) for the same con-
solidation conditions, the stress conditions at large strain are approximately at the same value of mean
effective stress p in compression and in extension, indicating that the critical state concept is not
affected by initial anisotropy; (d) as the M parameter is not the same in compression and in extension,
the deviatoric stress obtained at large strain in extension is smaller than that obtained in compres-
sion. According to Equations 2.5 and 2.7 and the values of Mc and Me, the ratio should be equal to
(3  sin )/(3  sin ).

3.2 Limit state curves of natural clays


For natural soils that may be microstructured, i.e. with bonds at the interparticle contacts, determination
of the limit state curve has to be obtained from tests with initial conditions inside the limit state curve

36
09031-02[1].qxd 18/Oct/02 12:09 AM Page 37

300
σ 'a
0.54
=
σ' r
200

ion
ss
re
mp
100
Co
q (kPa)

σ'r = σ 'a
0
Ex
te
ns

-100 σ'
io

r =1
n

.85
σ'
a

-200

-300
0 100 200 300 400
p' (kPa)

Figure 3.1. Stress paths followed during compression and extension undrained triaxial tests after consolidation in
the normally consolidated range under effective stress ratio r / a of 0.54, 1.0 and 1.85 (after Ladd & Varallyay, 1965).

(essentially in the elastic domain) and stress paths going towards and possibly outside the limit state
curve, where large plastic strains start developing. One possibility is to consolidate the soil close to in
situ effective stresses and follow radial stress paths; another possibility is to combine isotropically or
anisotropically consolidated drained or undrained triaxial (CID, CAD, CIU or CAU) tests and
anisotropic compression (K  r / a  cst) triaxial tests.
The clay chosen to show the determination of the limit state curve of natural clay comes from Saint-
Alban (Tavenas & Leroueil, 1977). Undrained triaxial compression (CIU) and anisotropic consolidation
tests were performed. Figure 3.2a shows the variation in deviatoric stress and pore pressure as a func-
tion of strain for three CIU tests performed in the overconsolidated range. At the start of the test, the
clay shows a strong resistance to compression until a peak is reached at a strain below 1%; beyond the
peak, large strains develop, associated with strain softening. The stress state corresponding to the pas-
sage through the peak can be considered as a gross yield point (see Section 4) for the clay. The stress
paths associated with these three tests are shown in Figure 3.2c. The volumetric strains seen during
K  cst tests are shown as a function of ( a  r )/2 in Figure 3.2b. The passage from the overconsol-
idated state to the normally consolidated state, that is the gross yield point, is well defined in all cases.
The stress states corresponding to gross yielding are assembled in Figure 3.2c, so that the limit state
curve of the Saint-Alban clay from a depth of 3 m can be completely defined.
A number of remarks can be made in relation to the limit state curve and the behaviour of natural
clays. They are summarized in the following paragraphs.
The limit state curve is approximately centred on the Konc line in the ( a – r)/2 vs ( a  r)/2 dia-
gram. This is certainly attributable to the anisotropy of stresses which exists during the formation of the
deposit (see Section 16.7, Figure 16.6 and Equations 16.9 to 16.11).

37
09031-02[1].qxd 18/Oct/02 12:09 AM Page 38

Figure 3.2. Triaxial testing on Saint-Alban (3 m) clay: (a) CIU tests; (b) anisotropic consolidation; (c) limit state
curve (after Tavenas & Leroueil, 1977).

The oedometer test also gives a gross yield point characterized by a vertical effective stress equal to
the preconsolidation pressure p 2. Unfortunately, in these tests, the lateral stress is not usually mea-
sured and the corresponding stress state is not known. The only thing that can be stated is that the stress
state lies on the line a  p . However, the definition of the preconsolidation pressure and experience
show that the state of stress at p is generally between points P and A in Figure 3.2c.

2
In its geological meaning, the preconsolidation pressure is unique and constant. On the other hand, the “gross
yield stress” separating small to moderate strains from large strains depends on strain rate, temperature and
microstructure. It must thus be considered as a rheologic parameter. It could have a particular appellation such
as “vertical yield stress” or “apparent preconsolidation pressure”, but engineers often refer to it as “precon-
solidation pressure”. So, this term is also used in this paper.

38
09031-02[1].qxd 18/Oct/02 12:09 AM Page 39

250
σ' (kPa) ✛

p
200 191
241
310
q (kPa) 150
380

100



50 ✛

0
0 50 100 150 200 250 300 350
p' (kPa)
Figure 3.3. Limit state curves for Winnipeg clay (from Graham et al., 1983b).

In general, whatever the degree of microstructuration, the stress paths followed during undrained tri-
axial tests (CIU or CAU) are at about constant p (Skempton’s parameter A  u/( 1  3)  0.33)
when the OCR is about 2 to 3. The stress path tends to bend towards lower effective stresses (parameter
A increases) when the overconsolidation ratio is smaller; it tends to bend towards larger effective
stresses (parameter A decreases) when the overconsolidation ratio is larger. These tendencies become
more marked as the stress level increases and failure approaches. This latter aspect of soil behaviour,
which has also been observed on reconstituted clays (Wroth & Loudon, 1967) was pointed out by
Cotecchia (2002*)3 and Leroueil et al. (2002*).
The upper part of the limit state curve is also the strength envelope of the overconsolidated soil. It
corresponds to the Hvorslev surface described in Section 2 (Figures 2.2 and 2.3) for non-microstructured
soils; it can be higher than the Hvorslev surface in microstructured soils (see Section 11). The undrained
shear strength of the overconsolidated soil is also on the peak strength envelope, at a location that
depends on the overconsolidation ratio. Ladd et al. (1977) proposed to express the undrained shear
strength of the soil overconsolidated, su, as follows:

su  s 
  u  ⋅ OCR m
ac
′  ac
′  nc (3.1)

Where ac is the axial consolidation stress, (su/ ac)nc is the strength ratio of the normally consolidated
soil and m a parameter that depends on the soil and the type of test. Parameter m is often close to 1.0,
which implies that su/ p is constant in the overconsolidated range. This characteristic is often used in
practice.
A general observation made is that the compression index obtained in anisotropic consolidation tests
depends on the effective stress ratio K. As indicated in Figure 3.2b, it is much larger for K values close
to Konc (equal to about 0.5) than for isotropic loading.
The limit state curves obtained at different depths at the same site have the same shape (Figure 3.3)
and can be normalized using the preconsolidation pressure. In fact, all the behaviour inside the limit
state curve can be normalized using the preconsolidation pressure, at least for clays having the same
geological history and similar microstructure. Thus, samples taken from different depths and consoli-
dated under stresses c such that c / p is the same, show the same relations ( a – r)/ p  f( ) and
u/ p  g( ) when they are sheared (Figure 3.4). This type of normalization was included in the NSP
concept of Ladd (1969) and Ladd & Foott (1974), and is often used in practice (Ladd et al., 1977).

3
2002* refers to a contribution to this workshop.

39
09031-02[1].qxd 18/Oct/02 12:09 AM Page 40

Compiling limit state curves obtained on natural clays from 10 different countries, Diaz-Rodriguez
et al. (1992) found the results summarized in Figure 3.5. Figure 3.5a shows extreme cases obtained on
Winnipeg clay (nc  17.5°) and Mexico City clay (nc  43°). As shown in Figure 3.5b, the geomet-
rical characteristics of the limit state curves ( a – r)max/2 p and piso/ p (where piso is the gross

Figure 3.4. Normalized stress-strain and pore pressure-strain relations in triaxial tests on intact Champlain sea clay
(from Leroueil et al., 1985c).

Figure 3.5. Summary on the limit state curves of natural soft clays (after Diaz-Rodriguez et al., 1992).

40
09031-02[1].qxd 18/Oct/02 12:09 AM Page 41

yield stress on the isotropic axis) respectively increases and decreases when nc increases. This indi-
cates that the shape of limit state curve of natural clays is essentially controlled by the friction angle of
the normally consolidated soil or critical state friction angle. Characteristics of limit state curves of spe-
cial interest for this Workshop have been added to Figure 3.5. It can be seen that they are in general
agreement with the data previously compiled. As indicated in Section 3.1 (Figure 3.1) and in Figure 3.6,
the shape of the limit state curve is reflected by the stress paths followed in undrained compression and
extension triaxial tests. In this latter figure from Ladd et al. (1977), the influence of plasticity and, indi-
rectly, of the normally consolidated friction angle on the shape of the limit state curve was clearly indicated.
The failure criterion (Figures 2.6, 2.7, and 3.12) implies a strength in extension lower that the
strength in compression. This fact becomes even more important with anisotropy. As a result, the ratio
of the undrained shear strengths obtained in extension and in compression increases from about 0.3 for
low plasticity soils to almost 0.8 for high plasticity clays as shown in Figure 3.7. As indicated in
Sections 16.5 and 16.6 (Figures 16.3 and 16.4), plasticity index is not a good parameter for describing

Figure 3.6. Effective stress paths followed in undrained compression and extension tests on normally consolidated
clays (after Ladd et al., 1977).

Figure 3.7. Undrained strength ratios in extension and compression versus plasticity index (modified after
Kulhawy & Mayne, 1990).

41
09031-02[1].qxd 18/Oct/02 12:09 AM Page 42

Figure 3.8. Limit state curves of Bothkennar clay, intact and after isotropic loading to high stresses (after McCinty
et al., 2000).

mechanical behaviour of clayey soils, and it can be expected that the scatter observed in Figure 3.7
would be smaller if the strength ratios were to be plotted as a function of nc.
The fact that the shape of limit state curves depends on effective stress history implies that it evolves
depending on the effective stress path followed. This is called kinematic hardening. For a natural clay
loaded under one-dimensional compression (horizontal strain equal to zero) in the normally consoli-
dated range, the shape of the limit state curve would remain essentially the same; on the other hand,
under an isotropic loading in the normally consolidated range, the shape of the limit state curve would
progressively evolve from its initial shape (Figure 3.5) to a shape corresponding to an isotropic material
(Figure 2.3).
McCinty et al. (2000) performed triaxial tests on the intact Bothkennar clay taken from a depth of
10.95–11.73 m and obtained the limit state curve shown with solid circles in Figure 3.8. Eight other
specimens of this clay were isotropically consolidated to an effective stress of 210 kPa, 2.6 times the
isotropic gross yield stress of the intact soil. They were then unloaded, and reloaded under different
stress ratios in order to define the new limit state curve. The results are also shown in Figure 3.8 with
triangles. They show a limit state curve that rotated towards the isotropic axis. Other samples consoli-
dated in the normally consolidated range under different stress ratios and then unloaded and reloaded
show yielding conditions that generally confirm the rotation of the limit state curve with consolidation
stress ratio and thus kinematic hardening. Futai et al. (2001) and Almeida & Marques (2002*) observed
a similar rotation of the limit state curve of Sarapui clay when isotropically consolidated.

3.3 Limit state curves of other soils and soft rocks


Limit state curves have been determined on a variety of residual soils: volcanic agglomerate by Uriel &
Serrano (1973); residual soil from gneiss by Sandroni (1981); residual soil from basalt by Maccarini

42
09031-02[1].qxd 18/Oct/02 12:09 AM Page 43

150 te
l sta
a
itic line

( σ'a - σ 'r )/2 (kPa)


Cr
100

50

0
0 50 100 150 200 250 300
(σa' + σ'r )/2 (kPa)

Figure 3.9. Limit state curve obtained for a residual soil from gneiss (from Sandroni, 1981).

Figure 3.10. Failure and yield of a porous silty mudstone (after Ohtsuki et al., 1981).

(1987); etc. An example is shown in Figure 3.9. Machado & Vilar (2002*) also show a limit state curve
for a weathered colluvium. In all these cases, the limit state curves appear to be fairly isotropic.
Limit state curves were also determined on soft rocks such as chalk and calcareous rocks (Addis &
Jones, 1990; Elliot & Brown, 1985; Lagioia & Nova, 1995), mudstones (Ohtsuki et al., 1981), volcanic
tuff (Pellegrino, 1970; Aversa et al., 1991) and marls (Kavvadas et al., 1993, 2002*).
An interesting case was provided by Ohtsuki et al. (1981). These authors studied a weak volcanic
mudstone with a high void ratio of 1.40 and a high apparent preconsolidation pressure of 5 MPa. The
limit state curve, shown in Figure 3.10, is strongly anisotropic and very similar to those obtained on soft
clays (see Figure 3.2c). Leroueil & Vaughan (1990) hypothesized that the high void ratio of this mud-
stone indicates that it has become microstructured early in its geological history while stresses were low
and water content high, and that the shape of the limit state curve of this strongly microstructured mate-
rial still reflects the anisotropy to be expected of a freshly sedimented clay.
Another interesting case is provided by Lagioia & Nova (1995). These authors report triaxial tests
performed on a calcarenite from the Bari area, in Italy. The results are very peculiar, showing a marked
transition from rock-like to soil-like behaviour and are worth being described here. Figure 3.11a shows
the isotropic compression curve. It can be seen that the behaviour appears to be essentially elastic until
collapse occurs at a mean effective stress of 2.4 MPa. After a significant decrease in void ratio at about
the same stress, the material regains stiffness. CID tests were also carried out. Figures 3.11b and c show
two typical deviatoric stress-axial strain and volumetric strain-axial strain curves for confining stresses
of 200 kPa and 1300 kPa. The corresponding stress paths are shown with dashed lines in Figure 3.11d.
For the tests with the confining stress at 200 kPa, a peak is reached at an axial strain of about 1.5%

43
09031-02[1].qxd 18/Oct/02 12:09 AM Page 44

Figure 3.11. Triaxial tests on calcarenite: (a) isotropic compression; (b) and (c) examples of CID tests; (d) limit
state curve (after Lagioia & Nova, 1995).

(Point (a) in Figure 3.11b) and is followed by some strain softening. For the test with a confining stress
of 1300 kPa, the deviatoric stress increases up to point (c) where there is yielding and breaking down of
bonds until point (d) under an essentially constant deviatoric stress. Then, at an axial strain of about 8%,
q starts increasing until a maximum value is reached at large strain. The limit state curve in this case is
clearly associated with the destructuration of the calcarenite. This limit state curve (Figure 3.11d) is
essentially elliptic in shape and centred on the p axis. Kavvadas et al. (2002*) also found a well-defined
limit state curve of elliptic shape for Corinth marl.
These observations and others summarized by Leroueil & Vaughan (1990), clearly show that the con-
cept of yielding initially defined for reconstitued clays also applies to microstructured natural geoma-
terials. The shape of the limit state curve of a given material reflects its geological history including
diagenetic processes. Whereas it appears to be approximately isotropic for most residual soils and soft
rocks, it is clearly anisotropic for natural soft clays.

3.4 Three-dimensional aspects of limit state surface


The behaviour of soils and rocks described in this paper has been obtained mainly with 2  3, either
in oedometer or triaxial tests, and with the axial stress axis coincident with the vertical direction of the
material in in situ conditions. Few studies have been performed on the influence of the intermediate
principal effective stress, 2, expressed by b  ( 2  3)/( 1  3), and of stress axis rotation charac-
terized by , the direction of the major principal stress to the vertical. The effect of stress axis rotation
is examined in Section 6 on anisotropy. Hereunder, only the effect of b is considered.
Callisto & Calabresi (1998) tested Pisa clay in a True Triaxial Apparatus (TTA). The natural clay was
first consolidated to in situ stress conditions ( a  113.5 kPa; r  75.5 kPa). The overconsolidation
ratio was about 1.6. After consolidation, the clay specimens were subjected to radial effective stress
paths in a constant p plane (p  88 kPa). Conditions at failure are shown in Figure 3.12a with dots and
the bold line. It can be seen that the failure line of Pisa clay is more rounded than that given by the

44
09031-02[1].qxd 18/Oct/02 12:09 AM Page 45

Figure 3.12. Failure envelopes determined from tests in the TTA: (a) Pisa clay (after Callisto & Calabresi, 1998);
(b) Louiseville clay at p/ p  0.29 (after Boudali, 1995).

Mohr-Coulomb criterion. It is similar to the one given by Lade & Duncan (1975) criterion (Equation
2.10 and Figure 2.6). Kirkgard & Lade (1991) and Lade & Kirkgard (2000) performed TTA tests on
undisturbed, but normally consolidated, San Francisco Bay Mud and obtained similar results.
Boudali (1995) performed a series of TTA tests on the highly microstructured Louiseville clay. After
isotropic consolidation to c  0.29 p, radial effective stress paths were followed in a constant
p 0.29 p plane (also see Leroueil et al., 2002*). Results are shown in the octahedral plane in
Figure 3.12b. The failure surface is rounded and well outside the Mohr-Coulomb failure surface corre-
sponding to the normally consolidated friction angle of Louiseville clay. This clearly shows the effect of
microstructure on the 3D-limit state curve. Similar results were obtained by Mitchell & Wong (1973)
and Wong & Mitchell (1975) on Ottawa clay.

3.5 Flow rule


Wong & Mitchell (1975), Graham et al. (1983b), Félix et al. (1985) and many others examined the
development of plastic strains in natural clays. As an example, Figure 3.13 shows the plastic strain
increment vectors obtained at gross yielding of the Winnipeg clay (Graham et al., 1983b). The plastic
strain increment vectors shown on this figure are not perfectly perpendicular to the limit state curve.
Other researchers also found results slightly at variance with normality, but with variable tendencies.
So, considering how uncertain is the determination of plastic strain components and limit state curve,
normality appears to be an acceptable approximation.

3.6 A simple yielding model


Stating that “a soft soil cannot be submitted to stresses in any direction higher than the previous stresses
in that direction without yielding”, Larsson (1977) proposed that limit state curves of natural clays be
schematized by four segments: two corresponding to the strength envelopes in compression and in
extension; and two corresponding to a  p and r  Konc p . Limit state curves of natural soft clays
have been described in Section 3.2. Their shape reflects the stress anisotropy prevailing during soil depo-
sition. However, a good fit between limit state curves and the schematic limit state curves suggested by
Larsson (1977) is obtained when the r  Konc p segment is changed to r  K* p, with K* being
slightly larger than Konc. The following expression was proposed by Leroueil & Barbosa (2000):

K*  0.85  sin(0.46nc) (3.2)

The validity of this approach is confirmed by Figure 3.14 where the schematized limit state curves
based on Equation 3.2 are shown with broken lines for Winnipeg and Mexico City clays. K* values

45
09031-02[1].qxd 18/Oct/02 12:09 AM Page 46

larger than Konc have probably to be associated with the fact that Ko increases during secondary con-
solidation or when strain rate decreases (Section 9.7). Also, limit state curves of soils are certainly more
rounded than those given by the model, but this latter represents a simple and useful approximation. The
parameter K*, as defined by Equation 3.2, is thought to represent the distribution of inter-particle or
inter-aggregate contacts in a soil and is more generally defined as KAL by Leroueil & Barbosa (2000).
The parameter KAL represents the stress ratio on the Anisotropy Line (AL) and reflects soil anisotropy;
as KAL does not seem to be significantly modified by bonding (microstructure) or partial saturation
(Section 14.7), this model [called GFY (for Given Fabric Yielding) by Leroueil & Barbosa, 2000] seems
to be applicable to microstructured and unsaturated soils.
The GFY model is well adapted for describing kinematic hardening. The two segments of the limit
state curve, in between the compression and extension strength envelopes, intersect on the AL line, at

0.7
mb
0.6 oulo
C
ohr-
M
0.4
q/ σ' p

0.2
GFY

0
0 0.2 0.4 0.6 0.8 1.0
p'/σ'p

Figure 3.13. Normalized limit state curve and plastic strain increment directions for Winnipeg clay (from Graham
et al., 1983b).

0.4
Winnipeg
( σ 'a- σ 'r ) / 2 σ p'

17.5 ˚
(φ'nc = 17.5˚) φ' nc =
0.2 AL

(a)
0
0 0.2 0.4 0.6 0.8 1.0
(σ a' + σ r' ) / 2 σ 'p
0.6
˚
Mexico City 43
c=
φ' n
( σ 'a - σ 'r ) / 2 σ p'

( φ''nc = 43˚)
0.4
AL

0.2

(b)
0
0 0.2 0.4 0.6 0.8 1.0
( σa' + σ r' ) / 2 σ 'p

Figure 3.14. Yielding and GFY curves of natural saturated clays (after Leroueil & Barbosa, 2000).

46
09031-02[1].qxd 18/Oct/02 12:09 AM Page 47

a given deviatoric stress when the soil is anisotropic or at zero deviatoric stress when the soil is
isotropic. For example, when applied to the isotropic weathered colluvium studied by Machado & Vilar
(2002*), the GFY model gives the limit state curve shown in Figure 14.22.
The authors also tried to apply the GFY model to the results obtained on Bothkennar clay and presented
in Figure 3.8. The GFY model captures relatively well the behaviour of the intact soil, even if the gross
yield points indicate a KAL value larger than that indicated by Equation 3.2. For the soil consolidated
isotropically in the normally consolidated range, the GFY model captures the limit state curve in the com-
pression domain (q  0) well, but overestimates yield stresses in the extension (q  0) domain. This is pos-
sibly because the soil kept some memory of its initial fabric, even after isotropic consolidation to 2.6 times
the isotropic gross yield stress of the intact soil. It also possibly shows the limits of this simple model.
Figure 3.13 shows the plastic strain increment vectors at yielding for Winnipeg clay. The dashed line
is the limit state curve given by the GFY model. It can be seen that the plastic strain increment vectors
are essentially perpendicular to this latter limit state curve, indicating that normality could also be appli-
cable to the GFY model.

3.7 Some simple applications


The concepts of limit and critical states can be used for understanding the behaviour of natural soils and
soft rocks under a variety of test conditions (Leroueil & Vaughan, 1990). The cases of isotropically con-
solidated drained (CID) triaxial tests and one-dimensional compression tests are described hereunder.
Figure 3.15 shows CID test results obtained on a Champlain Sea clay and on an oolitic limestone.
Tests labelled “1” are at low confining stresses. They show a well-defined peak where the limit state

Figure 3.15. CID tests (from Leroueil &Vaughan, 1990).

47
09031-02[1].qxd 18/Oct/02 12:09 AM Page 48

Figure 3.16. One-dimensional compression tests (from Leroueil & Vaughan, 1990).

curve coincides with shear failure; at larger deformations, there is strain-softening and the material
moves towards its critical state. Tests labelled “3”, at high confining stresses, show stiff behaviour until
the limit state curve is reached, followed by a deviatoric stress progressively increasing towards the crit-
ical state line. In tests labelled “2”, the stress path reaches the limit state curve and the critical state line
concurrently, and the deviatoric stress remains constant after yielding. It is worth noting that these two

48
09031-02[1].qxd 18/Oct/02 12:09 AM Page 49

Figure 3.17. Ko triaxial tests on Saint-Louis-de-Yamaska clay.

materials of completely different origin, and with strengths differing by more than two orders of mag-
nitude, qualitatively present the same behaviour.
Figure 3.16 shows the compression curve and the corresponding stress path obtained from one-
dimensional (Ko) tests performed on four different materials. Within their limit state curves all the
materials show stiff behaviour, with stress paths consistent with “elastic” materials. Theoretically,
 3/ 1  /(1  ) for isotropic elastic materials with Possson’s ratio . When the stress path
reaches the limit state curve, there is yielding and compressibility increases markedly. The stress path
then migrates towards the Ko line corresponding to Konc  (1  sin ). This is a transition phase; the
stress path remains on this Konc line thereafter. The four materials which show these similar patterns of
behaviour are soft clay, artificially bonded soil, marl and chalk, with apparent preconsolidation pres-
sures of 75, 170, 1250 and 40000 kPa, respectively. Again, the behaviour is congruent although the
materials are very different.
This latter behaviour is confirmed by a series of triaxial tests performed at Université Laval on Saint-
Louis-de-Yamaska clay. On this clay, with a preconsolidation pressure of about 180 kPa, two triaxial Ko
tests were performed after isotropic consolidation to 10 and 100 kPa respectively. The stress paths are
shown in Figure 3.17. The three phases are clearly shown with elastic response (I–a), transition (a–b) and
plastic response (b–c). However, while the effective stress ratio r/ a increases during the transition
phase in test 1 ( c  10 kPa) before reaching the Konc line, it decreases in test 2 ( c  100 kPa).

4 BEHAVIOUR INSIDE THE LIMIT STATE CURVE

4.1 Generalities
In previous sections, it has been assumed that the behaviour inside the limit state curve was elastic. It is
not the case. With the development of devices allowing the accurate measurement of local strains (e.g.
Jardine et al., 1984; Lo Presti et al., 2001), it has become possible to examine in more detail the behaviour
of soils inside their limit state curve. Jardine et al. (1991), Jardine (1992) and Hight & Higgins (1994)
schematized this behaviour as shown in Figure 4.1. It can be described as follows for a stress path such
as IABCD:
– There is the outer yield curve Y3 (Figure 4.1a) that is associated with a change in fabric and coinci-
dent with the limit state curve. Soils experiencing stress paths that reach this curve undergo large
plastic strains (Figure 4.1c). Inside Y3, strains are small to moderate.
– Within the inner yield curve Y1 surrounding the current effective stress conditions (zone 1), the
behaviour is linear-elastic and characterized by small strain elastic properties, in particular the small
strain shear modulus Go.
– When a stress path crosses Y1, but remains inside zone 2 (between A and B), the behaviour is non-
linear elastic.
– Between Y2 and Y3 (between B and C in zone 3), soil develops plastic strains. As indicated in
Figure 4.1c, the ratio of plastic to total strain progressively increases as the stress path approaches C

49
09031-02[1].qxd 18/Oct/02 12:09 AM Page 50

Figure 4.1. Scheme of multiple yield surfaces and soil response (modified after Jardine et al., 1991; Jardine, 1992;
and Hight & Higgins, 1994).

on Y3 and stiffness decreases. Passage through Y2, point B, also corresponds to the strain at which
pore pressures start to build up in resonant column or cyclic triaxial tests. From data compiled by Lo
Presti (1989) the strain on Y2 increases with Ip, and also with OCR.
– When Y1 and Y2 are crossed by the effective stress path, they are dragged with the effective stress
state. However, an elastic behaviour is obtained only if there is a sharp change in stress path direction
or if the soil can benefit from some aging.
As an example, Figure 4.2 shows the yield curves Y1, Y2 and Y3 around in situ stress conditions for
Bothkennar clay (Smith et al., 1992; Hight et al., 2002b*).
The yield curves Y1 and Y2 are reached at strains that are smaller than 0.1%, and even smaller than
0.001% in cohesionless or low plasticity soils. The study of soil behaviour in these zones thus necessi-
tates equipment with extremely high resolution. In particular, to avoid system compliance and bedding
and seating errors, it is essential to measure strains locally over the central part of the specimen. Readers
interested to get information and references on this aspect can refer to Lo Presti et al. (2001).
Small strain shear modulus, Go, can also be deduced from shear wave propagation velocity in soils,
by using the following equation:

Go  V s2 (4.1)

where  is the mass density and Vs is the shear wave velocity.


The shear wave velocity can be measured in the laboratory with piezoceramic bender elements
(Dyvik & Madshus, 1985; Viggiani & Atkinson, 1995b) or miniature geophones in the case of large size
triaxial specimens or calibration chamber specimens (Nishio & Tamaoki, 1988; Stokoe et al., 1991).
In situ, shear wave velocity can be obtained from seismic tests. The most widely used invasive techniques

50
09031-02[1].qxd 18/Oct/02 12:09 AM Page 51

Figure 4.2. Yield surfaces identified for Bothkennar clay under in situ stresses (after Smith et al., 1992; Hight
et al., 2002*).

Figure 4.3. Compliance matrix for cross-anisotropic materials (Tatsuoka & Kohata, 1995).

are Cross Hole (CH), Down Hole (DH; Stokoe & Wood, 1972) and seismic cone (SC; Campanella et al.,
1986) tests.
The small strain shear modulus can also be directly measured in resonant column (RC), torsional
shear (TS) or hollow cylinder (HC) apparatuses.

4.2 Small strain behaviour (inside zone 1)


4.2.1 Compliance matrix for cross-anisotropic materials
Inside zone 1, soil behaviour is linear-elastic and can thus be characterized by a linear relationship
between the effective stress tensor and the strain tensor. The compliance matrix linking effective stresses
and strains is as shown in Figure 4.3 for cross-anisotropic materials having the axis of symmetry in the
vertical direction. This matrix comprises 8 parameters that will be discussed in the next paragraphs.

4.2.1.1 Young’s moduli


The elastic Young’s modulus Ex is defined as d x/d x where x and x are respectively the axial stress
and the axial strain in the direction x. Examples are shown for a variety of geomaterials in Figure 4.4.

51
09031-02[1].qxd 18/Oct/02 12:09 AM Page 52

Figure 4.4. Summary of the effects of strain rate on the very small strain Young’s modulus EV (Tatsuoka et al.,
1997; Tatsuoka, 2000).

It can be seen that Ev typically varies from 105 kPa for normally consolidated clays to almost 108 kPa for
hard rocks. In any particular direction x, Ex is a unique function of the normal stress x in the x direc-
tion, independently of the normal stresses acting in the other orthogonal directions. This was suggested
by Hardin (1978) and has been confirmed since. It is well illustrated by the set of data obtained on
Nerima gravel and reported by Kohata et al. (1997), (Figure 4.5). Ev and Eh can thus be written:

m
 ′ 
E v  (E v )o  v  (4.2)
 o′ 

m
 ′ 
E h  (E h )o  h 
 o′  (4.3)

where, o is a reference effective stress and (Ev)o and (Eh)o are Young’s moduli at this reference stress.

52
09031-02[1].qxd 18/Oct/02 12:09 AM Page 53

Figure 4.5. Relationship between: (a) vertical Young’s modulus and vertical stress; and (b) lateral Young’s modulus
and lateral stress, at isotropic stress states and triaxial stress states (Kohata et al., 1997).

Figure 4.6. Summary of relationships between elastic Young’s modulus Eo and confining pressure for uncemented
granular materials (from Kohata et al., 1997).

Figure 4.6 shows normalized Young’s modulus as a function of confining stress for a variety of unce-
mented granular materials. It can be seen that the different curves are essentially parallel, corresponding
to an m value approximately equal to 0.5.
From Equations 4.2 and 4.3, the ratio Ev/Eh can be written as:
m
Ev (E )  v′ 
 v o  ′  (4.4)
Eh (E h )o h

53
09031-02[1].qxd 18/Oct/02 12:09 AM Page 54

Figure 4.7. Stress-system-induced anisotropy for sands and gravels (from Kohata et al., 1997).

Equation 4.4 shows the influence of the stress ratio ( v/ h) and of the inherent anisotropy reflected by
(Ev)o/(Eh)o, which is the ratio of the moduli when stress conditions are isotropic. Figure 4.7 shows some
Ev/Eh ratios versus v/ h relationships for granular soils. From these curves, it can be seen that, as
tested, Ticino sand would be isotropic whereas Chiba gravel would be highly anisotropic with
(Ev)o/(Eh)o  2.5. Other (Ev)o/(Eh)o values can be found in the literature, in particular for clayey mate-
rials (Pennington et al., 1997; Jovicic & Coop, 1997; Rampello et al., 2002*).
It is worth noting that Equations 4.2 and 4.3 are valid only for geomaterials that are not significantly
microstructured. For microstructured soils or soft rocks, Equations 4.2 and 4.3 have to be modified so
that Ev and Eh may have finite values under zero confining stress. The following forms are proposed:

m
 ′ 
E v  E vmo  (E v )o  v  (4.5)
 o′ 

m
 ′ 
E h  E hmo  (E h )o  h  (4.6)
 o′ 

with m is smaller than 0.5 (see Section 11.2.3, Figures 11.5 to 11.7, and also Tatsuoka et al., 2002*)

4.2.1.2 Additional conditions provided by the symmetry of the compliance matrix and the
isotropic properties in the horizontal plane
As indicated by Tatsuoka et al. (1997; 2001), the compliance matrix is often considered symmetric,
which implies:

vh/Ev  hv/Eh (4.7)

and

Ghh  Eh/2(1  hh) (4.8)

54
09031-02[1].qxd 18/Oct/02 12:09 AM Page 55

Figure 4.8. Summary of relationships between normalized Go and applied stresses for reconstituted clays (Hight,
1998).

As shown by Stokoe et al. (1991) and Bellotti et al. (1996) for sands, and by Lo Presti et al. (1999)
and Pennington et al. (1997) for reconstituted clays:

Gvh  Ghv (4.9)

which is expected for a continuum medium.


However, Butcher & Powell (1995) and Pennington et al. (1997) indicated that Equation 4.9 may not
be valid for very stiff or layered clays.
With Equations 4.7, 4.8 and 4.9 (when valid), the number of independent variables of the compliance
matrix decreases to 5: Ev, Eh, vh (or hv), hh and Gvh. Very often Gvh is referred to as Go or Gmax the
small strain shear modulus.

4.2.1.3 Shear modulus


Hardin (1978) suggested that for clays, the small strain shear modulus, Go, depends on applied stresses,
void ratio and overconsolidation ratio (OCR). It has however been shown that the effect of OCR is, to a
large extent, taken into account by the effect of void ratio and could be neglected. The empirical equa-
tion describing the influence of the controlling factors on Go can then be written as follows:

Go  SF(e)( v . h)n Pa(12n) (4.10)

where F(e) is a void ratio function that will be described later on, n is a parameter indicating the influ-
ence of stress, Pa is the atmospheric pressure and S a dimensionless parameter characterizing the con-
sidered soil.
Figure 4.8 shows the shear modulus, normalized with respect to a void ratio function F(e)  e1.3 as
a function of ( v  h ) for a variety of reconstituted clays mainly tested in RCT. It can be seen that n
is close to 0.25 for most of the clays, except for some of the Japanese clays for which it appears to be
slightly larger. Values slightly larger than 0.25 have also been observed by Jamiolkowski & Lo Presti
(2002*) on Messina sand and gravel (n  0.27) and by Miura & Yagi (2002*) on coarse-grained
Touhoro volcanic soil (n  0.34). It may be worth mentioning that the materials showing an n value
larger than 0.25 seem to have very high friction angles (nc of 28–38° for Japanese clays (Tanaka,

55
09031-02[1].qxd 18/Oct/02 12:09 AM Page 56

Figure 4.9. Normalized maximum shear modulus versus void ratio (after Lo Presti, 1989; from Jamiolkowski et al.,
1991).

2002); 35° for Messina sand and gravel; and larger than 40° for Touhoro volcanic soil), indicating that
this latter parameter could influence the value of n.
Also, as for Equations 4.2 and 4.3 for Young’s moduli, the form of Equation 4.10 applies only for
soils that are not microstructured. The reader is referred to Section 11 to find information on the influ-
ence of microstructure on Go.
Figure 4.9, modified after Lo Presti (1989), shows Go values as a function of e for a variety of soils.
The slope which reflects the void ratio function F(e), does not seem to be significantly influenced by the
type of soil. Lo Presti (1989) deduced from this figure the function F(e)  1/e1.3. Several other void
ratio functions have been proposed: Hardin (1978) proposed 1/(0.3  0.7e2); some are in the form
F(e)  1/ex, with x varying between 1.1 and 1.55 (Jamiolkowski et al., 1995a; Shibuya & Tanaka,
1996). Some are of a different form, such as the one proposed by Romo & Ovando (1994) for Mexico
City clay which is expressed as a function of Ip and Ir ((wL  w)/Ip).
S may vary with the type of soil and also, due to the form of Equation 4.10, with the void ratio func-
tion. If, as in the functions previously mentioned, F(e) takes a value of 1.0 for a void ratio equal to 1.0,
then S would be the Go value (in atmospheres) at a void ratio of 1.0 under an isotropic confining stress
of 1 atmosphere. As shown on Figure 4.9, and from other data from the literature, S values are gener-
ally between 200 and 500 for clays that are not significantly microstructured and between 300 and 750
for sands. Weston (1996) and Brignoli (1997) examined the stiffness of Ticino sand and gravel and
found an influence of the uniformity coefficient Cu  (d60/d10). According to the parameters provided
by Jamiolkowski & Lo Presti (1998), S, as defined in Equation 4.10, would decrease from 612 to 372
when Cu increases from 2 to 74, thus essentially in the previously mentioned range.
Hardin (1978) suggested that for clays, the small strain shear modulus, Go, depends on applied
stresses, void ratio and overconsolidation ratio (OCR). Viggiani & Atkinson (1995a) showed that Go
could be defined as follows for isotropically consolidated reconstituted clays:

n m
G0  p′   pp′ 
 A    (4.11)
pr′  pr′   p′ 

where pr is a reference pressure, pp is the stress at the intersection of the swelling line with the isotropic
normal consolidation line and A, n and m are non-dimensional soil parameters. The ratio pp/p is thus
the overconsolidation ratio defined in terms of isotropoic stresses.

4.2.2 Influence of stress history and stress path


When soil follows a stress path such as (1) in Figure 4.10, the yield curves Y1 and Y2 are dragged with
the stress state. Shortly after stopping at point A, the shapes of Y1 and Y2 are as indicated by the dashed

56
09031-02[1].qxd 18/Oct/02 12:09 AM Page 57

Figure 4.10. Development of Y1 and Y2 zones by different recent stress-time histories (from Tatsuoka et al., 1997).

Figure 4.11. Definition of soil stiffness.

line (1). If the stress path followed to reach A is (2), the shape of Y1 and Y2 is as indicated by the dashed
line (2), different from (1). The small strain behaviour of soils is thus kinematic (Atkinson et al., 1990;
Jardine et al., 1991; Jardine, 1992). However, with decreasing strain rate (see Section 9.3) and with
microstructuration, zones 1 and 2 grow in size and may be less dependent on previous stress history
(Tatsuoka et al., 1997; Clayton and Heymann, 2001).

4.3 Non-linear stress-strain behaviour (zones 1 to 3)


When stress conditions in a soil element move out of zone 1 (Figure 4.1), the stress-strain behaviour
becomes non-linear and non-elastic. The non-linearity depends on several factors including the type of
soil, previous stress and strain history, stress path and strain rate. Each practical problem and associated
stress paths have thus to be considered individually. However, there are several aspects of stress-strain
behaviour that seem to be general and will be described in the following paragraphs.
It is first essential to specify some parameters used to describe soil stiffness. As shown in Figure 4.11,
several shear moduli can be defined. Monotonic loading enables to define at a point such as P a secant

57
09031-02[1].qxd 18/Oct/02 12:09 AM Page 58

modulus Gs and a tangent modulus Gt. In case of cyclic loading, these moduli are relevant only to the
loading part of the first cycle. For the subsequent cycles, the defined shear modulus is implicitely the
equivalent shear modulus Geq. As a consequence, the relationship between shear modulus and shear
strain defined in monotonic (static) loading test may be different from the equivalent curve defined in
a cyclic loading test. As indicated in Section 9.3 regarding the influence of strain rate, this phenomenon
could be increased by strain rate effects. This is schematically shown in Figure 4.12a. Figure 4.12b
shows an example obtained on Ticino sand isotropically consolidated to 49 kPa and subjected to cyclic
and monotonic tests. It can be seen that at strains smaller than about 103, in the elastic range, the shear
modulus is essentially the same for all tests. At larger strains, however, the curves associated with cyclic
loading tests are slightly above those obtained in monotonic loading tests. However, several authors
have not observed significant differences between the two types of curves (Tatsuoka et al., 1997).
Figure 4.13 shows the variation of the undrained secant modulus Eu as a function of axial strain for
3 British geomaterials, Chalk, London clay and Bothkennar clay. The initial modulus value varies from

Figure 4.12. Modulus reduction for monotonic and cyclic loadings.

Figure 4.13. Undrained modulus as a function of axial strain for three different British geomaterials (after Clayton
& Heymann, 2001).

58
09031-02[1].qxd 18/Oct/02 12:09 AM Page 59

one material to the other, but also the general shape of the curves and in particular the strain at which
the modulus becomes non-linear.
Figure 4.14 shows normalized shear strain modulus G/Go versus shear strain  for two extreme geo-
materials, namely the highly plastic Mexico City clay and the Messina sand and gravel. Whereas
Mexico City clay shows shear modulus decay at a shear strain of about 0.1%, such a decay starts at a
strain that is less than 103 % for the sand and gravel.
Examining the cyclic test results obtained by different researchers on a variety of clayey soils,
Vucetic & Dobry (1991) produced Figure 4.15 showing normalized shear modulus as a function of the
cyclic shear strain for plasticity index varying from 0 to 200. The tendency indicated by Figure 4.15 has
generally been confirmed since, in particular by the results obtained on the extreme soils that are the
Mexico City clay (Ip  147–238) and the Messina sand and gravel (Figure 4.14). However, as already
indicated by Vucetic & Dobry (1991) and confirmed many times since, “It is important to note that soils
having a very sensitive structure, such as quick clays or cemented sands, may have G/Go curves quite
different from those in Figure 4.15, independently of their Ip”.
Considering a linear strain threshold 1 at G/Go ratio of 0.95, d’Onofrio et al. (1998) defined the
correlation between 1 and plasticity index shown in Figure 4.16. 1 typically increases from 0.001%
for cohesionless materials (Ip  0) to about 0.03% for clayey soils with Ip  100%. The data obtained
on Mexico City clay (Romo, 1995, and Diaz-Rodriguez, 2002*) as well as on Messina sand and gravel
are consistent with the tendency given by Figure 4.16.

Figure 4.14. Normalized shear modulus versus shear strain, curves for two geomaterials.

Figure 4.15. Relation between G/Gmax and for normally and overconsolidated clays (modified after Vucetic &
Dobry, 1991).

59
09031-02[1].qxd 18/Oct/02 12:09 AM Page 60

Figure 4.16. Correlation between plasticity index and threshold strain (defined at G/Go  0.95) (after d’Onofrio
et al., 1998).

The bulk modulus, K  p/ vol is also non-constant (Smith et al., 1992; Zdravkovic & Jardine,
1997). Results of probing tests performed on Bothkennar clay are shown in Figure 4.17a where bulk
modulus normalized with respect to p is plotted as a function of the volumetric strain. They confirm
that K varies with strain, and also indicate that K is stress path dependent. This latter aspect is also
shown in Figure 4.17b. K/p is seen to be greatest along paths where q is constant (  0 and 180°) and
smallest when a change in q dominates (  70° and 110°), (Smith et al., 1992). Other data on non-
linearity of bulk modulus for Gault clay and London clay are reported respectively by Dasari & Bolton
(1998) and by Hight et al. (2002a*).

4.4 Other factors influencing stress-strain behaviour inside the limit state curve
In addition to the previously mentioned factors (applied stresses, type of soil, void ratio, stress history and
stress path), there are other factors such as strain rate and time, temperature, microstructure, matric suc-
tion and discontinuities that may also influence the small strain behaviour of soils. The influence of these
factors will be examined later on in Sections 9.3 for strain rate and time, 10.3 for temperature, 11.2.3 for
microstructure, 14.8 for matric suction and 12 for discontinuities. It may however be of interest to men-
tion here that, due to secondary compression and development of microstructure, soil stiffness generally
increases with time. This is often quantified by the following parameter (Afifi & Richart, 1973):

G0 (t)  G0 (t  t p) 1
NG  ⋅ (4.12)
G0 (t  t p) log (t/t p )

where G0(t  tp) and G0(t) are the small strain shear moduli at the end of primary consolidation (t  tp)
and at time t respectively. Values of NG are given in Section 9.3 and Figure 9.16.
Considerations on soil stiffness in in situ conditions and on relationships between characteristics
deduced from different tests are presented in Part II of this paper, more precisely in Section 17.5.1, and
in the paper by Hight & Leroueil (2002*).

4.5 Stress-strain curves and stress paths inside the limit state curve
4.5.1 Representation of stress-strain curve
Researchers have looked for mathematical representations of the pre-failure stress-strain curve. An
attractive and simple one is the hyperbolic function (Kondner, 1963; Duncan & Chang, 1970). Such a
function can be written in the form:

( 1  3)  /(a  b ) (4.13)

60
09031-02[1].qxd 18/Oct/02 12:09 AM Page 61

Figure 4.17. Variation of K/p during drained probing tests on Bothkennar clay: (a) Ktan/p versus ( vo1; (b) varia-
tion of tangent and secant K/p at vo1  0.01% and 0.1% (after Smith et al., 1992).

that can also be modified as:

/( 1  3)  a  b (4.14)

where a and b, parameters that can be easily defined in a diagram of /( 1  3) vs where Equations
4.13 and 4.14 correspond to a straight line (Kondner, 1963). However, more accurate data show that in
most cases, the hyperbolic function cannot capture the soil response from very small to intermediate
strains, from zone 1 to Y3 (Tatsuoka & Shibuya, 1991; Fahey, 1998).
Another important practical aspect, already mentioned in Section 3.2 (Figure 3.4), is that, in homo-
geneous deposits, for the same OCR, the stress-strain curves can be normalized with respect to the
preconsolidation pressure.

61
09031-02[1].qxd 18/Oct/02 12:09 AM Page 62

4.5.2 Stress paths inside the limit state curve


An observation made on reconstituted clays (see Wroth & Loudon, 1967), but also on microstructured
clays (Cotecchia & Chandler, 2000; Cotecchia, 2002*; Leroueil et al., 2002*) is that stress paths corre-
sponding to linear elastic behaviour, i.e. at essentially constant p(Skempton’s A parameter  0.33), are
obtained in undrained tests in stress planes such that p is equal to 0.3 to 0.5 p, independently of the
microstructure of the soil.

5 BEHAVIOUR OF COHESIONLESS SOILS AND CRITICAL STATE

The state of sedimented clayey soil is primarily controlled by mineralogy, stress history and aging
processes. The state of cohesionless soils depends not only on stress history, but also on the mode of
deposition or of mechanical compaction, and is generally characterized by the relative density.

5.1 Behaviour of cohesionless soils under drained and undrained conditions


Because of their relatively high hydraulic conductivity, the drained behaviour of sands and gravels
attracted the most attention historically. Researchers became interested in the undrained behaviour of
sands and gravels when they realized that such hydraulic conductivity may not be sufficient to prevent
the development of excess pore pressures and possibly failures (liquefaction under gravity or earthquake
loading).

5.1.1 Drained conditions


The drained behaviour of sands received great attention in the sixties (Ladanyi, 1960; Rowe, 1962;
Bishop & Green, 1965; Lee & Seed, 1967; Vesic & Clough, 1968). A very detailed study was performed
on the Sacramento sand by Lee & Seed (1967). These authors performed 4 series of drained triaxial
compression (CID) tests at confining stresses varying from 0.1 to 13.7 MPa on sand specimens pre-
pared at four different relative densities from 38 to 100%. Figures 5.1a and b show the principal stress
ratio and the volumetric strain as a function of the axial strain for the soil prepared at a relative density
of 38%, and Figures 5.1c and d show the same curves for the soil prepared at a relative density of 100%.
For the loosest soil (Figures 5.1a and b), the stress strain curves generally do not show any peak and
the volumetric strain is generally in compression. This behaviour is mostly associated with loose mate-
rials and is referred to as “loose behaviour”. The test results obtained on the dense sand (Figures 5.1c
and d) show, for confining stress smaller than 2 MPa, stress-strain curves generally reaching a peak fol-
lowed by strain softening, and a volumetric strain that is negative. It is also worth noting that the rate of
dilation (d vol/d 1) is at its maximum when the peak is reached (see arrows). This type of behaviour is
referred to as “dense behaviour”.
Closer examination of Figure 5.1a shows that, under small stresses ( 3c  0.5 MPa) the soil with
Dr  38% shows a “dense behaviour” with 1/ 3 ratio at the peak larger than the same ratio at large
strain. Figure 5.1c shows that the dense soil (Dr  100%) does not present any peak, and thus presents
a “loose behaviour” under large stresses ( 3c  4 MPa). For all the tests, 1/ 3 at large strains is about
the same, at a value of 3.4.
The general drained behaviour is schematically shown in Figure 5.2. At a given density, the soil has
a curved peak strength envelope and a strength envelope at large strains (critical state line, CSL) char-
acterized by the slope M in the q–p diagram or cs (Figure 5.2a). For different densities, the peak
strength envelopes are as shown in Figure 5.2b, but the CSL is unique.
From the studies reported by Lee & Seed (1967) and others, it appears that the drained shear behav-
iour of sands depends mainly on density and stress. As previously indicated, the peak strength of cohe-
sionless soils is associated with the maximum rate of dilation and reflects the energy necessary for the
soil specimen to change its volume.
Rowe (1962) proposed the following equation to link stress conditions during shearing, with the rate
of dilation, including at large strains when there is no volume change (critical state).

1′  ′   d 
  1 ⋅ 1  vol  (5.1)
3′  3′  crit  d 1 

where ( 1/ 3)crit is the stress ratio at critical state, vol is the volumetric strain and 1 is the major prin-
cipal strain. (d vol/d 1) is the rate of dilation.

62
09031-02[1].qxd 18/Oct/02 12:09 AM Page 63

Figure 5.1. Typical drained triaxial test results on Sacramento River sand: (a) and (b) respectively principal stress
ratio versus axial strain and volumetric strain versus axial strain for loose sand; (c) and (d) respectively principal
stress ratio versus axial strain and volumetric strain versus axial strain for dense sand (after Lee, 1965).

Figure 5.2. Variation of drained shear strength envelope for sand with: (a) confining pressure, and (b) density.

63
09031-02[1].qxd 18/Oct/02 12:09 AM Page 64

When stress conditions at the peak are corrected for dilation (Equation 5.1), the peak strength
envelopes shown in Figure 5.2 collapse to come closer to the CSL. Rowe (1962), Lee & Seed (1967)
and others mention that some shearing energy also has to be attributed to particle rearrangement and
crushing.
It is important to mention that, although the influence of dilatancy on strength has been mostly
studied on sands, Equation 5.1 is general and also applies to other geomaterials.
Bolton (1986) examined drained test results obtained on 17 sands with friction angle at critical state,
cs, varying from 30 to 37°. To define the combined effect of relative density and stress, he proposed a
relative dilatancy index defined as follows:

IR  Dr(Q  ln p)  R  Dr(10  ln p)  1 (5.2)

where Dr is the relative density, p is the mean effective stress (in kPa in Equation 5.2), and Q and R are
empirical factors found equal to 10 (for quartz and feldspar minerals) and 1 respectively. The Q value
depends on the grain crushing strength and would decrease to 8 for limestone, 7 for anthracite and 5.5
for chalk (Bolton, 1986).
Bolton (1986) found the following correlations for IR values between 0 and 4.
For plane strain conditions:

max  crit  5IR (°) (5.3a)

For triaxial strain conditions:

max  crit  3IR (°) (5.3b)

The friction angles considered in these equations are secant friction angles obtained by dropping a
tangent from the origin on to a single Mohr circle of effective stress (c  0).
For both test configurations,

(d vol/d 1)max  0.3IR (5.4)

Figure 5.3 illustrates Equation 5.3b. Test results obtained for relative density of about 0.5 and about
0.8 are shown in the same figure. It can be seen that the agreement is quite good, with a typical depar-
ture of less than 2°.

5.1.2 Undrained conditions


By definition, there is no dilatancy in saturated soils when tested under undrained conditions. The range
of possible soil response under such conditions has been summarized by Robertson & Ahmadi (2002*)
and in Figure 5.4a, showing idealized stress-strain curves deduced from undrained compression tests on
specimens initially consolidated under anisotropic conditions. Line A represents the weakest type of
response in which the sample reaches a peak strength and then strain-softens to a minimum strength at

Figure 5.3. Dilatancy component for sands failing at various mean effective stresses (after Bolton, 1986).

64
09031-02[1].qxd 18/Oct/02 12:09 AM Page 65

ultimate state. Line D represents the strongest type of response in which the sample continues to strain-
harden throughout the test, eventually reaching a much larger strength at ultimate state. Lines B and C
represent the two intermediate responses. Line B illustrates the type of response associated with limited
strain-softening to a quasi-steady-state (QSS) and then strain-hardening towards ultimate state. The
change in behaviour occurs on what Ishihara et al. (1975) call the “phase-transformation line”. Along
line C, the sample continuously strain-hardens, even if, once the initial peak is reached, the stress-strain
curve shows a plateau. Types A and B behaviour present a minimum strength that is smaller than the
peak strength. Stress paths are shown in Figure 5.4b for tests A and D. After reaching the peak at point
P, below the critical state line, the stress path moves towards the ultimate state situated on the critical
state line. As for test D, after showing a decrease in p, the stress path moves up along the critical state
line towards its ultimate state. Soil response depends on void ratio, soil fabric, initial stress conditions
and the type of test (compression, extension or simple shear).

5.1.3 Strain path tests


Chu & Lo (1994) performed triaxial strain path tests with constant (d vol/d 1) and stress path tests with
constant effective stress ratio, on saturated Sydney sand. Figure 5.5 shows some results in a q–p

Figure 5.4. Possible responses to triaxial undrained monotonic loading of sandy soils (Figure a is from Robertson
& Ahmadi, 2002*).

Figure 5.5. Strain path testing results: (a) Stress paths and d vol /d a  cst lines; (b) Relationship between effective
stress ratio and strain increment ratio, Equation 5.5 (after Chu & Lo, 1994).

65
09031-02[1].qxd 18/Oct/02 12:09 AM Page 66

diagram. The effective stress path resulting from shearing along a constant strain increment ratio path
test asymptotically approaches a constant stress ratio line. The corresponding slope in the q–p dia-
gram was termed (q/p)asy. Chu & Lo (1994) showed that (q/p)asy depends on the strain increment
imposed to the soil (Figure 5.5b). The relationship shown in Figure 5.5b corresponds to the following
equation:

q  1 d 
 p′   M0 ⋅ 1  ⋅ vol  (5.5)
asy.  3 d 1 

These strain path test results apparently link in a more general manner than Rowe’s approach the rate
of dilation to strength envelope (or mobilized stress ratio 1/ 3 or q/p). However, Equation 5.5 found
by Chu & Lo (1994) is different from Equation 5.1 from Rowe (1962), which could indicate that the
influence of rate of dilation is stress path dependent.

5.2 Critical state, steady state and state parameter


According to Roscoe et al. (1958), critical state could be defined as that ultimate state of a sample at
which any arbitrary further increment of shear distortion will not result in any change in void ratio or
effective stress. Poulos (1981) defined the “steady state of deformation” as follows: “The steady state
of deformation” for any mass of particles is that state in which the mass is continuously deforming at
constant volume, constant normal effective stress, and constant velocity. The steady state of deforma-
tion is achieved only after all particle orientation has reached a statistically steady state condition and
after all particle breakage, if any, is complete, so that the shear stress needed to continue deformation
and the velocity of deformation remain constant.”
Although the two definitions appear very similar except, possibly, for the reference made to constant
velocity of deformation in the case of the steady state, there has been discussion on whether the critical
state and steady state lines were the same or not (Casagrande, 1975; Poulos, 1981; Sladen et al., 1985;
Alarcon-Guzman et al., 1988; Been et al., 1991). According to Been et al. (1991), “the differences
appear to lie in the methods of measurement. Critical state workers have generally relied on drained,
strain-controlled tests on dilatant samples to determine the critical state. The steady state is measured in
undrained tests, usually on loose (contractive) samples.” Been et al. (1991) performed drained and
undrained monotonic triaxial compression tests on Erksak sand and found that the critical and steady
state lines were coincident. The authors believe that both critical and steady states are identical. In a
e–log p diagram, the corresponding critical state line (CSL) or steady state line (SSL) is often char-
acterized by its slope (Figure 5.6).
To characterize the behaviour of sands, Been & Jefferies (1985) proposed the state parameter 
defined as the difference between the current void ratio eo and the void ratio on the steady state line ess
under the same mean effective stress p (Figure 5.6).
Void ratio e

eo A

ψ = eo - ess
e ss
1
λ

Steady state
line (SSL)

log p'

Figure 5.6. Definition of state parameter (after Been & Jefferies, 1985).

66
09031-02[1].qxd 18/Oct/02 12:09 AM Page 67

The state parameter  neglects a cohesionless soil characteristic considered as extremely important
historically, i.e. (emax  emin) in which emax and emin are respectively the maximum and the minimum
void ratios of the sand, as defined by usual standards. To take this aspect into account, Konrad (1988)
suggested a normalized state parameter, N, defined as follows:

N  /(emax  emin) (5.6)

In a manner similar to sedimented clays which are well characterized by their overconsolidation ratio,
sands could have been characterized by pcs/po, when po is the current mean effective stress and pcs is
the mean effective stress on the critical or steady state line at the same void ratio. As noted by Hird &
Hassona (1986), this ratio is related to the state parameter. It can be shown that:

pcs/po  exp(2.3/ ) (5.7)

The normalized stress paths shown in Figure 5.7 indicate that this could be a reasonable approach.
Test 45 shows a “normally consolidated behaviour” with a stress path that goes towards its critical state
CS by decreasing p; on the other hand, tests 103 and 113 show an “overconsolidated behaviour”, with
strain hardening towards the same critical state.
Previous comments are based on the assumption that the critical/steady state line is unique in an
e–log pdiagram. Several investigators have however questioned this uniqueness:
– Performing undrained compression and extension triaxial tests on loose Ottawa sand prepared by
water pluviation, Vaid et al. (1990) observed larger shear-induced pore pressures, and consequently
smaller p values and strength values at large strain in the case of extension tests. They concluded that
the critical state line is stress path dependent. This has been confirmed by Yoshimine et al. (1999),
and also by Shibuya & Hight (1987) and Hight et al. (1987) who performed undrained hollow cylin-
der tests on isotropically consolidated specimens of loose Ham River sand with b  ( 2  3)/
( 1  3)  0.5 and principal stress direction  from 0° to 90°. Hight et al. (1987) showed that the
large strain strength is strongly influenced by , progressively decreasing when  varies from 0 to 90°.
– On the basis of undrained compression triaxial tests, Konrad (1990, 1993) and Doanh et al. (1999)
indicated that the critical state line is not unique, but depends on the consolidation stress. As schema-
tized in Figure 5.8, for specimens consolidated at the same void ratio and over a range of stresses
leading to contractant behaviour, relatively small confining stresses (test B) give large strain shear
strength values which are smaller than those obtained for higher confining stresses (test A).
According to Konrad (1990, 1993), rather than being a unique line, the critical states define a band
in the e–log p diagram limited by upper and lower limits referred to as UF and LF lines in Figure 5.8.

5.3 Soil liquefaction and instability


Sladen et al. (1985) gave a definition of liquefaction in the context of field conditions. Adapted to lab-
oratory testing, liquefaction could be defined as a phenomenon in which there is a reduction in strength

3 350/2 (2% fines)


350/10 (10% fines)

2
45
q/p'cs

CS 37
1

113
103
0
0 1 2 3 4 5
p'/p'cs

Figure 5.7. Normalized stress paths for undrained triaxial tests on Kogyuk sand (after Been & Jefferies, 1985).

67
09031-02[1].qxd 18/Oct/02 12:09 AM Page 68

Peak envelope

q/p'UF
Failure
MP
envelope
1
MF
1 3

UF

LF
2

1 B C A
(a) p'/p'UF
log p'

a A
b B
c C
UF
lin e
LF
lin e
e

(b)

Figure 5.8. Conceptual framework for strain-softening behaviour of loose sand samples (after Konrad, 1993).

observed after failure reached by monotonic, cyclic or shock loading. Whereas up to the point of fail-
ure loading may be undrained, drained or partially undrained, the post-failure stage is generally
undrained or only partly drained. The behaviour shown by the test A in Figure 5.4, with a peak at P and
strain softening is a liquefaction process.
Sladen et al. (1985) studied the fine uniform Nerlerk sand (with different percentages of fines) and
the coarser Leighton Buzzard Sand. The specimens were prepared loose (apparently by moist tamping),
isotropically consolidated and subjected to undrained triaxial compression. The results obtained in the
different series were very similar. Figure 5.9 shows effective stress paths normalized with respect to the
mean effective stress at the critical state, pcs, at the same void ratio for Leighton Buzzard sand. For con-
solidation stress po equal to about twice pcs, CIU tests show almost no strain softening and reach the
critical state (CS on the figure) corresponding to an M value of 1.2 (i.e. c s  30°). For higher nor-
malized consolidation stresses, e.g. po/pcs  6.9, the deviatoric stress increases, reaches a peak at point
P and then decreases towards the critical state CS.
All the peaks of the normalized stress paths fall on the same straight line (Figure 5.9). In a q–p–e dia-
gram, this line forms a surface called the “collapse surface” by Sladen et al. (1985). The collapse
surface is shown in Figure 5.10 (Sasitharan et al., 1993). It can be seen that it comes as a complement
to the conventional critical state models. Alarcon-Guzman et al. (1988) and Sasitharan et al. (1993)
confirm the concept of collapse surface but their results indicate that, rather than being generated
by straight lines as suggested by Sladen et al. (1985), the surface would be curved and defined by mono-
tonic undrained stress paths. The collapse surface would then be a state boundary surface followed
by soil elements that reach it and are moving towards the critical state line.
Sasitharan et al. (1993) performed constant shear drained (CSD) tests, i.e. triaxial tests at constant
deviatoric stress and decreasing mean effective stress, on loose Ottawa sand. Figure 5.11 shows the

68
09031-02[1].qxd 18/Oct/02 12:09 AM Page 69

Figure 5.9. Series of monotonic undrained triaxial compression tests on loose Leighton Buzzard sand showing the
collapse surface and the critical state (modified from Sladen et al., 1985).

Figure 5.10. Typical collapse surface in q–p–e space (from Sasitharan et al., 1993).

Figure 5.11. State boundary or collapse curve at e  0.809 and CSD test (q  100 kPa) on Ottawa sand (after
Sasitharan et al., 1993).

69
09031-02[1].qxd 18/Oct/02 12:09 AM Page 70

results of one of these tests. The initial void ratio was 0.804 under p  300 kPa and q  100 kPa. The
specimen collapsed when p reached 140 kPa; the void ratio was then 0.809. It can be seen from
Figure 5.11 that the stress conditions at that time were very close to the collapse line corresponding to
this void ratio, confirming the possibility of soil collapse at a mobilized friction angle much smaller
than the friction angle at failure, cs for loose sands. According to Sladen et al. (1985), the collapse sur-
face is “the limit of stability if drainage is impeded under load-controlled conditions”.
From Sladen et al. (1985)’s data (Figure 5.9), it turns out that, if stress conditions on the collapse sur-
face are described by an “angle of mobilization mob” (e.g. corresponding to M6.9 for the test consoli-
dated at 6.9 pcs in Figure 5.9), this angle of mobilization would decrease from cs for po/pcs ⬵ 2 to a
value corresponding to M6.9 for po/pcs  6.9. Data presented by Castro et al. (1982) and Been &
Jefferies (1985) also indicate a decrease of mob as po/pcs increases. On the other hand, Vaid & Chern
(1985) and Lade (1993) came to the conclusion that mob would be a constant, that the peaks obtained
in CIU and CAU tests would be on a line, termed the instability line by Lade (1993), intersecting the
origin. However, instability can only occur above a cut-off value of ( 1  3) that increases with den-
sity (Vaid & Chern, 1985; Lade, 1993; Vaid & Eliadorani, 1998).
The state parameter  defined in Figure 5.6 has been used to evaluate the behaviour of soils. Soil ele-
ments with state conditions below the CSL (  0) will always have a dilatant behaviour in undrained
tests whereas soil elements with state conditions above the CSL (  0) will have a contractant behav-
iour. If  is large enough, some strain softening during undrained loading may occur. The state param-
eter has also been used for interpreting in situ tests and predicting debris flows (Lee et al., 1988).
However,  has been defined from undrained tests often performed after isotropic consolidation and
is thus associated with a given type of test. It may not apply to other types of test. This can be illustrated
by results of CSD (Constant Shear Drained) tests, triaxial tests performed at constant deviatoric stress
and decreasing mean effective stress.
Anderson & Sitar (1995) and Anderson & Riemer (1995) performed triaxial tests on a clayey collu-
vial soil from Briones Park, near San Francisco. Anderson & Sitar (1995) anisotropically consolidated
specimens above the steady state line in a e–log p diagram and carried out undrained compression
(CAU) tests. They observed contractant behaviour, with an ultimate deviatoric stress generally smaller
than the initial deviatoric stress. Anderson & Riemer (1995) consolidated the same soil to approxi-
mately the same conditions before subjecting it to an essentially constant deviatoric stress and decreas-
ing mean effective stress (Stress path such as IY in Figure 5.12a). As shown in Figure 5.13, dilatancy
was observed rather than the expected contractancy. Santos et al. (1996) observed similar behaviour on
a residual soil from Rio de Janeiro.
Leroueil (2001) explained this behaviour as follows (see schematic Figure 5.12). If initial stress con-
ditions are at point I, in Figures 5.12a and b, and if the mean effective stress is decreased, there is first
some volume change of the soil, generally swelling, and yielding is reached at a point such as Y. This
latter point can be on the critical-state line (CSL), as shown in Figure 5.12a, or above in microstructured
or dense soils, or below, on the collapse surface, as observed in loose sands (Figure 5.9). If initial con-
ditions are at point I1, well above the CSL, the soil at yielding, Y1, is above the CSL (Figure 5.12b); for

Figure 5.12. State paths followed in CSD tests (schematic). In (a), Y could be above, on or below the critical-state
line, depending on the soil considered.

70
09031-02[1].qxd 18/Oct/02 12:09 AM Page 71

undrained post-failure conditions, the ultimate strength, at C1 is smaller than the initially applied shear
stress, and yielding is followed by an undrained collapse and flow of the soil mass. If initial conditions
are at point I3, below the CSL, yielding will appear at Y3 (Figure 5.12b). The ultimate state in undrained
conditions, C3, then corresponds to strength larger than the shear stress at yielding and is thus stabilis-
ing. For the intermediate case with initial stress conditions, I2, above the CSL and yielding, Y2, below
(Figure 5.12b), the soil would be dilatant with a tendency to swell towards the CSL. The colluvium soil
studied by Anderson & Sitar (1995) and Anderson & Riemer (1995), and the residual soil tested by
Santos et al. (1996) seem to correspond to case 2, with contractant behaviour in undrained CAU tests
and dilatant behaviour in drained CSD tests. In such tests, the potential for soil instability is thus not
controlled by the state parameter I2 associated with initial state conditions but by the equivalent param-
 
eter at yielding, I2 (Figure 5.12b). To avoid confusion, Chu et al. (2002) call  the “modified state
parameter”.
It can be added that, if initial stress conditions had been at point I2, with a much larger state param-
eter, I2, the soil behaviour in a CSD test would have been the same, namely dilatant. The state param-
eter is thus not a parameter that can characterize soil behaviour in CSD tests. This could be the same for
other tests or stress paths.
This remark made by Leroueil (2001) led Chu et al. (2001, 2002) to examine soil instability in a dif-
ferent way. These authors examined instability of marine dredged sand in loose to dense state condi-
tions, associating instability to the development of large plastic axial and volumetric strains. CIU and
CSD triaxial tests were performed. Figure 5.14 shows some results in q–p and e–p diagrams. The

critical state line and the instability lines corresponding to  values of 0.060 and 0.186 are also shown
as references.
The behaviour of loose sand can be illustrated by test DR7 (also test DR10). The specimen was firstly
sheared to point A along a drained path. The deviatoric stress at point A was q  150 kPa. The mean
effective stress was progressively decreased along an essentially constant q path from A to B to C. There
were little axial and volumetric strain development until point B where both axial and volumetric strains

started to develop at a faster rate. B is on the instability line for this soil (  0.060). With further
reduction in p, the stress path moved further towards the CSL that would correspond to the failure line
for this loose specimen. However, the test had to be stopped when the axial and volumetric strains were
respectively 8% and 2% at point C. Therefore, instability with development of large plastic strains can
be observed under fully drained conditions. Leong (2001) and Chu et al. (2002) term this type of insta-
bility “conditional instability” in contrast with undrained instability termed “runaway instability”.
However, if the pore pressure cannot dissipate freely, the specimen would not be able to support the
imposed q anymore and “runaway instability” would occur. This is in particular the case if a specimen
is brought under drained conditions above the instability line and has its drainage valve closed; as indi-
cated by Chu et al. (2002) and Lade (1995), instability can develop due to increase in pore pressure by
creep.

Figure 5.13. State paths in CSD tests performed on colluvial soil from Briones Park (from Anderson & Riemer,
1995).

71
09031-02[1].qxd 18/Oct/02 12:09 AM Page 72

Figure 5.14. General instability behaviour of loose and dense sand along a CSD path (after Chu et al., 2002).

Test DR39 (also test DR40) was conducted on a dense specimen under q  303 kPa from initial
stress conditions below the CSL (point D). As shown in Figure 5.14a, p was progressively decreased
from D to E to F to G. Axial and volumetric strains started to increase at point E (Figure 5.14b) in a dilat-

ant manner, at  close to 0.186. As shown in Figure 5.14a, instability (defined as the development of
large plastic strains) is above the CSL. From E, the mean effective stress continued to decrease until the
failure line associated with its rate of dilation was reached, at point F. If during this process, dilation
could not be accommodated, such as in undrained conditions, negative pore pressure would tend to
develop, leading to an increase in effective stress and the specimen would remain stable.
From CIU and CSD tests, Chu et al. (2002) were able to define q/p on the instability line, (q/p)ins,
 
as a function of  (Figure 5.15). The instability curve defined that way crosses the   0 line at a q/p
value of 1.35, corresponding to the critical state friction angle. It can be seen that the instability curve
seems to approach asymptotically some limiting q/p values when goes towards large positive or nega-

tive values. From the data available, (q/p)ins would be approximately constant and equal to 0.8 for 
values greater than 0.08. This seems to be in agreement with (and possibly explains) the results of Vaid
& Chern (1985) and Lade (1993) that show an essentially constant mobilized friction angle at peak,
mob, above a cut-off value of ( 1  3).
It is worth noting that the difference between instability and failure also exists in natural clays for
which ( 1  3 )max is lower than ( 1  3 )max (Bjerrum & Simons, 1960) and that the concept of
collapse surface could be applicable to geomaterials other than sand.

72
09031-02[1].qxd 18/Oct/02 12:09 AM Page 73

Figure 5.15. Instability curve from undrained and CSD tests on sand (after Chu et al., 2002).

Figure 5.16. Effect of specimen reconstitution method on undrained simple shear response (after Vaid et al., 1995).

5.4 Influence of specimen preparation and soil fabric


The influence of specimen preparation on the cyclic stress-strain behaviour of sands has been well
established (e.g. Mulilis et al., 1977; Miura & Yagi, 2002*, for coarse-grained volcanic soils). It is, how-
ever, only recently that similar effects have been well documented for static behaviour. Vaid et al. (1995)
studied the undrained behaviour of loose fine uniform Syncrude sand on specimens reconstituted by
moist tamping (MT), air pluviation (AP) and water pluviation (WP). Figure 5.16 shows comparative
results obtained in simple shear tests performed on specimens with the same void ratio. The moist
tamped specimen is highly contractive; the air pluviated specimen is also strain softening but to a much
smaller degree. In contrast, the water pluviated specimen behaves in a strain hardening (dilative) man-
ner. At a shear strain of 15%, the shear stress mobilized on the WP specimen is 10 times larger than the
shear stress mobilized on the MT specimen. Vaid & Sivathayalan (2000) observed similar behaviour
when performing undrained triaxial compression and extension tests on Fraser River sand.

73
09031-02[1].qxd 18/Oct/02 12:09 AM Page 74

Hoeg et al. (2000) compared the undrained behaviour of natural silt and a silty sand tailings in both
undisturbed and reconstituted conditions. The natural silt was from a 10,000-year-old fluvial deposit
and the mine tailings were deposited hydraulically 5 years prior to the investigation. About half of the
26 undrained compression triaxial tests presented were performed on reconstituted specimens, gener-
ally prepared by moist tamping. In all cases, the undisturbed specimens showed dilative and ductile
behaviour, whereas in all but a few cases, the reconstituted specimens showed contractive behaviour. A
typical comparison is shown in Figure 5.17 for silt. Even if the relative density was smaller (e  0.81,
Dr  66%) than that of the reconstituted specimens (e  0.73, Dr  73%), the undisturbed specimen
showed a dilative behaviour with continuous strength increase, whereas the moist tamped specimen
showed a low peak and brittle behaviour. Hoeg et al. (2000) also showed that with the same grain size
distribution and void ratio, reconstituted silty sand tailings specimens had Go values that were approxi-
mately 25% lower than those of undisturbed specimens.
Differences in behaviour are associated with differences in fabric between specimens prepared by
different methods or between undisturbed and reconstituted soil specimens. In particular, it is thought
that moist tamping generates aggregates of particles with relatively large voids in between and this can
be viewed as a potentially collapsible fabric. This is similar to what happens when soils are compacted
at degrees of saturation smaller than the one at the optimum (see Figures 14.6 and 19.2 as well as related
comments). Vaid & Sivathayalan (2000) consider that water pluviation more closely replicates the fab-
ric of water deposited in situ sands. Indeed, for two sands, they observed very similar undrained soil
response on undisturbed and WP specimens.
Mimura (2002*) reports for 4 Japanese sands comparative triaxial monotonic loading and cyclic
loading tests performed on undisturbed specimens sampled by the freezing technique and on soil spec-
imens reconstituted to the same void ratio by air pluviation. In all cases, the undisturbed soil shows a
higher resistance, in particular to liquefaction by cyclic loading. This can be due to aging, but can also
be due to differences in fabric.

Figure 5.17. Comparison of typical stress-strain curves and pore pressure development for undisturbed, moist
tamped and slurried specimens of natural silt (after Hoeg et al., 2000).

74
09031-02[1].qxd 18/Oct/02 12:09 AM Page 75

All these observations show that the behaviour of cohesionless materials is extremely sensitive
to fabric.

5.5 Influence of anisotropy and stress axis rotation


This aspect is examined in Section 6.

5.6 Influence of grain crushing


Compression of granular materials to high confining pressures and subsequent shearing induce grain
crushing (Lee & Seed, 1967; Vesic & Clough, 1968; Colliat-Dangus et al., 1988). This is illustrated in
Figure 5.18b showing grain size distributions of Chattahoochee River sand (Vesic & Clough, 1968)
after isotropic compression to 2.1, 21 and 63 MPa (dashed lines) and after shearing under confining

Figure 5.18. Drained triaxial shear tests performed on loose and dense Chattahoochee River sand: (a) Isotropic
normally consolidated zone and critical state line; (b) Grain size distributions after compression and shear (after
Vesic & Clough, 1968).

75
09031-02[1].qxd 18/Oct/02 12:09 AM Page 76

0.80 e max

e cs 1
λ
0.75

Void ratio e
C
ec ψ
Void ratio e

0.70 1
p'
eo Log cs λc
ψ>0ψ<0 O p'o
0.65 Undrained
Load-contr., compr.
Strain-contr., compr.
0.60 CSL
Strain-contr., ext.
Drained, compr.
e min
0.55
10 10 2 10 3 10 4 Log p'o Log p'c Log p'cs
p' (kPa) Log p'
a) Erksak 330/0.7 sand (after Been etal., 1991) b) Schematic behaviour(after Konrad, 1997)

Figure 5.19. Influence of crushing on critical state lines.

stress of 21 and 63 MPa. This phenomenon influences the measured friction angle (Lee & Seed, 1967);
it also influences the critical state line in the e–log p diagram. The compression lines for the loose
and dense Chattahoochee River sand as well as the critical state line deduced from the drained triaxial
shear tests reported by Vesic & Clough (1968) are shown in Figure 5.18a. Even if the critical state line
is approximate, there are several noteworthy observations: (a) it shows a knee somewhere between 1 and
2 MPa, and this is probably associated with crushing (Figure 5.18b); (b) the critical state line and the
normal compression line are curved under high stresses, which is logical considering the fact that the
void ratio becomes very small and cannot take negative values; (c) the ratio of the mean effective
stresses on the normal compression line, pe , and on the critical state line, pcs, at the same void ratio
seems approximately constant, which is consistent with CSSM.
The effect of crushing on the critical state line has been examined by Been et al. (1991) (Figure 5.19a)
and Konrad (1997). As schematically shown in Figure 5.19b from Konrad (1997), the slope of the crit-
ical state line abruptly changes at a mean stress pc corresponding to the crushing stress of soil particles.
This crushing stress typically varies between 500 and 2000 kPa, depending on the mineralogy, the angu-
larity and the grain size distribution of the material. The critical state line can be characterized by for
mean stresses smaller than pc and c for higher stresses.
The bilinearity of the critical state line has consequences for the meaning of the state parameter .
Again, one may suggest that the pcs/po ratio should be considered in place of or in combination with
 to describe the mechanical behaviour of cohesionless soils.
Materials particularly prone to crushing are coarse-grained volcanic soils and carbonate sands that
are described in this Workshop by Miura & Yagi (2002*) and Coop & Airey (2002*) respectively. When
uncemented, these materials behave essentially as sands. Shimizu (1999) and Miura & Yagi (2002*)
indicate however that volcanic soils show some particular features that can be summarized as follows:
•  is generally high, between 40 and 50°.
• The friction angle has a tendency to decrease as the confining stress and thus crushing increase. The
evolution of crushing with confining stress depends on particle hardness.
• Due to crushing, the intermediate effective stress 2 developed in plane strain conditions is generally
smaller than in other sands.
• Due to crushing, Go may increase with stress more rapidly than in other sands. Due to the fact that
particles may contain internal voids, the void ratio functions such as those described in Section
4.2.1.3 may not apply in coarse volcanic soils (Miura & Yagi, 2002*).

6 INFLUENCE OF ANISOTROPY

6.1 Introduction
Civil engineering construction almost invariably involves changes to the stress state in the ground,
resulting in movements and requiring stability under the new stress regime to be considered. We are

76
09031-02[1].qxd 18/Oct/02 12:10 AM Page 77

r0

r/r0
1.5 0.5 0.5 1.5
0
0.2

0.25
0.1 0.1
0.3 0.5

0.3

0.15 0.25 1.0 0.15


0.2 Z/r0 0.2
(a) Contours of ( ∆σ1-∆σ3)
2p

+ ve
α
∆σ1

(b) Contours of α (c) Contours of ∆ b


Figure 6.1. Description of stress changes beneath vertically loaded circular footing on isotropic linear elastic
foundation (  0.45).

familiar with these stress changes in terms of the changes in deviator or shear stress that may take place;
an example is presented in Figure 6.1a as contours of ( 1   3)/2p associated with a pressure, p,
applied to a circular footing on the surface of an isotropic elastic half space;  1 and  3 are the
changes in major and minor principal stresses, respectively. We are probably less familiar with both the
changes in the directions of these principal stress increments and the relative magnitude of the interme-
diate principal stress increment,  2. Figure 6.1b shows contours of the direction of  1, expressed
as an angle,  to the vertical, and Figure 6.1c shows contours of ( 2   3)/( 1   3), or b, both
for the same case of the uniformly loaded circular footing. (For each parameter the distributions are
symmetrical about a vertical axis through the centre of the footing.)
The principal stress increments shown in Figure 6.1 are superimposed on the initial stresses in the
ground, and the combination determines both the magnitude and direction of 1, 2 and 3. Figure 6.2
shows the directions of 1 predicted to occur at failure around the failure surface beneath a long
embankment constructed on soft clay, in which the initial stresses involved K0 ( ho/ vo) of 0.49.
Significant deviations of the 1 direction from the vertical are apparent. In a soil with pronounced fab-
ric, such as a laminated clay, the response of the clay will obviously vary around the failure surface with
changing direction of the major principal stress. How significant are the variations in response to load-
ing direction in the more homogeneous soils which are the subject of this review, i.e. how significant is

77
09031-02[1].qxd 18/Oct/02 12:10 AM Page 78

90o
0o
80o 4o
50o 12o
24o 18o

Figure 6.2. Orientation of major principal stress along failure surface beneath embankment.


6
Principal stress ratio,

45°
90°
R=σ'1/σ'3

4
σ1 65°
α δ
2

Bedding planes
(a) 0-12%
0
0 2 4 6 8 10 12
Axial strain (%)
250

200
45°
σ1-σ3 (kPa)

65°
150
90°
100
σc'=80kPa
50 OCR=1
e0.05=0.654 - 0.657 (b) 0-0.5%
0
0.0 0.1 0.2 0.3 0.4 0.5
10

8
σ1-σ3 (kPa)

2
(c) 0-0.005%
0
0.000 0.001 0.002 0.003 0.004 0.005
Local axial strain, (%)

Figure 6.3. Plane strain compression tests on air pluviated Toyoura sand (Park,1993).

78
09031-02[1].qxd 18/Oct/02 12:10 AM Page 79

50

φ'max

45 Dr = 81%

φ'max
φ'o

40
Dr = 53%

φ'cv
35

σc'=100 kPa
30
0 15 30 45 60 75 90

Major principal stress direction αº

Figure 6.4. Anisotropy of  in Toyoura sand in plane strain (Park, 1993).

the anisotropy of stiffness and strength in soil? This question is addressed in this Section. The impor-
tance of anisotropy in engineering problems is considered by Hight and Leroueil (2002*).
We are limited in our ability to measure the anisotropy of strength and stiffness in soils, by control-
ling both the direction of the major principal stress, 1,which we shall define in terms of its angle to the
vertical, , and the relative magnitude of the intermediate principal stress, 2, which we shall define in
terms of b (( 2  3)/( 1  3)). We can run triaxial compression tests, in which  equals 0° and b
equals 0, or triaxial extension tests in which  equals 90° and b equals 1. Comparison of data from the
two types of test on the same soil does not allow the effects of changes in  and b to be separated.
Initially data will be presented on the anisotropy of strength and the effect on it of b for reconstituted
soils, obtained from four different pieces of apparatus: the true triaxial apparatus (TTA); the plane strain
compression apparatus with flexible boundaries (PSC), used to test tilted samples, i.e. samples in which
the initial direction of the bedding relative to the fixed direction of 1 can be varied; the Directional
Shear Cell (DSC); and the Hollow Cylinder Apparatus (HCA). The way in which the observed pattern
of anisotropy is likely to be modified by the fabric and structure of natural soils is then discussed, before
introducing the limited data on anisotropy in natural soils. A presentation of data on the anisotropy of
stiffness at small and intermediate strains in reconstituted and natural soils is then introduced.

6.2 Anisotropy in reconstituted soils


6.2.1 Initial, induced and evolving anisotropy in sands
Figure 6.3 presents the results from drained PSC tests on isotropically consolidated tilted samples of
Toyoura sand to illustrate how anisotropy evolves during shear. The data is presented in three separate
plots, covering increasing ranges of axial strain. At small strains (Figure 6.3a), the response is almost
isotropic, with little difference between samples sheared with their bedding planes at different angles to
the major principal stress directions. As strain levels increase (Figure 6.3b), the responses diverge and
anisotropy is apparent. At large strains (Figure 6.3c), there is strong anisotropy in stiffness and strength.
The evolution of anisotropy involves different facets of structure coming into play. At small strains, one
can imagine that particle (or aggregate) contacts play a significant part; at larger strains, changing par-
ticle arrangement and void closure play a part. Thus, anisotropy at small strains is unlikely to reflect
anisotropy at large strains.
Data from PSC tests of this type define the anisotropy of plane strain strength in pluviated sands
(Figure 6.4). The level of anisotropy reduces as relative density reduces and disappears at large strains.
Variations in peak friction angle are very significant in a wide range of dense sands (Figure 6.5). At this
point, it is perhaps appropriate to present a summary of the effect of b on peak friction angle, as measured

79
09031-02[1].qxd 18/Oct/02 12:10 AM Page 80

50
Toyoura
Silver Leighton Buzzard
Silica (Dr=50%)
Monterey
Ticino
Hostun
φ'max 45 Karlsruhe

40

σ3'=80 kPa OCR=1 Dr=80-90%


35
0 3 0 60 90
Major principal stress direction, αo (horizontal bedding)

Figure 6.5. Anisotropy of max in air pluviated sands in plane strain compression (Park & Tatsuoka, 1994).

60

55

50
φ' (º)

45

40

35
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
b
Lade&Duncan (1973) - dense Monterey No.0 sand, Dr = 98%; CP = 60kPa
Lade&Duncan (1973) - loose Monterey No.0 sand, Dr = 27%; CP = 60kPa
Ochiai&Lade (1983)- dense Monterey No.0 sand, Dr = 93%; CP = 100kPa
Symes (1983) - dense HRS, Dr = 85%; p' = 200 kPa
Lam&Tatsuoka (1988) - dense Toyura sand, Dr = 80%, CP = 100 kPa; = 0º
Lam&Tatsuoka (1988) - dense Toyura sand, Dr = 80%, CP = 100 kPa; = 30º
Lam&Tatsuoka (1988) - dense Toyura sand, Dr = 80%, CP = 100 kPa; = 60º
Lam&Tatsuoka (1988) - dense Toyura sand, Dr = 80%, CP = 100 kPa; = 90º

Figure 6.6. The effect of b on peak friction angle (from Zdravkovic, 1996).

80
09031-02[1].qxd 18/Oct/02 12:10 AM Page 81

in TTA and HCA experiments on isotropically consolidated samples (Figure 6.6). Included are a set of
curves for Toyoura sand from Lam and Tatsuoka (1988) in which b was varied in tests with different incli-
nations of the bedding planes to the vertical. The data from Symes (1983) was from HCA tests on dense
Ham river sand with  at 45°. It is apparent from Figure 6.6 that the peak value of  increases when b
increases from 0 to 0.5 and remains constant or decreases slightly at b of 0.9 to 1.0. The effect of b becomes
less at larger values of , i.e. when the bedding plane inclination changes from horizontal to vertical.
The DSC, introduced by Arthur et al. (1977, 1981), is a plane strain device in which the direction
of the major principal stress is controlled during shear by applying both normal stress and shear stress
to four faces of a cube (Figure 6.7). The DSC can be used in two configurations, as shown in Figure 6.8.
In the first (Figure 6.8a), it can be used to shear samples in a plane parallel to the bedding planes.

Rigid Backing Plate


Reinforcing
rubbed pulling
sheets Pressure bag
σa retaining vanes
σa

Acrylic Triangular τa
Prism τb
σb σb σb
Soil Sample
σb
Embedded Shot τa
for Radiography τb

σa Pressure bag σa
Unreinforced
rubber strips

0 50mm

Figure 6.7. Application of normal and shear stresses in the UCL directional shear cell (Arthur et al., 1981).

Direction of deposition Direction of deposition

Elevation

First loading
direction
 ψ Subsequent  
 loading direction

Plan Elevation

(a) To examine induced anisotropy (b) To examine initial anisotropy

Figure 6.8. Directional shear testing.

81
09031-02[1].qxd 18/Oct/02 12:10 AM Page 82

5
ψ = 0o 90o
o
50

Stress ratio, R=σ'1/σ'3


20o
4 60o 70o 80o

First loading cycle


3

(a) Induced
1

5
α = 0o
70o 90o
Stress ratio, R=σ'1/σ'3

o o
40 60
4

3 Isotropic plane

(b) Initial
1
0 2 4 6 8 10 12 14
Major principal strain, ε1 (%)
Figure 6.9. Induced and initial anisotropy in loose Leighton Buzzard sand (Arthur et al., 1981).

When sheared in this plane with different directions of the major principal stress, samples of loose
Leighton Buzzard sand were found to be isotropic. A typical response in this plane is shown in
Figure 6.9a as the curve labelled “First loading cycle”.
The fact that anisotropy can be induced has been demonstrated in these drained DSC tests by first
applying a loading cycle in the horizontal plane with the principal stress direction fixed, then reloading
the specimen with the principal stress at different angles, , relative to the direction of the first loading
(Figure 6.8a). The response was no longer isotropic but varied with the direction of reloading, as indi-
cated in Figure 6.9a. The induced anisotropy resulted in the response being stiffer and stronger than
originally for some directions of loading, but softer and weaker for other directions.
In the second configuration for the DSC, shear and normal stresses can be applied to control the prin-
cipal stress directions in the vertical plane, i.e. the plane containing the bedding planes (see Figure 6.8b),
to examine the initial or inherent anisotropy. The results of such tests on the same Leighton Buzzard
sand are shown in Figure 6.9b. Clearly, the sand is strongly anisotropic in the vertical plane; the fact
that the sand is also isotropic in the horizontal plane is a demonstration that the sedimented sand is
cross-anisotropic.
Comparing Figures 6.9a and b, it can be seen that, for some directions of loading, the response in
the vertical plane is stiffer and stronger than the initial response in the horizontal plane, and softer and
weaker for other directions. In fact, the pattern of initial anisotropy in the vertical plane is similar to that
induced in the horizontal plane by the first loading cycle.

6.2.2 Undrained anisotropy in isotropically consolidated loose sands


In the hollow cylinder apparatus, a sample for which is shown in Figure 6.10, control over the major
principal stress direction, , and of b, is achieved by independently varying the inner and outer cell

82
09031-02[1].qxd 18/Oct/02 12:10 AM Page 83

MT

P1 Po

σz σ1
α

τzθ σ3
σθ = α
σr σ2
Figure 6.10. Idealised stress conditions in the wall of a hollow cylinder sample.

100
α = 0o α = 0o
30o
(σ1-σ3)/2 (kPa)

15o 15o
o
30
50
45o
45o

60o 90o 90o 60o

0
0 50 100 150 200 0 4 8 12

(σ1'+σ2'+σ3')/3 (kPa) Octahedral shear strain (%)

Figure 6.11. Undrained HCA tests on water pluviated Ham river sand. Isotropic consolidation, b  0 (Shibuya,
1985).

pressures, pi and p0, the axial load, W, and the torque, MT. The HCA is the only apparatus in which both
 and b can be independently controlled.
Figure 6.11 compares the behaviour of water pluviated and isotropically consolidated Ham river sand
sheared undrained with b  0 and  held at values between 0° and 90° (Shibuya, 1985). Note, firstly, the
major differences in peak resistance mobilised at intermediate strains. This peak in resistance is reached
well before the friction angle is fully mobilised in each test. Note, also, the difference in stiffness and how
the character of the material changes. At   0°, which corresponds to a triaxial compression test, the
material shows its strongest response, it is ductile and there is a tendency to dilate. The tendency to dilate
at large strains reduces as  increases, and at  of 90° no such tendency appears to exist.

83
09031-02[1].qxd 18/Oct/02 12:10 AM Page 84

150
α =0 o 15
o

(σ1-σ3)/2 (kPa) 15
o
α = 0o
100

o
30 30
o
o
45 45
o

50
o o
60 60
o o
90 90
e=0.89-0.91
0
0 50 100 150 200 0 5 10 15

(σ1'+σ2'+σ3')/3 (kPa) ε1-ε3 (%)

Figure 6.12. Undrained HCA tests on water pluviated Fraser River sand. Isotropic consolidation, b  0
(Uthayakumar & Vaid, 1998).

100
100
(σ1-σ3)/2 (kPa)

α=0° (σ1-σ3)/2 (kPa)


α=15°
α=0°
α=30°
50 α=15°
α=45°
α=60° 50 α=45° M1
α=60° α=30°
α=75° M2
α=90° M3
M4
M5
0 M6
0 50 100 150 200 0 0 5 10 15
(σ1'+σ2'+σ3')/3 (kPa) Octahedral shear strain (%)

Figure 6.13a. Undrained HCA tests on water pluviated Ham river sand. Isotropic consolidation, b  0.5 (Shibuya,
1985).

100 100 M11


M11 M14
(σ1-σ3)/2 (kPa)

M14 M15
(σ1-σ3)/2 (kPa)

M15 M16
M16

50 50
α=30° α=30°
α=45° α=60° α=45°
α=60°
α=90° α=90°
0
0 0 5 10 15
0 50 100 150 200
Octahedral shear strain (%)
(σ1'+σ2'+σ3')/3 (kPa)

Figure 6.13b. Undrained HCA tests on water pluviated Ham river sand. Isotropic consolidation, b  1.0 (Shibuya,
1985).

The results of a similar set of tests in a different HCA on water pluviated Fraser River sand are shown
in Figure 6.12 and the pattern of behaviour is almost identical.
Figures 6.11 and 6.12 illustrate the effect of principal stress direction on the undrained behaviour of
sand for the condition of b  0. The effect of b can be seen in Figure 6.13 which presents undrained
HCA tests on sedimented Ham river sand with b  0.5 and 1.0. Putting the data together from
Figures 6.11 and 6.13, we can examine in Figure 6.14 the variation in strength ratio, tp/p, with  and b,
where tp is the peak resistance at intermediate strains. It is clear that, in these isotropically consolidated
loose sands, the effect of  is more significant than that of b, when considering undrained resistance.
There is a factor of more than 2 in terms of the anisotropy of undrained strength in these materials.
Stacking the effective stress paths observed in a series of HCA tests along an  axis, latched on to
conventional stress path axes, as shown in Figure 6.15, defines a surface which portrays the initial
anisotropy of the material. For normally consolidated material, this is the local bounding surface. The
surface has several uses, one of which is illustrated in Figure 6.16. By tracking the effective stress path

84
09031-02[1].qxd 18/Oct/02 12:10 AM Page 85

K0
0.6
b=0.3
Isotropic
b=0
b=0.5
b=1
0.4
tp / σ'vc

0.2

0.0
0 20 40 60 80

Major principal stress direction, α 0

Figure 6.14. Anisotropy of undrained stress ratio (intermediate strains) in Ham river sand.

100 100
(σ1 - σ3)/2
(kPa)
50 50

0 0
15
50 30
100 45
(σ ’ 60
1 + σ
2’ + σ
150 75 α (deg)
3 ’)/3
(kPa 200 90
)

Figure 6.15. Local bounding surface for isotropically consolidated water pluviated Ham river sand, b  0.

100 100

(σ1 - σ3)/2
(kPa) INE
AKL
50 PE 50

0 0
15
50 30
100 45
(σ ’ 60
1 + σ 150 75 α (deg)
2’ +σ
3 ’)/3 200 90
(kPa
)

Figure 6.16a. Undrained principal stress rotation at constant t leading to failure.

85
09031-02[1].qxd 18/Oct/02 12:10 AM Page 86

100 100

(σ1 - σ3)/2
E
(kPa) K LIN
P EA 50
50

0 0
15
50 30
(σ ’ 100 45
1 +
σ 60
2’ +σ 150 75 α (deg)
3 ’)/3
(kPa 200 90
)

Figure 6.16b. Principal stress rotation leading inside the local bounding surface.

100 100
ion
(σ1 - σ3)/2 reg
nt’
lata
(kPa) ‘Di
se
50 Collap
u r f a c e 50
s

0
0
(σ 50
1’ + 100 15
σ’ 45
30
2 +σ 150 60
3 ’)/
3 75 α (deg)
(kP 200 90
a )

Figure 6.17a. Extent of collapse surface in relation to complete local bounding surface for water pluviated Ham
river sand.

in a problem, it is possible to predict the response of the material to rotation of principal stress direc-
tions. The example shown in Figure 6.16a is for an element which is sheared initially at   0°. At a
certain level of shear stress, the principal stresses are rotated and the shape of the surface dictates that
the material will fail at a certain value of , simply as a result of principal stress rotation. In Figure
6.16b, the initial principal stress direction is at 60°; subsequent rotation of the principal stress direction
to a lower  brings the stress path inside the surface, resulting in a stiff response.

6.2.3 Collapse surface


A second use of such surfaces of initial anisotropy, shown in Figure 6.17a, is that they help to identify
regions where the response is brittle, i.e., where the surface descends after reaching a peak, and where
there is no dilation at large strains and even a tendency to collapse. In water pluviated sands, the sam-
ples tend to dilate at large strains when sheared with low values of . At high values of , the behaviour
is brittle at large strains and the local bounding surface descends from a peak point.
The key features of a collapse surface are summarised in Figure 6.17b, which shows a constant 
section through the surface; these features are:
• an undrained approach to the collapse surface results in a brittle response; this occurs well before the
full angle of shearing resistance, , is mobilised

86
09031-02[1].qxd 18/Oct/02 12:10 AM Page 87

(σ1-σ3)/2
e
lin
re
ilu
Fa

ce
surfa
l lapse
Co

(σ’1+σ’3)/2

Undrained
Drained Strain controlled
Drained
Load controlled -drained -undrained

Figure 6.17b. Constant  section through collapse surface showing conditions leading to collapse.

• an approach under load maintained conditions results in collapse and runaway failure whether the
approach is undrained or drained, unless pore pressures can dissipate immediately during the col-
lapse process
• drained strain-controlled approaches to the collapse surface allow the effective stress path to cross
the surface, although this is marked by the onset of large strains.
The surface in a constant  section was referred to as the liquefaction line by Shibuya and Hight
(1991); others have referred to it as the collapse surface (refer to Section 5.3, Figures 5.9 and 5.10). The
key thing to note is that, in water pluviated sands, the collapse surface only exists for a limited range of
 values. Its location can be modified by induced anisotropy.

6.2.4 Undrained anisotropy in K0 consolidated loose sands


Further evidence of induced anisotropy is available from a comparison of the results of hollow cylinder
tests run on water pluviated Ham river sand which has been consolidated under K0 conditions, with 1
vertical (Figure 6.18), with the data presented earlier on the isotropically consolidated material. A direct
comparison can be made in Figure 6.19 in which the solid lines are the effective stress paths observed
in K0-consolidated samples sheared with different values of , in most cases after undrained unloading
from the consolidation stress. The broken lines are the effective stress paths from the tests on isotro-
pically consolidated samples. It can be seen that K0 consolidation distorts the anisotropy locally to the
direction of 1 during consolidation. Effective stress paths at high values of  are coincident. Undrained
strength ratios at intermediate strains after K0 consolidation have been added to Figure 6.14; this
emphasises how the increase in strength induced by K0-consolidation magnifies the anisotropy. Now
the maximum difference in strength is a factor of 3.

6.2.5 Undrained anisotropy in K0 consolidated reconstituted silt


Zdravkovic and Jardine (2000) describe the anisotropy of undrained shear strength and yield of a dense
K0 consolidated quartzitic silt (HPF4), formed by pluviation and investigated using an HCA. The silt is
an industrially produced rock flour with highly angular and elongated particles. The results of the tests
on the normally consolidated silt, carried out with b  0.5, are presented in Figure 6.20. Note that to
explore the shape of the three-dimensional bounding or limit state surface it was necessary to partially
unload some of the samples before shearing with the principal stress direction fixed. The initial
response of the silt was stiff and highly non-linear. Yield at the Y3 surface was reached with little change
in p, after which there was a stable contractant stage until the phase transformation point was reached,

87
09031-02[1].qxd 18/Oct/02 12:10 AM Page 88

150
α=0o
α=0o
(σ1-σ3)/2 (kPa) o
15
100 o
15 o
30
tion
so lida
con
Ko
o
30
50
45o 45o
o o
90 90

0
0 50 100 150 200 0 4 8 12
(σ1'+σ2'+σ3')/3 (kPa) ε1-ε3 (%)

Figure 6.18. Undrained HCA tests on water pluviated Ham river sand. K0 consolidation, b  0.3 (Shibuya, 1985).

Ko consolidation (b=0.3) α=0o


Isotropic consolidation (b=0.5)
100

o
(σ1-σ3)/2 (kPa)

15

o
0
o
50 15
o
45 45o
o
90 90
o

0
0 50 100 150 200

(σ1'+σ2'+σ3')/3 (kPa)

Figure 6.19. Comparison of anisotropy in isotropically and K0 consolidated Ham river sand.

150
M1
M2 α=0o
M3
M4
M5
M6 α=15o
100
(σ1-σ3)/2 (kPa)

α=30o

α=45o

50 α=70o
α=90o

0
0 50 100 150 200
(σ1'+σ2'+σ3')/3 (kPa)

Figure 6.20. Undrained HCA tests on pluviated HPF4 silt. K0 consolidation, b  0.5 (Zdravkovic & Jardine, 2000).

referred to as Y4 yield by Zdravkovic & Jardine (2000). After phase transformation all samples showed
a strong tendency to dilate.
Figure 6.21 summarises the variations with the major principal stress direction, , of the friction
angle, , and the normalised shear stress, t, at phase transformation. Also shown are the corresponding
values from triaxial compression and extension tests. There is clearly a major effect of b on  at phase

88
09031-02[1].qxd 18/Oct/02 12:10 AM Page 89

60

55 b = 0.5; shearing after K0 consolidation


PT values of φ'

b = 0.5; shearing after inclined consolidation


50 PT values of φ'

estimated plain strain envelope


45 b = 0.3
φ' (º)

40

35

TXC PT = Phase Transformation


30

25 TXE

20
0 15 30 45 60 75 90

α (º)

Figure 6.21a. Anisotropy of  mobilized at phase transformation in K0 and c consolidated HPF4 silt.

also b = 0.3

0.65
b = 0.5; inclined consolidation to p0' = 400 kPa
0.6 TXC
0.55

0.5

0.45

0.4
tpt / pc'

0.35

0.3

0.25

0.2
b = 0.5; K0 consolidation to p0' = 200 kPa
0.15 TXE
0.1

0.05

0
0 15 30 45 60 75 90

α (º)

Figure 6.21b. Anisotropy of normalized shear stress at phase transformation in K0 and c consolidated HPF4 silt.

89
09031-02[1].qxd 18/Oct/02 12:10 AM Page 90

transformation, being 52–53° at b  0.5, 42° at b  0.3 and 32° at b  0 for   0°. There is also sig-
nificant anisotropy in  and t/p0 at b  0.5.

6.2.6 Undrained anisotropy in K0 consolidated reconstituted Boston Blue clay


The undrained anisotropy of reconstituted Boston Blue clay has been investigated in the DSC by
O’Neill (1985) and Seah (1990), testing clays with OCRs of 4 and 1, respectively. The results of
O’Neill’s experiments are presented in Figure 6.22. These confirm that the sedimented clay is cross-
anisotropic and that the pattern of anisotropy in the vertical plane is similar to that seen in the pluviated
sands. Again, strengths and stiffnesses are higher than in the isotropic plane for low values of , and
lower for high values of . The source of anisotropy is differences in dilational tendencies in this over-
consolidated clay, rather than any differences in .

Figure 6.22. Constant volume DSC tests on resedimented Boston Blue Clay. K0 consolidation, OCR  4 (O’Neill,
1985).

Ham river Sand


HK clayey sand
0.4 KSS clay-silt-sand mix
Boston Blue Clay (plane strain)
HPF4 silt
0.3
tp / σ'vc

0.2

0.1

0.0
0 20 40 60 80

Major principal stress direction, α0

Figure 6.23a. Anisotropy of undrained peak strength ratio in K0 normally consolidated soils (Shibuya, 1985;
Menkiti, 1995; Zdravkovic, 1996; O’Neill, 1985).

90
09031-02[1].qxd 18/Oct/02 12:10 AM Page 91

6.2.7 Typical levels of anisotropy in reconstituted soils


Figure 6.23a summarises the data that has been obtained at Imperial College and MIT on anisotropy
of undrained strength in K0-normally consolidated reconstituted sands, clays and silt. First of all, note
the similar pattern of anisotropy in each material, but how the variation is least in the artificial clay KSS
and greatest in the water-pluviated granular materials. Undrained strengths can vary with  by a factor
of up to 3 in these reconstituted soils.
Figure 6.23b compares the anisotropy in peak  for dense Ham river sand (HRS), dense silt (HPF4),
a clayey sand (HK), and the artificial clay (KSS). The clay is almost isotropic in terms of  but there
is strong anisotropy in the granular materials, the level of which depends on their density.
A more complete description of the variation in normalised undrained strength, tp/ vc, in Ham river
sand is presented in Figure 6.24, using axes of b and ; a similar presentation is used in Figure 6.25 for
the variation in  in dense Toyoura sand.

HK (b=0.5)
KSS (b=0.5)
50 HRS (b=0.3)
HPF4 (b=0.5)
φ'oult

40

30

0 20 40 60 80
Major principal stress direction
Figure 6.23b. Anisotropy of ult in K0 normally consolidated soils (Shibuya, 1985; Menkiti, 1995; Zdravkovic,
1996; O’Neill, 1985).

t /σ’
vc
0.5 0.5

Triaxial
Compression
0.25
0.25
Triaxial
Extension
0
0 1

30 0.5
60 b
σ (deg)
90 0

Figure 6.24. Variation of tp/ vc with b and  in Ham river sand, isotropic consolidation.

91
09031-02[1].qxd 18/Oct/02 12:10 AM Page 92

6.2.8 Anisotropy induced by ac consolidation


In an elegant series of tests on the HPF4 silt, Zdravkovic (1996) has generalised the findings regarding
the anisotropy induced by non-isotropic consolidation. She has shown that consolidation along constant
ratio stress paths, but with different directions of the major principal stress (paths R0, R15, R30, R45,
R70 and R90 in Figure 6.26), locally modifies the anisotropy of volume change behaviour, and, there-
fore, of undrained response. Jardine et al. (1997) introduce the term c to distinguish consolidation
paths with different principal stress directions during consolidation. In Figure 6.21b, the anisotropy of
undrained strength at phase transformation, tpt, normalised by the mean effective stress during consoli-
dation, pc, has been shown for the case of K0 consolidation to p0 of 200 kPa, i.e. with c equal to 0.
These strengths may be compared with those measured at the same  values, but after consolidation to
pc of 400 kPa with c equal to the value of  during subsequent undrained shearing. In each case, sim-
ilar levels of strength are induced in the c direction, but at the expense of increased brittleness.
In the dense silt, the induced anisotropy is restricted to the behaviour at intermediate strains, i.e. in
behaviour up to phase transformation. The values of  mobilised at large strains are relatively unaf-
fected by the value of c, (Figure 6.21a) but remain a function of the initial fabric and anisotropy. This
confirms that, in this dense granular material, fabric changes are relatively subtle and likely only to
affect response at small and intermediate strains, and then only local to the direction of 1 during con-
solidation.

50
φ’ 50

45
45
Triaxial
Compression 40
40 Triaxial
Extension
35
35 1
0
30 0.5
60 b
α (deg)
90 0

Figure 6.25. Variation in peak friction angle with b and  in Toyoura sand, isotropic consolidation.

R0 R15 R30
150 R45
R70 R90
125
)

100
t (kPa

75
50
25 400
0 300
20 200
Major 40 100
princip 60
80 Pa)
a l stres
s direc
0
p' (k
tion, ( o
)
Figure 6.26. Consolidation stress paths for c consolidation, R-series tests (Zdravkovic, 1996).

92
09031-02[1].qxd 18/Oct/02 12:10 AM Page 93

6.3 Anisotropy in natural soils


6.3.1 Modifications to the pattern of anisotropy due to fabric and structure
While the reconstituted materials that have been considered to date have their own particular fabric,
related to the way in which they were reconstituted, they are not necessarily representative of the fabrics
of natural sands and clays. Since anisotropy of stress-strain-strength behaviour is closely linked to fab-
ric, it is necessary to consider the structure – that is the fabric, lithification and bonding – of natural
materials – and see how that affects the pattern of anisotropy observed in reconstituted soils. Hight
(1998) speculated on the potential effects of depositional and post-depositional processes on the
anisotropy of undrained strength in natural soil; taking the solid lines in Figure 6.27 as representing the
anisotropy of the reconstituted soil, he suggested that they might be modified to the dotted and broken
lines. He presented evidence of increased anisotropy in soft clays as a result of ageing and diagenesis,
which would distort the picture as shown in Figure 6.27a. In soft laminated or varved clays, it was sug-
gested that more extreme levels of anisotropy were likely, in both drained and undrained strength, with
minimum values occurring with  between 45° and 70°, as shown in Figure 6.27b. It was further sug-
gested that burial compaction of clay-rich clays could lead to similar levels of fabric anisotropy and thus
to similar levels of strength anisotropy. In silt-rich clays this particular anisotropic phenomenon was

0.4
Reconstituted
Contact ageing
Ageing Isotropic cementing
0.3

Reco
Cup / σ'vc

nstit
uted
0.2

0.1

0.0
0.4

0.3
Ageing
Cup / σ'vc

Rec
0.2 onst
itute
d
Bedding
compaction
0.1 lamination

0.0
0 20 40 60 80
Major principal stress direction, α0
Figure 6.27. Potential effects on anisotropy of undrained peak strength ratio of ageing, bonding, bedding and
compaction.

93
09031-02[1].qxd 18/Oct/02 12:10 AM Page 94

150

Uniaxial compressive
strength (MPa)
100

50

0
0 30 60 90
Inclination of bedding to vertical (degrees)
(a) Anisotropy of uniaxial compressive strength in a slate
(after Hoek, 1965)
10
compressive strength (MPa)

8
Undrained triaxial

4
σ1 σ1 σ1 σ1 σ1
2

0
0 45 90 135 180
Inclination of bedding to vertical (degrees)
(b) Anisotropy of undrained triaxial compressive strength of
Sagamihara mudstone (after Tatsuoka et al, 1997)
Figure 6.28. Anisotropy of strength in mudrocks.

less likely to occur. The presence of low strength shear surfaces or discontinuities would also modify
anisotropy. The cross-anisotropic properties of reconstituted sedimented soils would be modified by
inclined bedding or by non-isotropic horizontal stresses, which would induce azimuthal anisotropy. As
shown in Figure 6.27a, the possibility of isotropic cementing was noted.
This speculation was aided by the existence of data from naturally bonded or cohesive materials in
which anisotropy could be examined in tests on samples cut at different inclinations and loaded in com-
pression. Figure 6.28 shows the results of two such investigations in mudrocks. Figure 6.28a presents
results showing the anisotropy of uniaxial compressive strength in a slate, which, of course, presents an
extreme example of fabric anisotropy and resulting strength anisotropy. The example in Figure 6.28b
is of another mudrock, the Sagamihara mudstone, which shows only minor anisotropy of undrained tri-
axial compressive strength; this mudrock has not experienced the same pressures as the slate and,

94
09031-02[1].qxd 18/Oct/02 12:10 AM Page 95

1.5

1
Cu0 / C uv

0.5

0
0 30 60 90

α (°)

London Clay (Ward et al, 1965)


Welland Clay (Lo, 1965)

Figure 6.29. Anisotropy of strength in two natural clays.

because of its significant silt content, has not developed a strongly anisotropic fabric. Two examples for
natural clays are presented in Figure 6.29; data for the heavily overconsolidated London Clay indicates
undrained strengths that are higher for horizontal loading, while data for the lightly overconsolidated
Welland Clay shows undrained strengths higher for vertical loading.

6.3.2 Investigation of anisotropy of a natural clay in an HCA


Investigations of the complete anisotropy of stiffness and strength of a natural soft clay are rare. One
reported by Leroueil et al. (2002*) on the Louiseville clay is cited here. A series of tests in a hollow
cylinder apparatus of the type used by Saada & Ou (1973) were performed (Leblond, 2002). The spec-
imens tested were taken from a depth of about 10 m. The specimens were carefully trimmed to an exter-
nal diameter of 70 mm, an internal diameter of 50 mm and a height of 160 mm. The specimens were
mounted in the HCA and consolidated under an isotropic stress equal to 0.3 p. They were finally
sheared in undrained conditions with the axis of the major principal stress at an angle  to the vertical

95
09031-02[1].qxd 18/Oct/02 12:10 AM Page 96

70
σ'1
β
60

50

40

Mohr - Coulomb
30 
criterion (φ' = 2 8˚
θ,˚

20

10

0
10 20 30 40 50 60 70 80
-10 β ,˚
-20

Figure 6.30. Inclination of failure surface with horizontal, , versus the angle of stress axis rotation (after Leblond,
2002) (  ).

that varied from 0° (conventional compression) to 75°. Other tests were performed at  angles of 0°, 22°
and 29° after isotropic consolidation to 0.19, 0.44 and 0.54 p. As the inner and outer cell pressures
were equal in the HCA used,  was linked to the intermediate principal stress 2 by the equation:
( 2  3)/( 1  3)  sin2.
Well-defined failure surfaces were observed in all tests, and the angle of these surfaces to the hori-
zontal, , was measured. The  values are plotted as a function of  in Figure 6.30. As can be seen, the
shear surface was at an angle  of 56° for   0, decreasing to 22° at   29°, and then abruptly
dropped to 0° for larger  values. According to the Mohr-Coulomb criterion,  is related to  as follows:

′
  45°   (4)
2

This criterion has been shown on Figure 6.30 for a friction angle of 28°. While the soil behaviour
approximately follows Equation 4 for  values smaller than 30°, there is a significant variance from the
Mohr-Coulomb criterion for larger  values.
The tests performed at   0, 22 and 29° after consolidation at different effective stresses gave the
strength envelopes shown in Figure 6.31. For the tests performed at   30°, stress conditions at fail-
ure on the horizontal failure plane were calculated and are plotted on Figure 6.31. It can be seen that this
gives a strength envelope that is below the strength envelopes deduced from the   0–29° tests. This
demonstrates that Louiseville clay has lower strength characteristics on the horizontal plane than on
inclined planes. It has been observed, using scanning electron microscopy, that the relief is more pro-
nounced on the vertical plane than on the horizontal plane (Lapierre et al., 1990). The difference
between the strength envelopes for horizontal and inclined planes may be due to this difference in
roughness.
The Corinth Marl described by Kavvadas et al. (2002*) is reported to be isotropic because of the high
strength of isotropic bonds, the presence of non-platey silt size particles, and because of the early devel-
opment of bonding, which has maintained an open structure and inhibited the development of stress
induced anisotropy.

96
09031-02[1].qxd 18/Oct/02 12:10 AM Page 97

0.5
29 Strength envelopes for
22
= 0, 22 and 29°
0 (Failure plane horizontal)
0.4

Strength envelope
0.3 37 for α = 37 - 75 °
(Failure plane horizontal)
τ/σp'

45
53 40 Stress conditions at failure
65 50 on horizontal plane
0.2
68 60
75

0.1

0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
σ' /σp'

Figure 6.31. Strength envelopes deduced from undrained HCA tests (after Leblond, 2002) (  ).

Stress system-induced
anisotropy
3
Ev/Eh

Inherent anisotropy

1
Toyoura sand
SLB sand
Perfectly isotropic material Ticino sand
Hime gravel
0 Chiba gravel
0 1 2 Nerima gravel
σv'/σh'

Figure 6.32. Small strain anisotropy in sands and gravels (Tatsuoka et al., 1997).

6.4 Anisotropy at small and intermediate strains


The anisotropy of Toyoura sand at small strains was shown in Figure 6.3a. Although it was seen to be
isotropic at small strains, this is not the case in general. Figure 6.32 shows that initial or inherent
anisotropy, measured under isotropic stress ( v/ h  1) in terms of the ratio Ev/Eh, varies with grain size
in granular materials, and can be as large as 2.5. This level of anisotropy is modified as the consolidation

97
09031-02[1].qxd 18/Oct/02 12:10 AM Page 98

600
Ghh
Gvh
E'h
500 E'v

400
Stiffness (MPa)

300

200

100

0
0 50 100 150 200 250
p' (kPa)
Figure 6.33. Variation of stiffness components with p in dense Ham river sand during K0 consolidation, OCR  1
(Jardine et al., 2001).

stress ratio v / h changes; this stress-system induced anisotropy increases the ratio Ev/Eh to 3.5 in a
gravel under v/ h equal to 2.
Cross-anisotropic elastic parameters that characterise soils inside the Y1 zone have been derived by
a combination of effective stress triaxial probing tests and shear wave velocity measurements, made on
the same specimen, under the same effective stress conditions (e.g. Jardine et al., 2001). The assumption
is made that the soil is linear elastic and rate independent inside the Y1 zone. One such set of parame-
ters for normally consolidated Ham river sand is presented in Figure 6.33 (Jardine et al., 2001). This
shows that during K0 consolidation:
• Ev/Eh varies between 1.8 and 2.1
• Gvh exceeds Ghh by 20 to 25%
Information on the anisotropy of small strain stiffness of dry Ticino sand, measured at different
consolidation stress ratios in a calibration chamber using wave velocities, has been taken from Bellotti
et al. (1996) and is presented in Figure 6.34.
A compilation of data on the anisotropy of small strain stiffness in natural and reconstituted soft and
stiff clays is presented in Figure 6.35 in terms of the ratio Ghv/Ghh or Gvh/Ghh versus consolidation stress
ratio v / h. In the mottled facies of the Bothkennar Clay, the fabric of which has been rendered
isotropic by bioturbation, the ratio is unity in an unconfined state. In the bedded facies the ratio is 0.8
in an unconfined state, consistent with its anisotropic fabric. Jamiolkowski et al. (1994) found Gvh/Ghh
ratios under isotropic stress conditions of 0.7 and about 0.65 for the natural Pisa & Panigalia clays
respectively. In situ data for the stiff and very stiff London and Gault Clays is very similar, although
under isotropic stress conditions the Gault shows a lower ratio for Gvh/Ghh. Nash et al. (1999) found
ratios of 0.7–1.9 over the range of void ratio tested for the one-dimensionally reconstituted Gault clay
under isotropic stress.
One of the most comprehensive data sets on stiffness anisotropy at intermediate strains has been pres-
ented by Zdravkovic & Jardine (2000), from whom Figures 6.36, 6.37 and 6.38 have been taken.

98
09031-02[1].qxd 18/Oct/02 12:10 AM Page 99

2
DR = 40 - 45 %
σ'h = 100 kPa
1.75

1.5
Modulus ratio

1.25
Gvh / Ev
Ghh / Eh
Ghh / Gvh
1 Eh / Ev
G/E isotropic medium
1.2 K0.25
1.2 K0.25
0.75

0.5

0.25
0.25 0.5 0.75 1 1.25 1.5 1.75 2 2.25 2.5 2.75 3
Consolidation stress ratio K = σ'h / σ'v

Figure 6.34. Small strain stiffness anisotropy in Ticino sand (Bellotti et al., 1986).

1.0
Gvh/Ghh or Ghv/Ghh

Bothkennar in situ (Gvh)


Bothkennar intact (Gvh)
Bothkennar intact (Ghv)
London Clay in situ (Gvh)
London Clay intact (Gvh)
0.5 London Clay reconstituted (Ghv)
Gault Clay in situ (Gvh)
Gault Clay intact (Ghv)
Gault Clay reconstituted (Ghv)
Pisa Clay intact (Gvh)
Panigaglia Clay intact (Gvh)
0.0
0 1 2 3
σv'/σh'

Figure 6.35. Small strain stiffness anisotropy in natural and reconstituted clays.

99
09031-02[1].qxd 18/Oct/02 12:10 AM Page 100

140,000
shearing stages
M0

t (kPa)
120,000 M0
M15
M30 M15
M45 M30
100,000 M70
M45
M90 A
TE M70

80,000
Goct (kPa)

M90

60,000
p' (kPa)

40,000

20,000

0
0.001 0.01 0.1 1 10
εd (%)

Figure 6.36. Torsional shear stiffness in HPF4 silt (Zdravkovic & Jardine, 2000).

400000
HCA compression, b=0.3
HCA compression, b=0.5
300000 HCA extension, b=0.5
Triaxial compression
Triaxial extension
Eu (kPa)

200000

100000

0
0.001 0.01 0.1 1 10
εs (%)

Figure 6.37. Undrained Young’s modulus for different combinations of b and  in HPF4 silt (Zdravkovic, 1996).

Figure 6.36 shows the decay curves for octahedral shear stiffness during undrained shear of the HPF4
silt with different principal stress directions and b  0.5. Stiffnesses for M0 to M70 are, in fact, similar
and much softer than for initial unloading. The effect of b on stiffness characteristics is illustrated in
Figure 6.37, in which octahedral shear stiffness is compared for compression with b  0, 0.3 and 0.5
and for extension with b  1 and 0.5. Figure 6.38 compares the stiffnesses determined from uniaxial
and torsional shear stages of drained tests on the silt at an OCR of 1.3.

100
09031-02[1].qxd 18/Oct/02 12:10 AM Page 101

200000
εd -E h'
εd-Ev'
150000
εd-Gz
Gz , Ghh, E' (kPa)

εd-Ghh

100000

50000

0
0.001 0.01 0.1 1 10

εd (%)

Figure 6.38. Uniaxial and torsional shear stiffness of K0 consolidated HPF4 silt at OCR  1.3 (Zdravkovic, 1996).

6.5 Concluding remarks


Anisotropy of undrained strength can result from the anisotropy of volume change characteristics, as in
loose sands, from anisotropy of stiffness, as in stiff clays, from anisotropy of effective stress strength
parameters, as in dense sands or in stiff fissured or jointed clays, in which the discontinuities have a
dominant orientation.
The anisotropy of natural soils in situ is determined by: the inherent anisotropy of the soil, which can
be considered to be that resulting from the structure of the material, i.e. its fabric and bonding at all lev-
els, and can be measured with the soil in an unconfined state; and the in situ stress state, which is likely
to be anisotropic and induces additional anisotropy.
As the fabric features that can influence anisotropy occur at different scales, so will anisotropy be
scale dependent. The level of anisotropy and its pattern will vary. Anisotropy at small strains bears no
relation to anisotropy at large strains. Because anisotropy is induced by strains, sampling can modify
inherent anisotropy (Hight, 2001).
Examples of the influence of small strain and large strain anisotropy are presented by Hight and
Leroueil (2002*). These confirm that efforts should be made to assess the likely level and pattern of
anisotropy and its subsequent influence on performance.
To take into account the effects of anisotropy, it was suggested by Hight & Higgins (1994) that the
scheme for yielding introduced by Jardine (1992) be extended as shown in Figure 6.39. As indicated,
the curves Y1, Y2 and Y3 decrease in size as the principal stress axis rotates from the vertical direction
to the horizontal one in K0 consolidated soil.

7 BEHAVIOUR OF INTERMEDIATE SOILS

Section 5 has considered the behaviour of clean sands. In nature, sands often contain a component of
fines which may have been included at the time of deposition or may have become incorporated subse-
quently. The fines may have a profound effect on behaviour, depending on the quantity, shape, plastic-
ity and distribution of the fines, as well as on the grading of the host sand. These effects are illustrated
in this Section and are used to illustrate the transition from “sand behaviour” to “clay behaviour”. The
section also reviews the behaviour of silts and an extreme form of fines in sand, namely the presence of
mica in micaceous sands.

7.1 The effect of clay fines on permeability and compressibility


Perhaps the most important effect of the addition of clay fines to a sand is on its permeability and chang-
ing its relevant behaviour from drained to partially drained or even to undrained. Figure 7.1 illustrates

101
09031-02[1].qxd 18/Oct/02 12:10 AM Page 102

t/p'e Y3

0 p'/p'e

αo
Y2

90

t/p'e

ace
surf
ng
ndi
Bo u
Y3

Y2
Y1
Y3

p'/p'e
Projection on t/p’e – p’/p’e plane

Figure 6.39. Framework for soil behaviour extended to cater for principal stress directions.

102
09031-02[1].qxd 18/Oct/02 12:10 AM Page 103

Figure 7.1. Effect of clay fines on the permeability of sand (Hight et al., 1994).

how the addition of less than 5% of clay can reduce permeability by more than 3 orders of magnitude.
It is apparent that the effect depends on the activity of the clay. The effect of these changes in perme-
ability on in situ behaviour is discussed by Hight & Leroueil (2002*).
The effect of clay fines on compressibility is, by contrast, generally limited. Hight et al. (1994) report
that the compressibility of Ham river sand/kaolin mixtures was dominated by the compressibility of the
sand for clay contents up to 20%. With a clay fraction of 20% compressibility was only increased by 1.3
to 1.4 times that of loose sand, although the compressibility of normally consolidated kaolin was 20
times that of loose sand over a stress range of 300–400 kPa.

7.2 The effect of clay fines on the behaviour of sand in shear


The effect of clay fines on the behaviour of sand was investigated by Georgiannou et al. (1990, 1991a,
and 1991b) and summarised by Hight et al. (1994) in connection with the foundation soils at the site of
the Gullfaks ‘C’ gravity platform in the Norwegian sector of the North Sea.
Both the natural clayey sands and a model clayey sand, comprising Ham river sand and kaolin, were
investigated. In the intact clayey sands from Gullfaks, the clay acted as both a separator at the contact
between sand grains and as a filler to the voids between sand grains, see Figure 7.2. This fabric sug-
gested contemporaneous deposition of the sand and clay, with the clay initially in suspension, rather
than being washed in subsequently to fill only the voids.
Contemporaneous deposition of sand and clay was modelled by raining sand slowly into suspensions
of the clay component in distilled water. This revealed a strong dependence of the settled bed height on
the clay content of the suspension, for a given volume of rained sand. Increasing bed height, which cor-
responds to a looser granular component, accompanied increases in clay concentration. In a parallel
study, the ability of the clay to separate (or loosen) the sand was shown to increase with the activity of
the clay and the angularity of the sand grains, for a given clay concentration.
The main features of the behaviour in undrained triaxial compression and extension of K0-normally
consolidated clayey sands, formed in this way, are summarised below, referring to Figures 7.3 to 7.6:
• The addition of relatively small quantities of clay introduces pronounced undrained brittleness in tri-
axial compression; this is illustrated in Figure 7.3 for samples having approximately the same gran-
ular void ratio, eg

103
09031-02[1].qxd 18/Oct/02 12:10 AM Page 104

Figure 7.2. Fabric of clayey sands from Gullfaks “C” site, North Sea.

104
09031-02[1].qxd 18/Oct/02 12:10 AM Page 105

• Strains to phase transformation also increase with clay content (Figure 7.3)
• In undrained triaxial extension, the increased tendency for contraction with the addition of clay leads
to complete liquefaction of the sedimented model clayey sand having clay contents of 3.5 and 7.5%
• At similar clay contents, the effects of reductions in eg after sedimentation are to reduce undrained
brittleness and to reduce the risk of liquefaction in triaxial extension (Figure 7.4)

Figure 7.3. Undrained triaxial compression and extension tests on K0 normally consolidated loose sand clayey
sand (sedimented HRS and kaolin) (Hight et al., 1994).

Figure 7.4. Undrained triaxial compression and extension tests on K0 normally consolidated clayey sands with
varying granular void ratio at similar clay content (Hight et al., 1994).

105
09031-02[1].qxd 18/Oct/02 12:10 AM Page 106

Clay content (%)


200 0
7 30
20
7.6 6
( σa - σ r)/2 (kPa)
100 8
10
Ham river sand pluviated into kaolin suspensions
except 20% and 30% (wet slurry)
0
100 200 300 400 500
(σa’ +σ r’)/2 (kPa)

-100

Figure 7.5. Effective stress paths for undrained triaxial compression on K0 normally consolidated clayey sands
with varying clay content (Hight et al., 1994).

100

80 Triaxial extension
b=1, α=90
o
Bittleness index, IB (%)

60
Triaxial compression
b=0, α=0o

40

20

prepared by wet slurry method


0
0 5 10 15 20 25 30
Clay content (%)

Figure 7.6. Undrained brittleness in sand sedimented into kaolin suspension.

Ham river sand pluviated into 7-7.8% kaolin suspension


eg = 0.78 - 0.80
200 200

OCR = 1.0
OCR = 1.0 OCR = 1.0
(σa - σr)/2 (kPa)

(σ a - σ r)/2 (kPa)

100 1.3 100

2.0 1.3 1.3

2.0 2.0
4.0
0 0
100 200 300 400 500 100 200 300 400 500
(σa’ +σ r’)/2 (kPa) (σa’ +σ r’)/2 (kPa)

-100 -100

Figure 7.7. Effects of stress history in clayey sand Figure 7.8. Effects of stress history in pluviated Ham
(Georgiannou et al., 1990). river sand (Ovando-Shelley, 1986).

106
09031-02[1].qxd 18/Oct/02 12:10 AM Page 107

• The maximum undrained brittleness in triaxial compression occurs at a clay content of 10%
(Figures 7.5 and 7.6); for higher clay contents, undrained brittleness reduces and, at a clay content of
30%, the material assumes the behaviour of normally consolidated clay.
The effects of stress history on the behaviour of the model clayey sand are illustrated in Figure 7.7.
Overconsolidation reduces undrained brittleness in triaxial compression but does not prevent liquefac-
tion in undrained triaxial extension. The effects of stress history on loose sedimented clean sand, involv-
ing K0 consolidation and swelling, are included as Figure 7.8 to demonstrate an essential difference
between clean and clayey sands. Overconsolidated clean sands show gross yielding before reaching the
normalised effective stress paths of the normally consolidated sands, i.e. the Y3 yield surface shrinks on
unloading. In clayey sands, the presence of only small quantities of clay fixes the structure giving the
soil a memory of its stress history.
A problem related to the characterisation of clayey sands discussed by Hight & Leroueil (2002*)
concerns their particular range of permeability, which can lead to a switch between undrained and
drained behaviour. An illustration of the effects of the switch is presented in Figure 7.9 which compares
the drained and undrained behaviour of the model clayey sand at OCRs of 1 and 1.3.
[o - T] /2 [kPa]

200

1,0

DCR = 1,3

0
0 200 400 500
['o - 'r] /2 [kPa]

400

Drained
300
[o - r] [kPa]

200
Undrained

100
0 0.5 1.0
Axial strain, a (%)

Figure 7.9. Comparison of drained and undrained behaviour of clayey sand in triaxial compression (Georgiannou
et al., 1990).

107
09031-02[1].qxd 18/Oct/02 12:10 AM Page 108

7.3 The effect of mica on the behaviour of sand


The behaviour of micaceous sands was studied in connection with flow slides which occurred during
construction of river training works for the Jamuna Bridge in Bangladesh (Hight et al., 1999). Details
are given by Hight & Leroueil (2002*), who use it as a case history to illustrate the importance of
anisotropy and collapse potential in sands.
Two batches (A and B) of the natural micaceous sand, having the gradings shown in Figure 7.10,
were tested, together with mixtures of a model sand and an industrial mica (MF60), which sat within the
grading envelopes shown in Figure 7.10. The mica in the natural sand comprised thin sand-sized plates
of biotite, derived from the crystalline rocks of the Himalayas; the quantity of mica, its distribution and
its orientation varied. Grain counting indicated mica contents of 5–10% by number of grains. SPTs at
the site had indicated relative densities between 40 and 60%.

7.3.1 Effect of mica content on void ratio


It had been demonstrated in the classic study by Gilboy (1928) that the addition of sand sized mica to a
dumped sand increased its porosity dramatically. We repeated Gilboy’s experiments, using the host sand
and MF 60 mica. The results are shown in Figure 7.11, in terms of the void ratio measured in the dumped

Figure 7.10. Particle size distributions of micaceous sands (Hight et al., 1999).

Sand sized mica


Angular silt
2
Kaolin
Void ratio

0
0 5 10 15 20
Percentage of fines

Figure 7.11. Effects of fines on the void ratio of a sand (Hight et al., 1999).

108
09031-02[1].qxd 18/Oct/02 12:10 AM Page 109

sand as the mica content by weight was increased from 0% to 20%. For comparison, Figure 7.11 also
shows the void ratio in the same sand when increasing percentages of angular silt were added, and in a
similar sand, Ham river sand, when sedimented into kaolin suspensions of increasing clay concentra-
tions. In contrast to the effects of adding silt or kaolin to the sand, the effect of adding mica is to increase
the void ratio monotonically with mica content. When dumped and under low stress, the micaceous
sand can be expected, therefore, to be in a loose state.

7.3.2 Effect of mica on volume change characteristics and collapse potential


Samples of the model sand-mica mix, with different mica contents by weight, have been tested in
constant volume simple shear (CVDSS), using the Geonor DSS device (Bjerrum & Landva, 1966). The
samples were formed by dumping the mix through a trap door into the reinforced membrane.
Figure 7.12a compares the behaviour of a clean sand and the same sand with just 1% of mica added. The
clean sand, although loose, is ductile, a tendency to dilate at large strains causing the effective stress

100
Horizontal shear stress (kPa)

80 0 % mica
e = 0.7 0 % mica

60
(a)
40

1 % mica
20 e = 0.74
1 % mica

0
0 20 40 60 80 100 120 0 4 8 12 16 20
Vertical effective stress (kPa) Shear strain (%)

80
Horizontal shear stress (kPa)

0
% mica 0
60

40
(b) 20
20
40 10
10 40
20 5 5
1
2.5
1 2.5
0
0 40 80 120 0 5 10 15 20
Vertical effective stress (kPa) Shear strain (%)

100
Horizontal shear stress (kPa)

80

% mica
0 (e = 0.76) % mica
60 0 (e = 0.76)

(c)
40
5 (e = 0.76) 5 (e = 0.76)

20
5 (e = 0.91) 5 (e = 0.91)

0
0 20 40 60 80 100 120 0 10 20
Vertical effective stress (kPa) Shear strain (%)

Figure 7.12. Effects of mica on the behaviour of sand in constant volume simple shear (Hight et al., 1999).

109
09031-02[1].qxd 18/Oct/02 12:10 AM Page 110

path to climb up the failure line. In stark contrast, the sand with 1% mica is brittle and shows a poten-
tial to collapse. The addition of just 1% of mica by weight has suppressed almost completely the
tendency to dilate. It is important to note that, because of the aspect ratio of the mica plates (of approx-
imately 50:1) compared to the rotund sand particles, 1% of mica by weight is approximately equivalent
to 25% of mica by number of grains.
Figure 7.12b shows the behaviour observed in a series of CVDSS tests on dumped sand samples
having mica contents ranging between 0% and 40% by weight. This illustrates that it is relatively small
quantities of mica that are critical, in terms of brittleness and collapse potential. Higher percentages of
mica increase the stability of the mix, despite the increase in void ratio. At 40% mica, the behaviour is
similar to that of a normally consolidated unstructured clay, with no tendency to dilate or collapse;
behaviour at large strains is clearly dominated by the mica, which gives rise to a lower  (23° cf 33°).
In this latter respect, the addition of mica is similar to the addition of kaolin described above.
It might be argued that the effect of the mica on the behaviour of the sand is simply a reflection of its
effect on void ratio. That this is not the case is demonstrated by the data in Figure 7.12c. This com-
pares the results of CVDSS tests on dumped sand samples having mica contents of 0% and 5% and void
ratios of 0.76 and 0.91, and a second sample with 5% mica, brought to a similar void ratio as the clean
sand. The two samples with 5% mica behave in a similar way, despite their differences in void ratio. The
clean sand and sand with 5% mica at similar void ratio behave quite differently, with the mica again sup-
pressing dilation.

7.3.3 Anisotropy of undrained strength in micaceous sands


A series of undrained triaxial compression and extension tests has been run on the model sand-mica
mix, using air-pluviated samples with increasing mica contents, and K0 consolidated before undrained
shear. The results are shown in Figure 7.13. In triaxial compression, the addition of mica actually
increases the initial stability of the structure, reducing undrained brittleness; this is in direct contrast to
the effects of adding kaolin to Ham river sand described in Section 7.2. In undrained triaxial extension,
small mica contents increase brittleness and give the sand an even greater potential to collapse; further
increases in mica content suppress dilation at large strains in extension and begin to stabilise the mate-
rial. With small mica contents, the sand is extremely weak when loaded in extension.
This huge difference in strength between undrained triaxial compression and extension was also found
in similar tests on the natural material and was seen to persist at relative densities of 55%. Figure 7.14a
shows the results of undrained triaxial compression and extension tests on samples from Batch A,
prepared by dry spooning at relative densities of 58% and 55%, respectively. In compression, the mate-
rial is ductile and has an undrained strength in excess of 180 kPa. In undrained extension, the material
is brittle and has a strength of only 10 kPa. Figure 7.14b provides a similar comparison, but between
samples prepared from Batch B by moist tamping at relative densities of 25 to 27%; again strengths in
extension are extremely low and there is no tendency to collapse in compression even at these low
relative densities.

mica content (%)

0
50 50
1
2.5
(σa - σr)/2 (kPa)

10
25 25
20
40

0 0
0 25 50 75 100 -30 -20 -10 0 10 20 30
Axial strain (%)
(σa' + σr')/2 (kPa)

-25 -25

Figure 7.13. Effects of mica on the behaviour of sand in undrained triaxial compression and extension (Hight et al.,
1999).

110
09031-02[1].qxd 18/Oct/02 12:10 AM Page 111

7.3.4 Effect of fabric in micaceous sands


Not only does the mica content vary in the natural materials, but, as noted above, so does its distribu-
tion and orientation within the sand. The method of preparing specimens of the micaceous sands for
testing was varied, therefore, as this changed the orientation and distribution of the mica. Thus, in sam-
ples prepared by air pluviation or dry spooning, the mica was relatively uniformly distributed and
aligned horizontally. In samples prepared by moist tamping, the mica distribution was less uniform and
its orientation more random; in addition, following the observations of Jang and Frost (1998), the local
porosity would have shown a greater variation, with the presence of larger and more collapsible pores.
Figure 7.15 compares the behaviour in undrained triaxial compression of samples prepared by dry
spooning and moist tamping. Figure 7.15a shows data for Batch A and Figure 7.15b for Batch B of the
material. The effect of preparation method is critical. A test specimen of material from Batch A formed
by dry spooning at a relative density of 40% dilates strongly while a specimen formed by moist tamp-
ing at a similar relative density collapses. The reverse pattern is observed in Batch B, with the speci-
mens formed by dry spooning showing the potential to collapse. The effect of preparation method of
Batch B in constant volume simple shear is illustrated in Figure 7.16. Again, the effect of fabric is most
pronounced, although both batches show contractant behaviour when samples are prepared by moist
tamping.
At similar relative density, a micaceous sand may, therefore, be ductile or brittle, with a potential to
collapse, depending on its fabric, i.e. on its method of placement and grading, and, therefore, on the dis-
tribution and orientation of the mica.

7.3.5 Collapse surface in micaceous sands


The concept of a collapse surface and its key features were described in Sections 5 and 6. Some of these
features in the artificial micaceous sand with 2.5% mica are illustrated by the test results shown in

Dr (%) Dr (%)
200
58.3 58.3
(σa-σr)/2 (kPa)

100

0
0 54.6 100 200 300 -20 54.6 -10 0 10 20
(σa'+σr')/2 (kPa) Axial strain (%)
-100

Figure 7.14a. CAU triaxial compression and extension tests on micaceous sand. Batch A, dry spooned (Hight et al.,
1999).
(σa-σr)/2 (kPa)

100 100

Dr (%) Dr (%)
26.9 26.9

50 50
(σa-σr)/2 (kPa)

24.9 24.9

0 0
0 26.8 50 100 150 -10 26.8 0 10 20
(σa'+σr')/2 (kPa) Axial Strain (%)

-50 -50

Figure 7.14b. CAU triaxial compression and extension tests on micaceous sand. Batch B, moist tamped (Hight
et al., 1999).

111
09031-02[1].qxd 18/Oct/02 12:10 AM Page 112

120
Moist tamped
Dry spooned Dr (%)
39.6

(σa-σr)/2 (kPa)
80 Collapse

40

37.1

0
(a) Batch A
120

Dr (%)
31.2
(σa-σr)/2 (kPa)

80
26.9

40 24.9

26.3 27.8
36.7
0
0 40 80 120 160
(σa'+σr')/2 (kPa)
(b) Batch B
Figure 7.15. CAU triaxial compression tests on micaceous sand. Effect of preparation method and batch (Hight
et al., 1999).
Horizontal shear stress (kPa)

60
Dr (%)
Moist tamped 44.7
Dry Spooned
30.4
40

43.7
20

31.4
0

Figure 7.16. Constant volume simple shear tests on micaceous sand. Effect of preparation method, Batch B (Hight
et al., 1999).

112
09031-02[1].qxd 18/Oct/02 12:10 AM Page 113

50
Undrained triaxial compression
Constant t
Collapse

(σa-σr)/2 (kPa)
Drained

25
e=0.934
Undrained

e=0.917
0

20
Axial strain (%)

10

0
0 25 50 75 100

(σa'+σr')/2 (kPa)

Figure 7.17. Collapse in sands with 2.5% mica prepared by moist tamping (Hight et al., 1999).

Figure 7.17. The samples were formed by moist tamping and, as can be seen, moist tamping produces
samples that collapse under triaxial compression; this allows the features of the collapse surface to be
investigated in relatively simple tests. The fact that the effective stress path from the compression test
defines part of the collapse surface is demonstrated by a second test on an identical moist tamped sam-
ple, which was held at constant t (( a  r)/2) and swelled back under stress controlled conditions.
The stress path followed and the response on crossing the collapse surface are shown in Figure 7.17. The
development of strain accelerated abruptly on crossing the surface. Drained conditions were maintained
and the stress path continued at constant t until the full friction angle was mobilised. Had drainage at
the time of collapse been impeded, then the effective stress path would have followed the collapse sur-
face as indicated in Figure 7.17.
On the basis of the results presented previously, the collapse surface for pluviated micaceous sands
would be restricted to high values of , and would be particularly low at  of 90°, i.e. for extension
loading.
It is evident from the data that has been presented that mica can introduce a collapse potential for cer-
tain directions of loading at values of relative density at which collapse would not normally be expected.
In the same vein Yoshimine et al. (1998) present data (refer Figure 7.18) that shows that with increasing
fines content the relative density at which there can be zero residual strength increases.

7.4 The effect of silt fines on sand behaviour


The effects of fines described to date relate to clay and to sand size mica plates. Adding rotund silt par-
ticles has a different effect, tending to stabilise loose sand. This is evident from data presented by
Kuerbis et al. (1988) for undrained triaxial compression and extension tests on isotropically consoli-
dated sand with increasing proportions of silt (Figure 7.19). The silty sands were formed by slurry depo-
sition and the silts presumably occupied the void space rather than increasing the separation between

113
09031-02[1].qxd 18/Oct/02 12:10 AM Page 114

Relative density for essentially zero residual strength,


0

10

20
Dr (%) at IB=1.0
30

40

50

Triaxial compression
60 Triaxial extension
Simple shear

70
0 4 8 12 16
Fines content (%)

Figure 7.18. Relative density for essentially zero residual strength (Yoshimine & Ishihara, 1998).

Figure 7.19. Undrained triaxial compression and extension tests on isotropically consolidated silty sands (Kuerbis
et al., 1988).

sand grains. This effect of silt has been confirmed for K0 consolidated sands. Figure 7.20 compares the
behaviour in undrained triaxial compression of a clean sand and the same sand with 1% mica, 2.5%
mica and 2.5% angular silt, all formed by slurry deposition. As illustrated by Figure 7.21, the effect of
fines on the stability of the structure will depend on the shape, location and quantity of the fines.

114
09031-02[1].qxd 18/Oct/02 12:10 AM Page 115

e
0 % mica 0.792
1 % mica 0.829
50 50
2.5 % mica 0.893

(σa-σr)/2 (kPa)
2.5 % HPF4 silt 0.780
(σa-σr)/2 (kPa)

2.5% silt

2.5% mica
25 25 e
0% mica 0 % mica 0.792
1% mica 1 % mica 0.829
2.5 % mica 0.893
2.5 % HPF4 silt 0.780
0 0
0 25 50 75 100 0 5 10 15 20 25 30
(σa'+σr')/2 (kPa) Axial strain (%)

Figure 7.20. Undrained triaxial compression of sands with 0% and 2.5% mica and 2.5% HPF4 silt.

(a) (b)

(c) (d)

Figure 7.21. Importance of shape and location of additives in sand.

7.5 Some features of the behaviour of silt


Silts are often regarded as difficult or intermediate soils. This stems in part from the difficulty in sam-
pling and handling them. Unlike sands, and because of their lower permeability, silts are able to sustain
a small suction, giving them some coherence. Being at a low effective stress when unconfined they will
naturally show a tendency to dilate. As with sands, silts can be deposited at a range of void ratios and
have a multiplicity of normal consolidation lines. Again, in common with sands and glacial tills, their
behaviour is not uniquely related to void ratio and stress history, but depends on initial porosity, stress
level and stress history. The effect of initial porosity of HPF4 silt, which has the grading shown in
Figure 7.22, is evident from data from undrained triaxial compression tests on isotropically consoli-
dated samples in Figure 7.23.
An investigation of the same silt in a dense state (e0  0.65, Dr  88%) is reported by Zdravkovic
and Jardine (2000). Figure 7.24 defines the behaviour of the silt after K0 consolidation and swelling of
samples prepared by pluviation. The silt has no memory. It is highly dilatant at large strains. The
mobilised friction angle at phase transformation is 32° in compression and 25° in extension. The max-
imum mobilised friction angle was 37° in compression and 30° in extension.

115
09031-02[1].qxd 18/Oct/02 12:10 AM Page 116

100

90

80
Silt A
Ham River Sand

70

60
% passing

50

40

30

20

10

0
0.001 0.01 0.1 1
Grain Size (mm)

Figure 7.22. Grain size distribution of HPF4 silt compared to Ham river sand.

400 400

37.7%
300 300 39.7%
(σa-σr)/2 (kPa)
(σa-σr)/2 (kPa)

200 39.7% 200


37.7%

100 42.6% 100


42.6%

0 0
0 100 200 300 400 500 0 4 8 12
(σa'+σr')/2 (kPa) Axial strain (%)

Figure 7.23. Effect of initial porosity on behaviour of isotropically consolidated HPF4 silt in triaxial compression
(Ovando-Shelley, 1986).

8 LOCALIZATION AND BEHAVIOUR AT LARGE DEFORMATIONS

8.1 Generalities on post-failure and localization


Critical state soil mechanics (CSSM) described in Section 2 considers that, during shearing, the soil
deforms homogeneously and reaches critical state at large strains or deformations, whether the speci-
men is normally or overconsolidated. When specimens of normally consolidated or slightly overcon-
solidated clays, or loose cohesionless materials are subjected to triaxial compression tests, they indeed
deform in an essentially homogeneous manner up to large strains. If there are some end restraints, the
specimens will take a barrel shape but there is no formation of well-defined shear zones. Conditions
reached at large strains are then in accordance with CSSM. On the other hand, overconsolidated clays
or dense cohesionless materials subjected to triaxial compression tests show the formation of one or

116
09031-02[1].qxd 18/Oct/02 12:10 AM Page 117

Figure 7.24. Effective stress paths for HPF4 silt in undrained triaxial compression and extension (Zdravkovic &
Jardine, 2000).

117
09031-02[1].qxd 18/Oct/02 12:10 AM Page 118

Figure 8.1. Unconsolidated undrained triaxial test on Todi clay showing post-rupture behaviour (from Burland,
1990).

several well defined shear bands (there is localization), followed by essentially rigid bodies sliding over
others (Viggiani et al., 1993; Tillard-Ngan et al., 1993; Desrues et al., 1996).
The phenomenon of localization is illustrated in Figure 8.1 using the results of an undrained triaxial
test carried out on stiff Todi clay. The instrumentation for this test included, in addition to an external
axial displacement transducer and a pore pressure transducer at the base, local strain transducers
directly on the soil specimen and a local pore pressure probe (see the schematic specimen in Figure 8.1).
For the test presented in Figure 8.1, the failure surface passed outside the local strain transducers and
close to the pore pressure probe. Burland (1990) made several remarks in relation to the results shown
in Figure 8.1, among which: (a) prior to peak strength the local strains are less than the overall strains,
as was observed by many other authors; (b) after peak strength is reached the local axial strains decrease
as a result of the unloading process; and (c) localization essentially coincides with peak strength.
When displacement along the shear surface progresses, depending on the shape of soil particles,
there may be an orientation of particles in the direction of movement and a progressive decrease in shear
strength towards the residual strength.
From what has been previously said, three different levels of strength can generally be defined for
soils: (a) the peak strength of the soil in the pre-yield (or overconsolidated) range, which reflects the
true friction angle and cohesion as expressed by Hvorslev (1937; see Section 2.1) and, in many natural
geomaterials, bonding between particles or aggregates defined here as microstructure; (b) the critical
state strength envelope associated with large and homogeneous deformations or post-rupture strength;
and finally, (c) the residual strength envelope associated with oriented particles and very large dis-
placements along a slip surface. As shown in Figure 8.2, these three levels of strength were already
identified by Tiedemann in 1937.

8.2 Localization in cohesionless materials


Yoshida & Tatsuoka (1997) investigated the deformation characteristics of shear bands in cohesionless
materials and their relation to the physical characteristics of particles. These authors performed a series
of plane strain compression tests on 12 poorly graded materials with uniformity coefficient Cu between
1.2 and 2.13 and mean particle diameter D50 between 0.174 and 6.80 mm. All the tests were performed
on dense specimens so that localization could occur. The shear deformation of the shear band necessary
after peak to reach residual or steady state conditions, (u*) s res, and the shear band thickness were

118
09031-02[1].qxd 18/Oct/02 12:10 AM Page 119

2.0

Peak strength (intact specimens)


Peak strength (remoulded specimens)
1.5 Residual strength (intact specimens)
Residual strength (remoulded specimens)
τ (kg/cm2)

1.0

0.5

0
0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5
σ (kg/cm2)

Figure 8.2. Failure envelopes of Weser-Elbe clay (after Tiedemann, 1937).

Figure 8.3. Influence of particle diameter D50 on shear band thickness (a) and (u*s)res (b) (from Yoshida & Tatsuoka,
1997).

determined; they are plotted in Figure 8.3 as a function of D50. It can be seen that both shear displace-
ment (u*)
s res and shear band thickness increase as D50 increases. Figures 8.3a and b also show that these
shear band characteristics decrease as the confining pressure increases. Yoshida & Tatsuoka (1997) also
indicate that shear band thickness becomes larger as particles are rounder or less crushable.

119
09031-02[1].qxd 18/Oct/02 12:10 AM Page 120

1.00
(D50 = 0.32 mm; Cu = 1.7)
σ'c = 60 kPa
0.90
Void ratio (global or local)

0.80 Void ratio evolution


Local Global
rfdt3
0.70 rfdt4
rfdt6
rfdt7
0.60 rfdt8
rflt2
rflt1
0.50
0 0.10 0.20 0.30 0.40 0.50
Axial strain (global)

Figure 8.4. Global and local evolution of the void ratio in loose and dense Hostun RF sand specimens subjected to
triaxial compression test (from Desrues et al., 1996).

Using computer tomography (X-ray scanner), Desrues et al. (1996) observed the local evolution of
soil density inside triaxial specimens of sand subjected to constant effective confining pressure com-
pression tests, and in particular the development of localization. Figure 8.4 shows the evolution of both
average and local (in the shear band) void ratios with global axial strain. Several conclusions can be
drawn: (a) on the basis of average void ratios observed at large deformation, which are equal to 0.85 and
0.77 for the loose and dense specimens respectively, the critical state concept would not apply; (b) the
void ratio in the shear bands observed in the initially dense specimens (mostly the black dots) starts
increasing at some stage to reach the value obtained at large strains in the loose specimens, indicating
that the principle of critical state applies when conditions in the shear bands are considered; when spec-
imens are contractant, the average void ratio is representative of critical state conditions. It is worth not-
ing that Casagrande & Watson (1938) stated these points more than sixty years ago. An important
conclusion of Desrues et al’s study is that the concept of critical state applies in dense sands, but locally;
the shear strength is then characterized by the critical state friction angle, cs.

8.3 Post-rupture strength in microstructured soils


On the basis of direct shear tests performed on intact stiff Todi clay, Calabresi (1980) distinguished two
stages of post-peak strength reduction: the first one occurs just after peak, at a relative displacement of
less than 1 mm; the second takes place gradually as displacement increases. These two stages were con-
firmed for several stiff clays by Burland (1990) and Burland et al. (1996). As shown in Figure 8.1, after
peak, the curve of deviator force versus notional overall strain falls abruptly to a kind of plateau. This
plateau was termed “post-rupture strength”. These authors also showed that this post-rupture strength lies
on or very close to the critical state or fully softened strength envelope. The strength reduction from peak
to post-rupture thus mostly corresponds to the breakage of the bonds between particles or aggregates.
These results are supported by laboratory and field observations made by Calabresi & Manfredini
(1973) and Skempton (1977). Measuring the strength along fissures and joints along which no dis-
placement had occurred previously, these authors measured strengths very close to critical state strength
(see Section 12.1 and Figures 12.2 and 12.3).
However, if the soil contains stiff aggregates or lithorelicts, the large deformation strength envelope
obtained in triaxial testing of a natural soil can be above the critical state strength envelope defined on
the same material, reconstituted or destructured by compression beyond the limit state curve. This has
been shown by Leroueil et al. (1997) for different clays, sensitive eastern Canada clays and softened clay
shales from southern Italy. As shown in Figure 8.5 for the Saint-Jean-Vianney clay, the large deformation

120
09031-02[1].qxd 18/Oct/02 12:10 AM Page 121

Saint-Jean-Vianney clay
300
Ip = 8 - 10%; σp' ~ 1000 kPa ˚ 5˚
44.4 30.
Natural soil ' = φ'=

φ
Overconsolidated
(σa' - σ'r )/2 (kPa)

Normally consolidated
200

Reconstituted soil
CSL
100

0
0 100 200 300 400 500 600
(σa
' + σ'r )/2 (kPa)

Figure 8.5. Large deformation strength of the natural and reconstituted Saint-Jean-Vianney clay (from Saihi et al.,
2002).

0.8
Friction coefficient τ /σ 'n

0.6
φ' or φ 'cr
n.c.
0.4

0.2 φ'
r
Rolling
shear Transitional Sliding shear
0
0 20 40 60 80 100
Clay fraction (%)

Figure 8.6. Ring shear tests on sand-bentonite mixtures (after Lupini et al., 1981, and Skempton, 1985).

friction angle of the reconstituted soil is in the order of 30° whereas the large deformation friction angle
of the natural and overconsolidated clay is about 44°. This was attributed to the fact that sensitive clays
from eastern Canada are made up of aggregates that are stiff and extremely angular. Soccodato (2002*)
observed a similar behaviour in Fucino clay.

8.4 Residual strength


Residual conditions, which are reached after large displacements along a slip surface, are characterized
by a friction angle, r, and zero cohesion. In soils made of rotund particles, shearing does not introduce
any preferential orientation of the particles and r is equal to or only slightly smaller than cs (turbulent
shear according to Lupini et al., 1981); in soils made up of platy particles, shearing induces particle
orientation in the direction of shearing and r is significantly smaller than cs (sliding shear according
to Lupini et al., 1981). Values of r in clayey materials are given in Section 16.6 and Figures 16.4
and 16.5.
The behaviour of soils containing both rotund and platy particles has been studied by Lupini et al.
(1981). As shown in Figure 8.6, / n values obtained at residual conditions, and thus r values, are

121
09031-02[1].qxd 18/Oct/02 12:10 AM Page 122

high when the percentage of platy particles is small (turbulent shear dominant), and small when the per-
centage is high (sliding shear dominant). In between, there is a range of percentage in which the behav-
iour is transitional: there are rotund particles that impede the development of a perfectly smooth shear
surface, with a resulting residual friction angle intermediate between those of the soils that would be
made up of rotund or platy particles. The development of a perfectly smooth shear surface may also be
impeded in clayey soils, as clay shales, that contain hard lithorelicts. The residual strength envelope of
the natural soil may then be above the residual strength envelope of the same reconstituted material.
This is shown in Figure 12.5 for the Laviano tectonized clay shale from southern Italy.

8.5 Brittleness
Brittleness is associated with strain softening behaviour, i.e. a reduction in shear strength as strains
increase. It can be observed in most natural soils: stiff plastic clays and shales with residual strength
smaller than peak strength (Bishop, 1967); microstructured soils when bonds between particles are bro-
ken due to shearing; dense sands during drained shearing; loose sands during undrained loading (Sladen
et al., 1985); soils subjected to stress axis rotation during shearing or consolidated under an inclined
major principal stress (see Section 6); collapsible unsaturated soils. Also, as indicated in Hight &
Leroueil (2002*), brittleness has important practical implications since it controls progressive failure
and post-failure phenomena.
Bishop (1967) characterized brittleness by a brittleness index, IB, defined as follows:

p  r
IB  ,% (8.1)
p

where p and r are the peak and residual strength defined under the same effective normal stress.
However, as indicated by Vaughan & Hamza (1977) and by Chandler (1984), the brittleness index
alone is not sufficient to characterize the susceptibility of a soil to progressive failure or to flow; the rate
at which the strength decreases from peak to ultimate strength, and thus the energy necessary to go from
peak to ultimate conditions are also important. D’Elia et al. (1998) proposed a generalised brittleness
index, IGB, defined as follows:

 p   mob
IGB  ,%
p (8.2)

where mob is the mobilised shear stress at the considered strain or displacement. IGB thus varies with
strain or displacement from 0 at the peak to a value equal to IB at large strain or displacement
(Figure 8.7). In the generalised form proposed by D’Elia et al. (1998) and shown in Equation 8.2, the
Shear stress

τp 0

τ p - τ (ε or ∆ I )
IGB =
τp

τr τp - τ r
IGB = IB = τp

1
Strain or displacement

Figure 8.7. Brittleness of soils (from D’Elia et al., 1998).

122
09031-02[1].qxd 18/Oct/02 12:10 AM Page 123

brittleness index is associated with a given stress path or test, and must not be seen as a fundamental or
intrinsic characteristic of a soil.

9 INFLUENCE OF STRAIN RATE AND TIME

9.1 One-dimensional (and isotropic) compression


It is well known that when a soil specimen is loaded under a constant total vertical stress in one-
dimensional conditions, it continues to settle after full dissipation of excess pore pressures. This is
the secondary consolidation phase usually characterized by the secondary compression index
Ce  e/ log t. Mesri & Godlewski (1977) and Mesri et al. (1995) showed that Ce is related to the
compression index Cc of the soil and, more precisely, that the ratio Ce/Cc is a constant for a given soil.
Figure 9.1 shows typical results. This has been confirmed for a large variety of geotechnical materials
by Mesri and co-workers, and many other researchers. An interesting aspect is that, as indicated in Table 9.1,
Ce/Cc remains within a very narrow range for each soil type, varying from 0.01–0.03 for granular soils to
0.05–0.07 for peat (Mesri, 1987; Mesri et al., 1995). Coop & Airey (2002*) found 0.013 for carbonate
sands. A direct consequence of secondary consolidation is that, when the soil is reloaded, it shows an
apparent preconsolidation pressure larger than the previously applied stress, associated with its lower
void ratio. As schematically shown in Figure 9.2a, after 10 years, the preconsolidation pressure would
have increased from vo to p10; it would be p10 000 after 10 000 years. An indirect consequence is that,
as shown by Crawford (1964), Bjerrum (1967), Sällfors (1975) and Leroueil et al. (1983a, 1985a), when
the soil is reloaded slowly, a reduction of the apparent preconsolidation pressure of the soil is observed

Figure 9.1. Values of secondary compression index Ce and compression index Cc for Saint-Hilaire clay (from
Mesri et al., 1995).

Table 9.1. Viscous parameters for geotechnical materials.

 ′
/  log  (%)
Material Ce/Cc m ′

Granular soils, including rockfill 0.02  0.01 10033 2.37.2


Shale and mudstone 0.03  0.01 5025 4.79.6
Inorganic clays and silts 0.04  0.01 3320 7.212.2
Organic clays and silts 0.05  0.01 2517 9.614.8
Peat and muskeg 0.06  0.01 2014 12.217.5

Note: Types of material and Ce/Cc values are from Mesri et al. (1995).

123
09031-02[1].qxd 18/Oct/02 12:10 AM Page 124

(dashed compression curve in Figure 9.2), i.e. the apparent preconsolidation pressure is dependent on
the rate of strain during reloading.
Leroueil et al. (1985a) performed a variety of oedometer tests on different natural clays: Multiple
Stage Loading tests with reloading at the end of primary consolidation or after 24 hrs, Constant Rate of

Figure 9.2. Increase of the preconsolidation pressure and development of the limit state curve with secondary con-
solidation (after Tavenas & Leroueil, 1977).

Figure 9.3. Effects on one-dimensional compression of Saint-Polycarpe clay of: (a) strain rate; and (b) temperature
(after Marques, 1996).

124
09031-02[1].qxd 18/Oct/02 12:10 AM Page 125

Strain (CRS) tests, Controlled Gradient tests and long-term creep tests. They found that the behaviour
is controlled by a unique vertical effective stress-vertical strain-vertical strain rate ( 1– 1– 1) [or
( 1–e– e·)] relationship. This rheological model, originally proposed by Suklje (1957), was advocated,
in particular, by Lowe (1974). It has also been confirmed experimentally by Imai & Tang (1992) for
Yokohama Bay mud (see also Imai, 1995). It can be described by compression curves deduced from
CRS tests performed at different strain rates, as shown in Figure 9.3a. Leroueil et al. (1985a) also
showed that the ( 1– 1– 1) relationship can be described by two curves, one giving the variation of the
preconsolidation pressure P with strain rate (see footnote 2):
.
p  f( 1) (9.1)

and the other presenting the normalised effective stress-strain curve:


1/ p( 1)  g( 1) (9.2)

This means that the ratio between the preconsolidation pressures, or effective stresses at the same
strain, at two different strain rates is a constant regardless of the strain or void ratio considered. This also
means that the horizontal distance between two compression curves corresponding to different strain
rates is constant in a 1 (or e)–log 1 diagram.
Figures 9.4a and b show the two curves represented by Equations 9.1 and 9.2 for Berthierville clay.
As seen in Figure 9.4a for Berthierville clay, and as observed for other clays, Equation 9.1 can generally

Figure 9.4. One-dimensional compression of Berthierville clay: (a) preconsolidation pressure as function of strain
rate and temperature; and (b) normalized effective stress-strain curve (from Boudali et al., 1994; Kabbaj, 1985).

125
09031-02[1].qxd 18/Oct/02 12:10 AM Page 126

be approximated by a linear relation in a log p–log 1 diagram


.
log p  A  (1/m) log 1 (9.3)

in which A and m are constants.


Leroueil et al. (1983a, 1985a) and Leroueil (1996) showed that the variation of the preconsolidation
pressure with strain rate is very similar for numerous non-organic clays, i.e. 7 to 15% per logarithm
cycle. The corresponding m value is generally between 17 and 35.
As it can be demonstrated that:

m  1/(Ce/Cc)  Cc/e (9.4)

such m values are equivalent to Ce/Cc values between 0.059 and 0.029, which coincides with the
results obtained by Mesri and co-workers (Table 9.1). Equation 9.4 also implies that the Ce/Cc and the
strain rate (m) approaches used for describing the viscous behaviour of soils during secondary consol-
idation are equivalent.
In a viscous model such as the ( 1– 1– 1) [or ( 1–e–e·)] model previously described, soil behaviour
.
is controlled by the same rheological law, regardless of whether consolidation takes place during its pri-
mary or secondary phase.
Since the Ce/Cc  constant model seems to apply for a large variety of soils during secondary con-
solidation (Table 9.1), the ( 1 – 1– 1) model previously described should also be general. Cotecchia
(2002*) showed its application to the hard Pappadai clay; Edil & den Haan (1994) showed its validity
for a fibrous peat; tests performed at different rates of loading by De Waal (1986) indicate that it is valid
for sand; finally, tests performed on chalk by Ruddy et al. (1989) and on salt rock by Aubertin et al.
(1999) also indicate that the model may be valid for soft rocks.
The implications of the effects of strain rate have been described in detail elsewhere (Leroueil et al.,
1985a; Leroueil, 1988, 1996) and only the main aspects will be presented here.
Assuming that primary consolidation is completed before 24 hours in conventional multiple stage
loading oedometer tests, the strain rate at the end of the loading periods, which can be associated with
the 24-hrs compression curve, is given by the following equation (Leroueil, 1988):

Cc
 24 h (s1 )  2  107 (9.5)
1  e0

This strain rate is thus on the order of 5  108 s1 for low compressibility clays and on the order of
107 s1 for highly compressible clays. This is smaller than the strain rates used in CRS tests that are
generally between 1.0 and 4.0  106 s1. As a result, the preconsolidation pressure and the effective
stress at any strain or void ratio measured in CRS tests are larger than those measured in conventional
24-hrs oedometer tests (Figure 9.5). A compilation made by Leroueil (1996) for a variety of clays from
different countries indicates a ratio p CRS/ p conv typically equal to 1.25.
When a soil specimen is instantaneously loaded, the strain rate is high during the early stages of con-
solidation and, consequently, the effective stresses go to compression curves corresponding to these
high strain rates. However, with time, consolidation progresses and the strain rate decreases. As a result,
the effective stress-strain condition of the soil moves towards compression curves corresponding to
lower strain rates. It follows that, in multiple stage loading tests, the effective stress-strain relationship
really followed consists of a succession of steps as shown in Figure 9.6, very different from the com-
pression curve obtained by joining the points obtained at the end of the loading periods.
Even within a consolidating sample, the effective stress-strain relationships followed in different
parts of the sample depend on the local strain rate history. The test results presented by Mesri et al.
(1995) can be used for illustrating this point. These authors performed an isotropic compression test on
four 125 mm-long specimens of St-Hilaire clay connected in series. They present in detail the variation
with time of the axial strain of the four sub-elements of the 500 mm-long specimen and the excess pore
pressures measured in between for the pressure increment from 97 to 138 kPa during the primary con-
solidation phase. In these tests, the compression was isotropic, but drainage was one-dimensional. From
these data, it was possible to estimate the effective stress-void ratio curve for the different sub-elements.
These curves are drawn in Figure 9.7, from I (initial conditions) to P. Also drawn on the figure is a
fictitious secondary consolidation phase (PF). It can be seen that the compression curves followed
during primary consolidation vary with the position of the sub-element. Near the drainage boundary
(sub-element 1), the strain rate is very high during the early stages of consolidation and, as a result, the

126
09031-02[1].qxd 18/Oct/02 12:10 AM Page 127

Figure 9.5. Typical comparison between oedometer test curves obtained in CRS and conventional oedometer tests
(Batiscan clay, Leroueil et al., 1983b).

Figure 9.6. Multiple stage loading (MSL) oedometer tests.

effective stress goes to compression curves associated with high strain rates. On the other hand, for the
sub-elements that are far from the drainage boundary (sub-elements 3 and 4), the strain rates are much
smaller during the same period and the effective stress remains close to compression curves associated
with these strain rates. However, when the soil is approaching the end-of-primary consolidation (point P),
the strain rate becomes more uniform in the specimen and the compression curves deduced from the
different sub-elements converge. After the end-of-primary consolidation, the entire specimen would be

127
09031-02[1].qxd 18/Oct/02 12:10 AM Page 128

Effective σ ' (kPa)


100 stress,
90 110 120 130 140 150 160 170 180
1.90
I Primary consolidation
σ'vf
4 3 2 1 in the sub-elements
1.85 e - σ v'
-5 -1
e = 10 min
1.80 -6
-5 min 3 x 10
10 -6
10 -7
Void ratio, e

1.75 3 x 10
-7
10
1.70
1 EOP P

e (min )
Secondary -6

-1
10
500 mm

1.65 2 consolidation
(Cα e /Cc = 0.03 10-7
3 or m' = 33)
1.60 F -8
4
10
10
-10 10-9
1.55
Figure 9.7. Consolidation of the Saint-Hilaire clay for pressure increment from 97 kPa to 138 kPa (after Mesri
et al., 1995; from Leroueil & Marques, 1996).
Excess pore pressure u (kPa)

150
Void ratio e

Case 1
e 100
(min.-1 )
Case 2
Case 1 -3
10
Case 2 50
10 -4 Case 3
Case 3 -5
10

0
Vertical effective stress 0 5 10 15 20 25
2
(kPa, log scale) Time t (10 min.)
(a) (b)

Figure 9.8. Excess pore pressure generation during relaxation tests under one-dimensional conditions (from
Yoshikuni et al., 1994).

in secondary consolidation and would settle from P to F. This test allows one to define the complete set
of isotaches (lines of equal strain rate or change in void ratio) for the St-Hilaire clay. Imai & Tang (1992)
observed similar behaviour in a consolidating clay layer consisting of 7 sub-specimens in series.
While the strain rate model accurately describes the behaviour of clays when strain is increasing, it
is not the case when the axial strain remains constant, as in relaxation (constant strain) tests. Yoshikuni
et al. (1994 & 1995) performed special oedometer tests with different phases of consolidation (open
drainage) and relaxation (closed drainage). In one series of three tests, these authors stopped secondary
consolidation under an effective stress of 314 kPa by closing the drainage and monitored the excess pore
pressure developing with time. As shown in Figure 9.8, the higher the strain rate at which secondary
consolidation is stopped, the higher is the generated excess pore pressure, indicating that, when strain

128
09031-02[1].qxd 18/Oct/02 12:10 AM Page 129

log σ'
1

dε e du A (ε )
dε p
Cs
1 + eo

εv

Figure 9.9. Strain components in relaxation test.

rate decreases, the tendency to creep also decreases. Holzer et al. (1973) report similar results under
isotropic stresses. One explanation for the behaviour observed in relaxation tests is that the strain-rate
model (Equations 9.1 to 9.3) applies only to the plastic component of strain, so that, in relaxation tests
where the total strain is constant, the increase in plastic strain p associated with creep is compensated
by a decrease in elastic strain e, and thus a decrease in effective stress (Figure 9.9).
 e  p  constant (9.6a)

where e and p are respectively the elastic and plastic components of strain.
So, in relaxation tests:
·e   · p (9.6b)

The rate of change in pore pressure (u· ) following closing of the drainage when the soil is in second-
ary consolidation at a strain rate ·   · p (point A in Figure 9.9) would be:

2.3(1  e0 ) ′ p
u   (9.7)
Cs

in which  is the effective stress at the time of closing drainage and Cs is the recompression index.
When the applied load is decreased and soil swells, viscous effects are also observed. After primary
swelling associated with pore pressure equalization, secondary deformations in swelling or in compres-
sion develop. Such behaviour has been observed and discussed by many authors (Mesri et al., 1978;
Urciuoli, 1992; Feijo & Martins, 1993; Yoshikuni et al., 1993). It can be illustrated by the results
reported by Feijo & Martins. Special oedometer tests were performed on organic Sarapui clay from
Brazil (see Almeida & Marques, 2002*), in which several specimens were first loaded in the normally
consolidated range and then, at the end-of-primary consolidation, unloaded in one step to various over-
consolidation ratios. The specimens were left under constant stress for more than 5 months and the
observations are shown in Figure 9.10. For the specimens at OCR of 1.5 and 2, after some swelling, the
soil starts to compress again in a manner similar to what is observed with normally consolidated clays.
For OCRs of 6, 8 and 12, the soil shows some secondary swelling during the entire test. At an OCR of
4, secondary deformations are not significant. This kind of “neutral OCR” may vary from one soil to
another and certainly with the strain rate at the time of unloading.

9.2 Peak strength envelope and limit state curve


Sällfors (1975) and Tavenas & Leroueil (1977) indicated that the effects of time and strain rate on the
preconsolidation pressure can be generalized to the entire limit state curve of the soil. As indicated in

129
09031-02[1].qxd 18/Oct/02 12:10 AM Page 130

-8
OCR

(swelling)
12.0

-6
8.0

6.0
Volumetric strain, ε v (%)

-4
4.0

-2

2.0

0
(compression)

1.5

2
-1 2 3 4 5 6
10 1 10 10 10 10 10 10
Time (min.)

Figure 9.10. Volumetric strains observed after unloading of Sarapuí clay (after Feijó & Martins, 1993).

100
0

L
3 =

CS
(σ'a - σ'r )/2 (kPa)

σ'

CI
U,
50 0
CI D, 0

.5
CI .0 5 2 5
CI
U, .0

%
D,

/h
0 0

0 . %/ h / h
1
%
/h
%

0
0 50 100 150 200
(σ'a + σ'r )/2 (kPa)
Figure 9.11. Effect of strain rate on the strength of Saint-Vallier clay (after Lo & Morin, 1972).

Figure 9.2, secondary consolidation, which induces a decrease in void ratio and thus an increase in pre-
consolidation pressure also induces an increase in size of the entire limit state curve. Also, for an aged
natural clay, the slower the stress changes towards the limit state curve, the smaller the stress conditions
should be at gross yielding.
Triaxial tests performed by Lo & Morin (1972) on St-Vallier clay (Figure 9.11) demonstrate a strain
rate effect on the strength envelope of the overconsolidated soil, i.e. the upper part of the limit state

130
09031-02[1].qxd 18/Oct/02 12:10 AM Page 131

120
Numbers close to the points -6
CAU ε 1 = 1.4 x 10 s
-1
-8 140
are strain rates in 10 s-1 -6
CIU ε 1 = 1.4 x 10 s
-1
23
CAD ε 1 = 2.3 x 10 s
140 -7 -1

(kPa)
100 140 -9
CAU ε 1 = 6.0 x 10 s
-1
140 3
-6 0.6 CA-creep
σa' - σ 'r

10 0.2
From one-dimensional
2

80 ε vs (s -1)
10
-7 0.2
CRS and creep tests
-8 0.3
10 0.6 Compression tests at
-9
10 constant σ a' / σ r ' ratio
60
150 70 90 110 130 150
σa' +σ ' CS
L
r
(kPa)
2

li n e
(kPa)

100 -1 Ko
εvs = 10 -6
s
σa' - σ' r

εvs = 10 -9 s-1
2

50

0
0 50 100 150 200 250 300
σ'a + σ'r
(kPa)
2
Figure 9.12. Influence of strain rate on the limit state curve of Mascouche clay (after Marchand, 1982, and Leroueil
& Marques, 1996).

curve, similar to that on the preconsolidation pressure. Similar results were obtained for other eastern
Canada clays (Tavenas et al., 1978; Vaid et al., 1979; Leroueil & Tavenas, 1979).
Marchand (1982) reported triaxial tests performed on Mascouche clay, from eastern Canada. The
peaks obtained in CIU, CAU and CID tests, and the effective stress applied in triaxial drained creep tests
in which failure was reached are shown in the insert of Figure 9.12 with their associated strain rate (min-
imum strain rate reached before failure in creep tests; see Section 9.6). All the data are relatively well-
organized, confirming an effect of strain rate on the peak strength envelope of overconsolidated and
sensitive clay. The one-dimensional CRS and creep test results, also presented on the figure, indicate an
effect of strain rate on the preconsolidation pressure similar to that on the strength envelope. It is thus
the entire limit state curve that is strain rate dependent.
The viscous behaviour of Berthierville clay (see Figure 9.4 for its one-dimensional rheological law)
has been investigated by Boudali (1995). Continuous and step-loaded triaxial compression tests were
performed along K  3/ 1  constant stress paths. From the test results and for the different stress
paths followed, the following conclusions were drawn:
– Along the different stress paths, yielding at a given strain rate is well defined, and the mean effective
stress at gross yielding varies with the strain rate in the same manner as the preconsolidation pres-
sure (Equation 9.3). This implies that the limit state curves obtained for different strain rates have the
same shape, as shown in Figure 9.13a for Berthierville clay.
– The stress-strain curves obtained at different strain rates along a given K  constant stress path can
be normalised with respect to the gross yield stress value corresponding to the considered strain rate.
– The effective stress-strain-strain rate model proposed by Leroueil et al. (1985a) can be extended to
the entire stress diagram. Such a model can be schematized by a limit state surface having different
peels, each one being associated to a different strain rate (see Figure 10.8).

131
09031-02[1].qxd 18/Oct/02 12:10 AM Page 132

Type of ε vs = 10 -7(s -1)


40 40
test
5˚ 20˚C 35˚
ε -1
vs (s ) Triaxial
-6 Oedometer
10
-7 30
30 10
10 -8
T = 20 ˚C

q (kPa)
q (kPa)

20 20
4 3 43
0. 0.
= 0 .5 0 = 0 .5 0
K K
5 5
10 0 .6 10 0 .6

0 .8 0 0 .8 0

1.00 (a) 1.00 (b)


0 0
0 10 20 30 40 0 10 20 30 40
p' (kPa) p' (kPa)

Figure 9.13. Variation of the limit state curve of Berthierville clay with: (a) strain rate and (b) temperature (after
Boudali, 1995).

0.4 ˚
= 32
Start of shear φ'
Peak shear stress
0.3 Symbol ε (%/h)
OCR = 1
0.051
0.50
σ r' )/ 2 σ 'vm

5.0 OCR = 2
0.2 49.0
(σa' -

0.1 OCR = 4
OCR = 8

-0.1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
( σ'a + σ' r )/ 2 σ'vm

Figure 9.14. Normalized effective stress paths and strengths as obtained on resedimented Boston Blue clay in
undrained compression tests performed at different OCRs and strain rates (after Sheahan et al., 1996).

It is worth mentioning that the model previously described has limitations. In particular, Hight
(1983), Hight et al. (1987) and Sheahan et al. (1996) who performed anisotropically consolidated
undrained triaxial tests on reconstituted Lower Cromer till (Ip  13%) and Boston Blue clay (Ip  23.7%)
respectively did not observe strain rate effects on the strength envelope of the soil in the overconsoli-
dated range (Figure 9.14).

9.3 Stress strain behaviour


9.3.1 Small strain moduli
The effect of strain rate on small strain moduli has been the object of numerous studies (Isenhower &
Stokoe, 1981; Rampello & Silvestri, 1993; Zavoral & Campanella, 1994; Stokoe et al., 1995; Tatsuoka

132
09031-02[1].qxd 18/Oct/02 12:10 AM Page 133

Figure 9.15. Relationship between the secant Young’s modulus defined at different small strains obtained on
Metramo silty sand (after Santucci de Magistis et al., 1999; from Tatsuoka, 2000).

Figure 9.16. NG coefficient versus secondary compression index (from Lo Presti et al., 1996).

et al., 1995; d’Onofrio, 1996; and others). Some of these results were presented by Tatsuoka et al.
(1997, 2000), Lo Presti et al. (1997) and Tatsuoka (2000), and are shown in Figure 4.4. It can be seen
that the effect of strain rate on small strain Young’s modulus is not very important. From published data,
it appears that the strain rate effect is: (a) relatively small and less than 6% per tenfold increase in strain
rate in cohesive soils; (b) smaller in cohesionless materials than in clayey soils. Lo Presti et al. (1997),
Santucci de Magistris et al. (1999) and Tatsuoka et al. (2001) show that the effect is increasing with
strain. As shown in Figure 9.15 for the Metrano silty sand (Santucci de Magistris et al., 1999), almost
insignificant at a strain of 106, the effect of strain rate on the secant Young’s modulus increases with
strain to reach about 5% per logarithm cycle of strain rate at a strain of 1.5  105.
As indicated in Section 4.4, the increase in Go with time during secondary consolidation is often
characterized by NG (Equation 4.12). This parameter is soil dependent. Afifi & Richart (1973) indicated
an increase of NG when the size of the particles, characterized by D50, decreases; Kokusho et al. (1982)
indicated a relationship between NG and Ip; Mesri (1987) proposed an equation based on considerations
on secondary consolidation and Ce/Cc; Lo Presti et al. (1996) showed a relationship between NG and
Ce (Figure 9.16).

133
09031-02[1].qxd 18/Oct/02 12:10 AM Page 134

Figure 9.17. Results of undrained triaxial tests on NSF clay: (a) stress-strain relationship at small strains;
(b) Young’s moduli with stress (after Shibuya, 2001).

9.3.2 Moderate strains


The influence of strain rate at moderate strains is well illustrated by undrained triaxial test results
reported by Shibuya et al. (1997) and Shibuya (2001). Figure 9.17a shows the stress-strain curves
deduced from CIU tests performed on a reconstituted clay (called NSF clay) at strain rates of 0.011,
0.133 and 1.40%/min. It can be seen that, at small strains, smaller than about 0.002%, the modulus is
about the same for the three tests; the most rapid test keeps the same modulus up to strain of about
0.01% whereas the two other tests show a diversion from the linear behaviour at smaller strains; finally,
at strains larger than about 0.01%, the influence of strain rate on stress-strain behaviour is quite clear.
Figure 9.17b shows the relationships between Young’s modulus and strain. It also demonstrates the
strain rate dependency with, in particular, a threshold strain at which the stress-strain relationship stops
being linear (Y1, see Section 4 and Figure 4.1) which decreases with strain rate. From such observa-
tions, Tatsuoka and co-workers developed the model schematized in Figure 9.18. It implies that Y1 and
Y2 surfaces (Figure 4.1) increase in size with increasing strain rate (Tatsuoka et al., 1997), in fact as the
limit state curve (Y3) does.
Vaid & Campanella (1977) performed a variety of triaxial tests on the undisturbed Haney clay. In
particular, they performed undrained compression tests at different strain rates and undrained compres-
sion tests in which the strain rate was changed at a given strain. As shown in Figure 9.19a, when the

134
09031-02[1].qxd 18/Oct/02 12:10 AM Page 135

Figure 9.18. Framework of stress-strain relationship at small strains (Tatsuoka & Shibuya, 1992; Tatsuoka et al.,
2000).

Figure 9.19. Undrained triaxial compression tests on undisturbed Haney clay consolidated to c  505 kPa;
(a) influence of strain rate; (b) influence of undrained creep (after Vaid & Campanella, 1997).

strain rate was increased from 3.5  103 to 4.7  102%/min, the stress-strain curve moves from the
curve corresponding to the former strain rate to that corresponding to the new strain rate. The transition
corresponds to a high-stiffness zone. Based on such results in which the strain rate could increase or
decrease, and creep tests, Vaid & Campanella (1977) proposed the following model in which the devi-
atoric stress is a function of strain and strain rate:

q  f( s, ·s) (9.8)

Tatsuoka et al. (2000) show many other examples of shear tests performed on soft and stiff clays,
silty sand, sand, mudstone and gravel that generally confirm Equation 9.8. This latter equation thus

135
09031-02[1].qxd 18/Oct/02 12:10 AM Page 136

εvol 0.1 ε
εs

σv' or p '
q vol
0.01 ε vol
0.1 εs
C
0.01 εs Creep D
A
Creep B

(a) (b)

εs εvol

Figure 9.20. Influence of strain rate on the stress-strain behaviour of soilds: (a) shear behaviour; (b) compression
behaviour.

appears very general. Also clear from these results is the fact that when shear is started after some creep
or when the strain rate is suddenly increased, the soil initially shows high stiffness.
The behaviour described in Section 9.1 for oedometer tests refering to axial effective stress, strain
and strain rate can certainly be extended to mean stress, volumetric strain and volumetric strain rate as
follows:
p  f( , · ) vol vol (9.9)

Equations 9.8 and 9.9 are very similar and probably are two facets of the same viscous phenomenon.
They are schematized in Figures 9.20a and b respectively.
In some cases, the stress-strain curve obtained after a creep phase or when increasing the strain rate
may overshoot the reference curve corresponding to the considered strain rate, as indicated by the
dashed lines shown in Figures 9.20a and b. Figure 9.19b shows such an example where the stress-strain
curve obtained after undrained creep clearly overshoots the reference stress-strain curve. These
phenomena are associated with structuring phenomena and are described in Section 11.

9.4 Critical state line


Tests performed at different rates of strain on normally consolidated samples of clay do not show any
significant effect of strain rate on the normally consolidated friction angle when determined at
( 1/ 3)max (Bjerrum et al., 1958; Vaid & Campanella, 1977; Badra-Blanchet, 1981; Lefebvre &
Leboeuf, 1987; Sheahan et al., 1996). Tests performed on the Erksak sand by Been et al. (1991) indicate
no significant effect of strain rate on the critical state strength either. The critical state line thus does not
seem to be significantly influenced by strain rate in the q–p diagram.

9.5 Creep
On the basis of creep tests performed in the triaxial apparatus in drained and undrained conditions,
Singh & Mitchell (1968) showed that: (a) the logarithm of strain rate decreases with the logarithm of
time; (b) the slope of the log ·a–log t relationship is essentially independent of the applied stress;
(c) an increase in applied deviatoric stress shifts the relationship to higher strain rates; and (d) except at
very low and very high stress levels, there is a linear relationship between the logarithm of the strain rate
and the deviatoric stress at a given time. Singh & Mitchell (1968) proposed the following equation:
m
t 
 a  A ⋅ eq ⋅  1  (9.10)
 t

where a is the axial strain rate at any time t, q is the stress level equal to the applied deviatoric stress
divided by the deviatoric stress at failure, m is the slope of the log ·a–log t curve which generally falls

136
09031-02[1].qxd 18/Oct/02 12:10 AM Page 137

4-5 4
20 D 3

(σ 'a - σ'r )/2 (kPa)


5 F 2-3 2
G E C
10 B
A
Limit state curve
-1 1
0
10
Saint-Alban, 3 m
0
-10
0 10 20 30 40 50
(σ' + σ' )/2 (kPa)
a r
F D

-2
10
G
Axial strain rate εa (%/min.)

-3
10

-4
10

A σ' = 25.0 ;σ' = 16.5


1 3
-5
10 B σ' = 34.0 ;σ' = 16.5
1 3
C σ'1 = 41.0 ;σ' 3 = 16.5
D σ' = 52.8 ;σ' = 16.5
1 3
E σ' = 33.1 ;σ' = 10.3
1 3
F σ'1 = 33.9 ;σ'3 = 3.1
G σ'1 = 25.2 ;σ'3 = 2.0
B E
-6 A
10
2 3 4 5
1 10 10 10 10 10
Time t (min.)

Figure 9.21. Axial strain rate-time relationship for creep tests on the Saint-Alban clay (after Tavenas et al., 1978).

between 0.7 and 1.0, t1 is a reference time generally taken equal to 1 min, and  and A are creep param-
eters. Murayama et al. (1984) found a similar behaviour for Toyoura sand, with a m value of 1.0.
For the creep tests performed under undrained conditions at high stress levels, the strain rate reaches
a minimum value after some time and then increases up to failure of the soil specimen. Vaid &
Campanella (1977) showed on Haney clay that the relationship between this minimum strain rate and
the applied deviatoric stress is the same as that obtained in undrained shear between the strain rate used
and the deviatoric stress at the peak.
The studies on creep previously mentioned were mostly performed on normally consolidated
materials. A few studies were performed on overconsolidated clays, thus inside the limit state curve
(Bishop & Lovenbury, 1969; Larsson, 1977; Tavenas et al., 1978; d’Elia, 1991). Typical creep test
results obtained by Tavenas et al. (1978) on Saint-Alban clay are shown in Figure 9.21. They show a lin-
ear decrease of the logarithm of strain rate with logarithm of time and thus extend the validity of

137
09031-02[1].qxd 18/Oct/02 12:11 AM Page 138

Equation 9.10 to the overconsolidated domain. Tavenas et al. (1978) considered, however, that the stress
function (exp(q)) used in Equation 9.10 should be generalized and written as n( ), with n( ) being
a stress function reflecting the shape of the limit state curve of the soil. Equation 9.10 then becomes:

m
t 
  A ⋅ n( ) ⋅  1  (9.11)
 t

Equation 9.11 would be valid for both volumetric and shear strains. For high stress levels, Equations
9.10 and 9.11 are no longer valid after some time. As shown on Figure 9.21 for tests D, F, and G, the rate
of strain reaches a minimum value and then increases to reach failure. From the data presented by
Lefebvre (1981), the minimum strain rate is reached at a strain approximately equal to that obtained at
failure in conventional triaxial drained tests, indicating that accumulated strain is a major factor in the
control of failure.
The response of soils in terms of variation with time of volumetric strain in drained tests or pore pres-
sure in undrained tests depends on the overconsolidation ratio of the soil. For example, in Figure 9.21,
the void ratio decreased with time for test C, and increased with time for test G.
As described in this section, the viscous behaviour of soils would be time dependent whereas previ-
ous results indicated that soil behaviour is strain rate dependent. In order to clarify this aspect, Leroueil
(1998, 2001) established a simple model and simulated creep tests. The hypotheses considered were:
a hyperbolic stress-strain curve; a log-log variation of shear strength with time; and failure controlled
by a critical accumulated strain, 0.9% in that case. The model, with the considered equations and param-
eters, is shown in Figure 9.22a. Creep tests have been simulated for the deviatoric stresses indicated by
arrows in Figure 9.22a, and the results are shown in Figure 9.22b. It can be seen that the plots of loga-
rithm of strain rate against logarithm of time lines are linear, as observed in laboratory tests (Figure 9.21),
and as indicated by Equations 9.10 and 9.11. As indicated by Leroueil (2001), this means that, funda-
mentally, soil behaviour is strain rate dependent and that the time model proposed by Singh & Mitchell
(1968) reflects both this strain rate dependency and the test itself, in which stress conditions for creep
tests are suddenly applied at time zero. It is worth noting that the strain rate model shows a slight
decrease of the slope, m, of the log  against log t curves, as observed in laboratory creep tests.
Behaviour similar to that described for soft clays was also observed for stiff clays (D’Elia et al.,
1998) and soft rocks (Nishi et al., 1983; Adachi et al., 1998).

9.6 Some other implications of the viscous behaviour of soils


The viscous behaviour previously described has many possible consequences on soil behaviour. Some
are described in the following paragraphs.

Figure 9.22. Simulation of creep tests: (a) strain rate dependency of the stress-strain behaviour and input parame-
ters; (b) simulated creep tests for the conditions given in (a) (from Leroueil, 1998, 2001).

138
09031-02[1].qxd 18/Oct/02 12:11 AM Page 139

9.6.1 Undrained creep under constant deviatoric stress


Undrained creep of normally consolidated soils, under deviatoric stress or under isotropic stress condi-
tions, is accompanied by an increase in pore pressure and thus by a decrease in effective stress
(Arulanandan et al., 1971; Holzer et al., 1973). Arulanandan et al. (1971) carried out undrained creep
tests at different stress levels on San Francisco Bay mud. The effective stress conditions at different
times are shown in Figure 9.23. It can be seen that the locus of effective stress conditions at a given time
maintains essentially the same shape, but progressively move towards smaller stresses as time increases.
This behaviour can be explained in the same manner as for relaxation tests in one-dimensional condi-
tions (see Figure 9.8 and Equations 9.6 and 9.7).
As indicated, in particular by Sekiguchi (1977) and Sheahan (1995), there must be a lower limit to
stress conditions which can be reached during creep tests. This is supported by observations showing
that highly overconsolidated clays have a tendency to generate negative pore pressures in undrained
creep tests, or swelling in drained creep tests (see Figure 9.10). Sheahan (1995) calls the lower limit
“static yield surface”; “inviscid yield surface” has also been used in the literature.

9.6.2 Undrained relaxation (constant a) in triaxial conditions


These tests are characterized by the fact that the strain, the axial strain in triaxial tests, remains constant
with time. Such tests have been performed on clays and sands by Lacerda & Houston (1973), Akai et al.
(1975), Lacerda (1976), Hicher (1988) and Sheahan et al. (1994). The results show a linear decrease of
the deviatoric stress with the logarithm of time; they also show that pore pressure does not vary signif-
icantly during relaxation tests. This implies, as demonstrated in Figure 9.24, that the effective radial
stress remains approximately constant.

9.6.3 Undrained triaxial tests performed at different strain rates


If undrained triaxial tests are performed on normally consolidated clays at different strain rates, the gen-
erated pore pressure increases as the strain rate decreases. As a result, a rapid test will reach failure at a
point such as Rnc in Figure 9.25 whereas a slow test will reach failure at Snc. As the critical state strength
envelope is strain rate independent (see Section 9.4), Rnc and Snc are on the same strength envelope and
the difference in strength is due to different pore pressures at failure (Figure 9.25). For undrained triax-
ial tests performed in the overconsolidated domain, the strength results from the pore pressure gener-
ated during creep and a strength envelope that may be strain rate dependent (Figure 9.25). These effects
result in a decrease in undrained shear strength of natural clays with decreasing strain rates. Compiling
test results obtained on 26 different clays, Kulhawy & Mayne (1990) found a typical change in
undrained shear strength of 10% per logarithm cycle of strain rate (Figure 9.26). Graham et al. (1983a)
found similar variations in triaxial compression, but also in triaxial extension and direct simple shear
tests. However, studies performed by Hight (1983), Hight et al. (1987) and Sheahan et al. (1996) on
reconstituted Lower Cromer till and Boston Blue clay indicate that the effect of strain rate on the com-
pression undrained shear strength considerably decreases when the overconsolidation ratio increases.

Figure 9.23. Change in effective stress with time during undrained triaxial creep tests (( 1  3)  cst) (after
Arulanandan et al., 1971).

139
09031-02[1].qxd 18/Oct/02 12:11 AM Page 140

0.6

SR1-8
σ '3 ) / 2 σ'vm 0.4 SR1-9
SR1-5
( σ '1

0.2

0
0 0.2 0.4 0.6 0.8 1.0 1.2
' + σ '3) / 2 σ'
( σ1 vm

Figure 9.24. Effective stress paths followed during triaxial relaxation tests ( 1  cst) on normally consolidated San
Francisco Bay Mud (data from Lacerda, 1976; after Martins, 1992).

Undrained tests
(s 'a - s r') /2

Rapid e
Slow lin
f 'nc
Y: limit state curve at R nc
R oc
constant void ratio
Snc
YR Soc

YS I nc

I oc

(s a' + s r') /2

Figure 9.25. Undrained compression tests – Schematic behaviour.

Figure 9.26. Influence of strain rate on the undrained shear strength measured in triaxial compression (from
Kulhawy & Mayne, 1990).

140
09031-02[1].qxd 18/Oct/02 12:11 AM Page 141

9.6.4 Secondary consolidation and subsequent undrained compression tests


The influence of viscous effects on consolidation and subsequent shear testing is shown in Figure 9.27
for soil specimens consolidated at point A, in the normally consolidated domain (Figure 9.27a), and
subjected to undrained compression. If the specimen is tested immediately after reaching the end-of-
primary consolidation, the stress path is as AD in Figure 9.27a and corresponds to AD in Figure 9.27b,
with D on the critical state line of the soil. If, for a second specimen, the drainage valve is closed at the
end of primary consolidation and the soil left under constant total stresses during some time, pore pres-
sure develops (there is creep as shown in Figure 9.23) and the effective stresses move from A to B.
During shearing, the soil specimen first exhibits an “overconsolidated behaviour” with a stress path ini-
tially at constant p (assuming a perfectly linear elastic behaviour), and then moves towards D. Since the
void ratio is the same as for the first test, the shear strength is the same for both specimens. If, for a third
specimen, drainage is allowed for some time after the end-of-primary consolidation, the void ratio
decreases and soil conditions progressively move from A to C in Figure 9.27b. The new limit state
curve for the soil is then RN (Figure 9.27a) with its projection in the e–p diagram being RN. At point
C, the soil is overconsolidated and behaves as such during the initial part of the undrained shear test
(constant p); then, it progressively goes to failure, at point E (Figure 9.27a) and E (Figure 9.27b) cor-
responding to its void ratio. This was demonstrated by Shen et al. (1973).

Figure 9.27. Effect of time on consolidation and subsequent undrained compression (from Leroueil & Marques,
1996).

141
09031-02[1].qxd 18/Oct/02 12:11 AM Page 142

9.7 K0
As seen in Section 16.7, the coefficient of earth pressure at rest for normally consolidated soils,
Konc  h/ v, can be described by Equation 16.9 (also see Figure 16.6). One question raised by
Schmertmann (1983) was how, if at all, K0 varies with time during secondary consolidation or aging. In
fact, experimental data presented in the literature clearly show that K0 increases with time in normally
or slightly overconsolidated clays (Tavenas & Leroueil, 1987). Mesri & Hayat (1993) compiled K0/log t
values measured during secondary consolidation and found that they range between 0.003 and 0.05.

9.8 Residual strength


The residual strength envelope is characterized by a friction angle r and a cohesion equal to zero
(Section 8.4). Petley (1966), Kenney (1967), Lupini (1980) and Lemos (1986) studied the effect of the
rate of shearing on the residual friction angle of clays. The effect of rate of displacement is small, gen-
erally smaller than 3% per logarithm cycle for rates smaller than 101 mm/min. At faster rates (larger
than 1–10 mm/min), the residual strength under a constant normal stress has been observed to remain
constant or to increase or to decrease, depending on the soil and testing conditions. Tika et al. (1996)
studied these phenomena, and concluded that decrease in residual strength at fast rates would be due to
an increase in void ratio in the shear zone, and would exist if free water is available. For sands, Hungr &
Morgenstern (1984) did not find any rate effect.

10 INFLUENCE OF TEMPERATURE

Temperature has two major effects on soils: thermal expansion of solid particles and pore water, and
thermally induced modification of the strength of contacts between particles or aggregates. The com-
bined effect of these phenomena on the temperature-volume-effective stress behaviour of soils in both
drained and undrained conditions is explained by Mitchell (1993) and will not be described here in
detail. Figure 10.1, deduced from tests on illite samples initially isotropically consolidated under
200 kPa, illustrates the behaviour generally observed. In drained conditions, an increase in temperature
is associated with an expulsion of pore water from the clay sample whereas a decrease in temperature is
associated with absorption. As explained by Mitchell (1993), “the weakening effect of the higher tem-
perature is compensated by the strengthening effect of the lower void ratio”. However, due to changes
in the particle arrangement, the process is not reversible for soils normally or slightly overconsolidated.
In undrained conditions, heating is associated with an increase in pore pressure, and consequently, with

Effective stress (kPa)


100 125 150 175 200 225
Undrained
(Percent of initial sample volume)

46 32 19
0
59 46 31 I 19
Pore water drained

0.5 32
drained

1.0 19
46
32
1.5 46
19: temperature in ˚C
60
2.0

Figure 10.1. Effects of temperature changes on saturated illite under drained and undrained conditions (initial
isotropic effective stress  200 kPa) (after Mitchell, 1993).

142
09031-02[1].qxd 18/Oct/02 12:11 AM Page 143

a decrease in effective stress. In these conditions where there is no significant particle rearrangement,
the process is almost reversible.
The results presented and discussed in this Section are associated with temperatures from above zero
to 100°C, i.e. temperatures that are not low enough to modify soil fabric by the effect of frost and not
high enough to produce changes in mineralogy.

10.1 One-dimensional compression


In recent years, many studies have been conducted on the effects of temperature on the compressibility
of natural soft clays (Eriksson, 1989; Tidfors & Sällfors, 1989; Boudali et al., 1994; Marques, 1996;
Graham et al., 2001; Marques et al., 2002). Typical CRS test results, obtained on sulphide clay from
Lulea, Sweden, are shown in Figure 10.2. It can be seen that there is a significant effect of temperature
on the compressibility of the clay. With increasing temperature, the soil becomes more compressible in
the overconsolidated range, the preconsolidation pressure decreases, and the entire compression curve
moves towards smaller effective stresses. It can be seen on Figure 10.2b that the effect is more impor-
tant at temperatures below 35°C than at higher temperatures. The two CRS tests A and B performed on
the St-Polycarpe clay (Figure 9.3b) with changes in temperature at different strains from 5°C to 20°C
in Test A and from 50°C to 20°C in Test B, show the effective stress-strain curves jumping from one

Vertical effective stress (kPa)


10 100
0

5
Axial strain (%)

10

15

20

25
T( 5
°C)
45

15
35
25

30
(a)
Preconsolidation pressure (kPa)

80

70

60

50

40

30

20
0 10 20 30 40 50 60
Temperature (°C)
(b)

Figure 10.2. Oedometer tests performed on Luleå clay at various temperatures (from Eriksson, 1989).

143
09031-02[1].qxd 18/Oct/02 12:11 AM Page 144

constant temperature curve to another, and thus confirm the influence of temperature on one-
dimensional compression.
Boudali et al. (1994) performed CRS tests at different strain rates and temperatures on 3 clays,
among which was the Berthierville clay previously discussed (Figure 9.4). Their results show the fol-
lowing: (a) the preconsolidation pressure can be described as a function of strain rate and temperature
(Figure 9.4a); and (b) the effective stress-strain curves obtained at various strain rates and temperatures
reduce to a unique one when normalized with respect to the preconsolidation pressure corresponding to
the strain rate and temperature used in each test (Figure 9.4b). The model proposed by Leroueil et al.
(1985a) for strain rate effects can thus be extended for including temperature effects. Equations 9.1 and
9.2 then become:

p  f( 1, T) (10.1)

and

1  p( 1, T)  g( 1) (10.2)

The test results shown in Figures 9.3 and 10.2 fit this model, which also implies that the compression
index (or ) is independent of strain rate and temperature.
Figure 10.3 shows the preconsolidation pressure (or the effective stress at a given void ratio) nor-
malized with respect to the preconsolidation pressure (or the effective stress at the same void ratio)
measured at 20°C, as a function of temperature. The change in preconsolidation pressure with temper-
ature is about the same for all the clays considered, indicating a change in preconsolidation pressure of
almost 1% per °C between 5°C and 40°C, and much smaller at larger temperatures.
Equations 10.1 and 10.2 indicate that strain rate and temperature are two factors which are intimately
linked in the viscous deformation of soils and influence the maximum effective stress a soil can support
at a given void ratio. This results from fundamental physics and can be explained, at least qualitatively,
by the rate-process theory (Murayama & Shibata, 1961; Mitchell, 1964, 1993).

10.2 Limit state curve


To investigate the effect of temperature on the limit state curve, Boudali (1995) performed, in addition
to the oedometer tests previously described (Figure 9.4), isotropic compression tests and undrained tri-
axial tests at temperatures of 5, 20 and 35°C on Berthierville clay. The isotropic compression tests show
a behaviour very similar to that observed in oedometer tests with an isotropic yield stress decreasing
with increasing temperature (Figure 10.4). Burghignoli & Desideri (1988) on reconstituted Todi clay,

1.40
σ' p/σ ' p (T = 20 ˚C) or σ'1 /σ '1 (T = 20 °C)

Boudali et al., 1994


Despax, 1975
Campanella & Mitchell, 1968
1.20 Tidfors & Sällfors, 1989
Eriksson, 1989
Moritz, 1995a and b
Marques, 1996
1.00 Akagi & Komiya, 1995

0.80

0.60
0 20 40 60 80 100
Temperature (°C)

Figure 10.3. Variation of the normalized preconsolidation pressure, or vertical effective stress at a given void ratio,
with temperature (from Leroueil & Marques, 1996).

144
09031-02[1].qxd 18/Oct/02 12:11 AM Page 145

Graham et al. (2001) on illite and Marques et al. (2002) on St-Roch-de-l’Achigan clay observed simi-
lar effects of temperature. The isotropic yield stresses and preconsolidation pressures obtained on
Berthierville clay at a strain rate of 107 s1 are presented in Figure 9.13b (the preconsolidation pres-
sures deduced from Figure 9.4 have been plotted, in an approximative manner, on the K  0.5 line).
Even if the data available are limited, it seems that the effect of temperature on the limit state curve is
very similar to the effect of strain rate (Figure 9.13a).
Due to such effects of temperature, for a soil consolidated in the normally consolidated range at a
given temperature, cooling causes the soil to behave as if it were overconsolidated (Plum & Esrig, 1969;
Burghignoli & Desideri, 1988; Hueckel & Baldi, 1990). As a result, volumetric strains due to tempera-
ture changes in the overconsolidated range are relatively small and to a large extent reversible (Hueckel &
Baldi, 1990; Burghignoli et al., 1992; Boudali et al., 1994). Plum & Esrig (1969), Hueckel & Baldi
(1990) and Burghignoli et al. (1992) also showed that these volume changes depend on the overconsol-
idation ratio. For the natural Pasquasia clay (Figure 10.5) and for temperatures increasing from 21°C to

σ' (kPa)
0 20 40 60 80
0

T (˚C)
2 5
20
4 20
35
Volumetric strain (%)

10

12

14

16

Figure 10.4. Effective stress-volumetric strain relations at a strain rate of 107 s1 and at different temperatures,
isotropic compression isotropic compression of Berthierville clay (from Boudali, 1995).

100 OCR = 3.4


13.07 7.5 5.67

80
Temperature (˚C)

60

40

20
-0.3 -0.2 -0.1 0 0.1 0.2 0.3
Volumetric strain (%)

Figure 10.5. Drained heating tests at constant effective stress at various OCRs (in terms of isotropic stress). Natural
Pasquasia clay (from Hueckel & Baldi, 1990).

145
09031-02[1].qxd 18/Oct/02 12:11 AM Page 146

Figure 10.6. Limit state curves of Saint-Roch-de-l’Achigan clay (4.8–5.8 m) at temperatures of 10, 20 and 50°C
(after Marques et al., 2002).

more than 90°C, volumetric strains are positive for OCRs smaller than approximately 6, and negative
for higher OCRs. The similarity with the tendency for the soil to creep in compression when slightly
overconsolidated and in swelling when strongly overconsolidated (see Figure 9.10) is noteworthy. So, as
Feijo & Martins (1993) suggested for creep (see Section 9.1), a “neutral OCR” could also be thought of
for temperature.
Hueckel & Baldi (1990), Boudali (1995), Moritz (1995) and Marques et al. (2002) observed a gen-
eral tendency for the peak shear strength to decrease when temperature increases. However, as noted by
Hueckel & Baldi (1990), when temperature increase is associated with a significant void ratio decrease,
the strength gain due to this decrease in void ratio could compensate for the strength loss due to a higher
temperature.
Figure 10.6 summarizes test results obtained by Marques et al. (2002) on St-Roch-de-l’Achigan clay
from Québec. Below the critical state line, the effect of temperature is very clear. As for peak strength
envelopes, that at 10°C is the highest; those at 20°C and 50°C do not show a clear effect of temperature
but that could be due, as suggested by Hueckel & Baldi (1990), to a combined result of weakening of
the interparticle contacts and strengthening of the soil due to a decrease in void ratio when the soil is
heated from 20°C to 50°C.
Graham et al. (2001) performed a variety of triaxial tests on an isotropically consolidated illite at
temperatures of 28, 65 and 100°C. Figure 10.7 shows the peak strength envelope obtained in the over-
consolidated range, normalized with respect to pe (T), the mean stress on the normal consolidation line
corresponding to the test temperature, at the considered void ratio. There is some scatter but normal-
ization seems to be a reasonable approach.
Equation 10.1 indicates that the preconsolidation pressure is a function of both strain rate and tem-
perature. The results presented in Section 9.2 and here show that it can be generalized to the entire limit
state curve and surface. As schematized in Figure 10.8 in a e–p–q diagram, the limit state surface
appears to have different peels, each one being associated with different strain rate and temperature. The
results also indicate that the mechanical behaviour could be normalized with respect to p or pe, value
that would be a function of strain rate and temperature.

10.3 Small strain behaviour


Anderson & Richart (1974) performed low amplitude resonant column tests on cohesive soils at tem-
peratures of 22°C and 4°C. Seven soils with plasticity index varying from 12 to 64 were tested; five of
these soils were undisturbed and two were reconstituted. Anderson & Richart (1974) found shear wave
velocities at 4°C from 0% to 12.5% higher than at 22°C and concluded that temperature has a minor
influence on shear wave velocity. This is consistent with the very small effect of strain rate on small
strain shear modulus (see Section 9.3). However, at larger strains (0.1%), as for strain rate, there is
evidence that moduli decrease when temperature increases (Eriksson, 1989; Hueckel & Baldi, 1990).

146
09031-02[1].qxd 18/Oct/02 12:11 AM Page 147

Figure 10.7. Normalized peak strength envelope for overconsolidated specimens of illite tested at different tem-
peratures (modified after Graham et al., 2001).

Figure 10.8. Influence of strain rate and temperature on the limit state surface of soils.

10.4 Critical state line


Test results presented by Mitchell (1964), Hueckel & Pellegrini (1989), Hueckel & Baldi (1990),
Graham et al. (2001) and Marques et al. (2002) do not indicate a significant effect of temperature on the
friction angle of normally consolidated clays. Houston et al. (1985) observed a slight increase (1°)
between 4 and 40°C and a much greater increase at higher temperatures. Testing a sand-bentonite mix-
ture, Lingnau et al. (1995) found very similar strength envelopes at 26°C and 65°C, and a strength enve-
lope lying well above at 100°C. Hence, it seems that the strength envelope of normally consolidated
clays, thus the critical state line in the stress diagram, is essentially temperature independent, at least for
temperatures lower than about 50°C.
Sherif & Burrous (1969) report test results in which specimens were consolidated at room tempera-
ture and then sheared in undrained, unconfined compression tests at the same or higher temperature.
The results presented in Figure 10.9 demonstrate the influence of water content or void ratio; they also

147
09031-02[1].qxd 18/Oct/02 12:11 AM Page 148

44

42

Tr
ia
Moisture content (%)
40

xia
col
n so
lid
38

a
tio
n
cu
rv
e
36

34

7 5 0 ˚ ˚F
1 02 5 ˚F
1 50

˚F F
1
32
8 10 20 40 60 80 100
Compression strength (psi)

Figure 10.9. Effect of temperature on the undrained strength of kaolinite in unconfined compression (from Sherif
& Burrous, 1969).

confirm that, at a given water content, the higher the temperature, the smaller the undrained shear
strength. As the strength envelope does not change significantly at the considered temperatures, the
results also indicate that, at a given water content, the p value at the critical state decreases when tem-
perature increases. The data presented by Lingnau et al. (1995), Graham et al. (2001) and Marques et al.
(2002) confirm that point. The general behaviour of soils subjected to different temperatures can thus
be described, as shown in Figure 10.10 with: (a) a unique critical state line in the q–p diagram (Figure
10.10a); and (b) normal compression and critical state lines in an e (or w)–log p diagram that are tem-
perature-dependent.
Figure 10.10 schematically shows the behaviour of clay specimens that would be normally consoli-
dated and then subjected to undrained compression at different temperatures. The specimens are first
consolidated to point A, at temperature T1. At the end of primary consolidation, they are at point A, on
the normally consolidated line NCL1 corresponding to T1 (Figure 10.10b). A specimen sheared in
undrained conditions at this temperature reaches failure at point D (Figure 10.10a) corresponding to
point D on the critical state line CSL1 corresponding to T1 (Figure 10.10b). If, for a second specimen,
the valve is closed and the temperature increased to T2, the pore pressure increases and, due to the differ-
ence in thermal expansion of water and soil particles, the void ratio increases slightly. The soil conditions
then move to point B on the normally consolidated line NCL2 corresponding to the new temperature.
If the soil is sheared at T2, the soil conditions progressively move to point E, on the critical state line
CSL2 corresponding to temperature T2 (Figure 10.10b). In the stress diagram (Figure 10.10a), the corre-
sponding point E is on the same strength envelope as the first specimen since temperature has no influ-
ence on the friction angle of the normally consolidated soil, but at a lower undrained shear strength. For
a third specimen heated in drained conditions to T2, there is a decrease in void ratio from A to C on
NCL2 (Figure 10.10b). During undrained compression at T2, soil conditions move to F on CSL2.

10.5 Residual strength


Butcher (1975) observed on two different clays that the residual friction angle was not influenced by
temperature changes between 10 and 60°C. None or a small effect of temperature on r is consistent
with the observations made for the strength envelope of the normally consolidated soil (Section 10.4)
and the fact that r is only slightly influenced by strain rate (Section 9.8).

148
09031-02[2].qxd 18/Oct/02 12:12 AM Page 149

L
CS
D Stress paths
F
in CIU tests
E

B A
P'
(a) C

e 1 for T1
2 for T2
with T1 < T2

B'
E' A'
D'

F' NC
C' L
NC 1
L
CS 2
CS

L
L

1
2

(b) P'

Figure 10.10. Effect of temperature change on consolidation and subsequent undrained compression.

11 INFLUENCE OF MICROSTRUCTURE

11.1 Microstructure of natural soils and soft rocks


As indicated by many authors, in particular by Burland (1990), Leroueil & Vaughan (1990),
Schmertmann (1991), Clayton & Serratrice (1993) and Kavvadas (2000) in their synthesis papers, most
natural soils and soft rocks are microstructured (bonded), i.e. at a given void ratio, they can sustain
stresses higher than could the same material non-microstructured. This is evident for soft rocks; it is
now recognized for most soft clays (Tavenas & Leroueil, 1985; Burland, 1990); there is also evidence
of microstructure in sands (Mitchell & Solymar, 1984; Schmertmann, 1991) as well as in stiff overcon-
solidated clays and clay shales (Calabresi & Scarpelli, 1985; Picarelli et al., 2002*). Leroueil &
Vaughan (1990) concluded that microstructure is as important as initial void ratio and stress history in
determining the behaviour of natural geomaterials. This has been confirmed since that time and specific
aspects of microstructured soils have been studied in detail.
There are several causes for the development of microstructure in soils and soft rocks: secondary
compression, thixotropy, cementation, cold welding, etc. As indicated in Section 9, secondary com-
pression only gives a strength increase related to the decrease in void ratio. The other factors give the
soil a strength which cannot be accounted for by void ratio and stress history alone.

149
09031-02[2].qxd 18/Oct/02 12:12 AM Page 150

Ideal soil Structured soil

Log p' Log p'


e

e
Yield Yield

R R L
L Y
~ elastic ~ elastic

~ p la sti
~p
la s

c
tic

(a) (b)
σ'a - σ' r
σ'a - σ' r

Limit state e
φ ' lin
2
2

e or yield curve
Limit state φ' lin
u re
or yield curve Fa il ~ plastic

u re
Fa il C ~ plastic C
~ elastic ~ elastic
σ' a + σ'r σ' a+ σ 'r
(c) 2 (d) 2

Figure 11.1. Schematic behaviour of ideal and structured soils (from Leroueil, 1992).

Figure 11.1 summarizes the main differences in mechanical behaviour between microstructured soils
and “ideal soils” as described by CSSM (Section 2). Compared to the same soil non-microstructured
and at the same void ratio, the microstructured soil has higher preconsolidation pressure, undrained
shear strength and stiffness. These features are detailed in Section 11.2; Section 11.3 describes what
happens if the bonds between particles or aggregates are broken and the soil is “destructured”

11.2 Main features of microstructured soils and soft rocks


11.2.1 One-dimensional compression
In one-dimensional compression, microstructure can be shown by comparison of the compression
curves obtained on the natural, undisturbed, soil or soft rock and on the same geomaterial, remoulded
and reconstituted. This can be illustrated by test results obtained on Mexico City and the Grande
Baleine soft clays (Figure 11.2), hard Pappadai clay (Figure 11.3) and calcarenite (Figure 3.11a).
Because of microstructure, these soils can reach a domain of the e–log v space that is not permitted
for the non-microstructured soil; the apparent preconsolidation pressure of the soil is higher than the
value normally associated with its void ratio. For the case of Pappadai clay (Cottechia & Chandler,
1997; Cotecchia, 2002*), the vertical effective stress at the depth of sampling was 415 kPa while the
geological preconsolidation stress was about 1300 kPa (Point P on Figure 11.3). It can be seen that P lies
on the right of the compression curve of the reconstituted clay, indicating that the soil was already struc-
tured at the time of the formation of the deposit. The vertical effective gross yield stress deduced from

150
09031-02[2].qxd 18/Oct/02 12:12 AM Page 151

Figure 11.2. One dimensional compression curves.

1.3
Reconstituted clay

1.1
Preconsolidation
P Yield
Void ratio e

0.9

0.7
Natural
clay
0.5

0.3
2 3 4
10 10 10 10
σ v' (kPa)
Figure 11.3. One-dimensional compression behaviour of the natural and reconstituted Pappadai clay (after
Cotecchia & Chandler, 1997; Cotecchia, 2002*).

oedometer tests is about 2600 kPa, twice the geological preconsolidation pressure, confirming the
strong microstructure of Pappadai clay. In dense materials, however, the influence of microstructure on
gross yielding may be difficult to observe as indicated by Aversa et al. (1993) and Huang & Airey
(1993).
Burland (1990) showed that the void ratio-effective overburden stress relationship for normally con-
solidated deposits (sedimentation compression line) is different from the compression line obtained in
the laboratory with reconstituted clay (intrinsic compression line, ICL; see Section 16.3). As shown in
Figure 18.2, under a given effective stress, the natural void ratio is higher than that of the reconstituted
material.
Also associated with microstructure is the fact that the compression index Cc is generally larger than
that of the same soil reconstituted. This aspect is discussed in Section 18.4.

151
09031-02[2].qxd 18/Oct/02 12:12 AM Page 152

0.8
HP tests
MP tests

0.6

q /p 'e

0.4

0.2

0
0 0.4 0.6 0.8
p '/p 'e

Figure 11.4. Normalized stress paths and strength envelopes of Vallericca clay undisturbed (MP tests) and
destructured (HP tests) (from Amorosi & Rampello, 1998).

11.2.2 Limit state curve and peak strength envelope


The peak strength envelope, and in particular the effective cohesion, of soils in their overconsolidated
domain reflects their microstructure. Figure 11.4 shows the strengths envelopes of stiff Vallerica clay, as
deduced from two series of tests performed by Amorosi & Rampello (1998). In both series, the clay was
first anisotropically consolidated to a given vertical effective stress and then unloaded in approximately
Ko conditions before shearing in drained or undrained conditions. In the first series (MP tests), the max-
imum vertical stress reached during the tests was smaller than the vertical gross yield stress evaluated
from oedometer tests and the soil remained undisturbed. In the second series (HP tests), the soil was
first loaded to stresses well in excess of the vertical gross yield stress, and the soil was then destruc-
tured. In Figure 11.4, the strength envelopes are normalized with respect to pe to take into account the
influence of void ratio. It can be seen that the strength envelope of the undisturbed material (MP tests)
is well above the strength envelope of the destructured soil (HP test).
Burland et al. (1996) compared the strength envelopes of four different soils in intact and reconsti-
tuted conditions and clearly showed that, at the same void ratio, the strength envelope of the intact soil
is above that of the reconstituted soil (see Section 18.6 and Figure 18.7).

11.2.3 Small strain shear modulus


Microstructure reinforces the links between particles, and so increases the small strain shear modulus,
even at the same void ratio. This has been shown for clays (Rampello & Viggiani, 2001), soft rocks
(Nishi et al., 1989), sands (Katayatama et al., 1986; Hatanaka & Uchida, 1995) and gravels (Goto et al.,
1992). Examples are presented in Figures 11.5 and 11.6. In both cases, the cemented soil has higher
shear modulus than the uncemented one under the same confining stress. Also, it can be seen in Figure
11.5 that, as the cement content increases, the small strain shear modulus increases and the influence of
confining stress decreases. There are other indications in the literature that, in microstructured soils, Go
does not depend only on stresses as suggested by Equation 4.10 but also on the strength of the bonds
(Nishi et al., 1989; Rampello & Silvestri, 1993; Stokoe et al., 1994; Tatsuoka et al., 1997; D’onofrio
et al., 1998). Hardin (1978) suggested that for clays, the shear modulus at very small strains, Go,
depends on applied stresses, void ratio and overconsolidation ratio (OCR). Viggiani & Atkinson
(1995a) showed that Go could be defined as follows for isotropically consolidated reconstituted clays:
n m
Go  p   pp 
 A    (11.1)
pr  pr   pr 

152
09031-02[2].qxd 18/Oct/02 12:12 AM Page 153

Figure 11.5. Dependence of Gmax in sand on confining stress and cementing (unknown).

Figure 11.6. Effect of cementation of a sand and of unloading-reloading on shear wave velocity (after Fernandez &
Santamarina, 2001).

153
09031-02[2].qxd 18/Oct/02 12:12 AM Page 154

where p is the mean effective stress, pr is a reference pressure, pp is the stress at the intersection of
the swelling line with the isotropic normal consolidation line and A, n and m are non-dimensional
soil parameters. The ratio (pp /p) is thus an overconsolidation ratio defined in term of isotropic
stress, Ro.
If Go (at pp ) is the shear modulus when p  pp, then:

(nm)
Go  p 
  (11.2)
Go (at pp )  pp 

From the values obtained on reconstituted clays and reported by Viggiani & Atkinson (1995a), (n  m)
would vary between 0.45 and 0.56, with an average value of 0.51. Thus, Equation 11.2 could be approx-
imated by (Leroueil, 2000):

Go p 1
  (11.3)
Go (at pp ) pp Ro

Equation 11.3 is represented by the dash line in Figure 11.7.


From results presented in the literature and having Equation 11.3 as a reference, Leroueil (2000)
indicated that the influence of microstructure on Go could be summarized as shown in Figure 11.7; the
higher the degree of microstructuration, the less is the influence of applied stresses on Go.

11.2.4 Large deformation strength


A general observation made with microstructured soils is that stress conditions obtained in shear tests
at large deformation are not on the limit state curve, as schematized in Figure 11.1a (Point C), but
inside, as schematized in Figure 11.1b. This is shown for soft clays in Figures 3.2 and 11.8a, fine
grained tuff in Figure 11.8b, hard Pappadai clay in Figure 11.9, and microstructured sands (Cuccovillo
& Coop, 1997). It is due to the fact that the limit state curve reflects the strength of the bonds while, at
large strains, most of these bonds are broken.

Figure 11.7. Schematic figure showing the influence of microstructure on Go (after Leroueil, 2000).

154
09031-02[2].qxd 18/Oct/02 12:12 AM Page 155

Smith et al. (1992) demonstrated the same aspect by plotting radial stress paths starting from in situ
stress conditions in a normalized p/pe–q/pe diagram. As shown in Figure 11.10, the stress paths
travel from in situ stress conditions to gross yielding and then move back. Cuccovillo & Coop (1993)
observed a similar behaviour on Rankin calcarenite; Callisto & Calabresi (1998) also observed it on
Pisa clay but to a less extent, Pisa clay being less sensitive than Bothkennar clay; in the relatively insen-
sitive Queenborugh clay phnomenon is not observed at all (Jardine et al., 2002*).

11.3 Destructuration
By definition, destructuration corresponds to breaking down of the bonds between particles or aggre-
gates and thus to a general decrease in strength of the material. Even if destructuration is a progressive
process, it mostly occurs when a stress path reaches the limit state (or gross yield, Y3 in Figure 4.1)

Large Yield
Peak
120 strain
CIU
CID
100 e
Isotropic in
φ' l
consolidation
(σ' – σ' )/2 (kPa)

80

60
r

Yield curve
a

40 y

Large strain
20 envelope

0
0 20 40 60 80 100 120 140 160 180 200
(σ' + σ' )/2 (kPa)
a r
a) Saint-Vallier clay (after Lefebvre, 1970)
12
Large Yield
Peak
strain
10
CIU in e
φ' l
CID
(σ'a – σ' r )/2 (MPa)

8 Isotropic
consolidation

6
Yield curve

2
Large strain
envelope
0
0 2 4 6 8 10 12 14 16 18 20 22 24
(σ'a + σ' r )/2 (MPa)

b) Fine grained tuff (after Aversa et al., 1991)

Figure 11.8. Effective stress paths in undrained traxial tests.

155
09031-02[2].qxd 18/Oct/02 12:12 AM Page 156

2.8
Undrained shearing
Drained shearing
2.4 Iso. and aniso. compression

L
CS
2.0 Normalized
limit state
q /p 'e 1.6 curve

1.2

0.8

0.4

0
0 0.4 0.8 1.2 1.6 2.0 2.4 2.8 3.2
p'/p 'e

Figure 11.9. Normalized stress paths of the natural Pappadai clay consolidated to states before isotropic yield (after
Cotecchia & Chandler, 1997).

6
Outer state
boundary
surface
e

4
=
1.
8 0 .7 0
e= e=1
1

2 B
.6 0
q/p'e

e = 0.6
Final intrinsic
state boundary
0
surface

Post-yield contours
of void ratio
-2

-4
0 2 4 6 8
p'/p'e

Figure 11.10. Behaviour of Bothkennar clay normalized with respect to pe (from Smith et al., 1992).

156
09031-02[2].qxd 18/Oct/02 12:12 AM Page 157

curve of the material. As indicated by Leroueil & Vaughan (1990), destructuration can be obtained by
compression outside the limit state curve, by shearing or by swelling (Figure 11.11).

11.3.1 Destructuration by compression


Destructuration by compression has been shown for soft clays by Leroueil et al. (1979) and Tavenas &
Leroueil (1985). The effect of removing microstructure was examined by unloading test specimens after
anisotropic consolidation to axial stresses in excess of the preconsolidation pressure and testing them in
anisotropic compression under constant K  r / a ratio or in undrained conditions. The behaviour was
then compared with results obtained from the intact microstructured soil tested after consolidation to
similar overconsolidation ratios. Figure 11.12 shows typical undrained triaxial test results. The destruc-
tured clay is less stiff and reaches a smaller peak strength at a larger failure strain. The resulting failure
envelopes for the destructured clays were thus lower than those for the microstructured clays. As previ-
ously indicated, similar results have also been shown by Amorosi & Rampello (1998) (Figure 11.4).
(σ ' - σ' )/2

e
φ' lin
r

g yie ld
rin
a

a
S he

yield
io n
Swelling yield
(σ 'a + σ'r )/2
ss

re
o mp
Sh C
ea
rin g
yield
φ' li
ne

Figure 11.11. Schematic figure showing different possible zones of yielding (from Leroueil & Vaughan, 1990).

Figure 11.12. Stress-strain relationships from undrained triaxial tests on intact and destructured clays (after
Tavenas & Leroueil, 1985).

157
09031-02[2].qxd 18/Oct/02 12:12 AM Page 158

Figure 11.13. Results of oedometer tests on Laviano clay shale (from Picarelli, 1991b).

Destructuration by compression can also be observed in weak mudrocks (Banks et al., 1975;
Leroueil & Vaughan, 1990) and in clay shales (Picarelli, 1991b). Figure 11.13 illustrates well the phe-
nomenon of progressive destructuration. The Laviano clay shale was cyclically loaded and unloaded in
an oedometer, with maximum stress increasing at each cycle. It can be seen that the swelling index
Cs  e/log  increases with applied stress and accumulated strain to approach the Cs value of the
remoulded and reconstituted soil.
It is worth noting that destructuration by compression can also be reached by drying the soil, as
shown in the heavily overconsolidated Todi clay (Calabresi & Scarpelli, 1985; Rampello, 1989, 1991).
Figure 11.14a shows the strength envelopes obtained for the intact soil and for the soil first air dried and
then left to swell before shearing. The strength envelope obtained after drying is well below the strength
envelope of the intact soil, indicating some destructuration. To take into account the influence of void
ratio on strength envelopes, Rampello (1989) normalized them with respect to pe . The normalized
strength envelope of the undisturbed soil is then only slightly above that of the soil after drying-wetting,
indicating that interparticle bonding is existent but not strong in Todi clay.

11.3.2 Destructuration by shearing


Destructuration may also occur by shearing. This can be illustrated by triaxial tests performed by Hight
et al. (1997) on Crag sand, a very dense, often shelly, marine sand. These authors performed undrained
triaxial compression and extension tests with unload-reload loops at increasing levels of strain. As shown

158
09031-02[2].qxd 18/Oct/02 12:12 AM Page 159

1.0 1.2
Undisturbed
After free swell
0.8 1.0 After drying
Reconstituted o.c.
Reconstituted n.c.
0.8
0.6
q (MPa)

q/pe'
0.6
0.4
Undisturbed 0.4
After free swell
0.2 After drying
Reconstituted 0.2

(a) (b)
0 0
0 0.2 0.4 0.6 0.8 1.0 0 0.2 0.4 0.6 0.8 1.0 1.2
p ' (MPa) p'/p'e

Figure 11.14. Failure envelopes for undisturbed, free swollen and dried-wetted samples of Todi clay: (a) in an usual
stress diagram; (b) normalized with respect to the equivalent pressure at failure, pe (from Rampello, 1989, 1991).

Figure 11.15. Progressive destructuring in rotary cored samples of Crag sand by unload-reload compression cycles
(from Hight et al., 1997).

in Figure 11.15 for compression tests, as the previous maximum axial strain increases, the secant mod-
ulus decreases, demonstrating destructuring.
In general, it is relatively difficult to quantify the effect of destructuration by shearing because of
the possible formation of shear discontinuities in triaxial testing, and with possible orientation of parti-
cles (see Section 8). One way of inducing homogeneous shear strains in soils is by sampling them with
tubes driven or jacked into the ground. During sampling, the soil is subject to a compression-extension-
compression cycle of loading (Baligh, 1985). These aspects are described by Hight & Leroueil (2002*);
however, a few examples can be considered to illustrate the effects of destructuration by shearing.
Figure 11.16a compares CIU test results performed on samples of clay taken by 5 cm diameter tube and
hand-cut block. The effect of destructuration is similar to that observed in Figure 11.12 where the soil
was destructured by compression. Figure 11.16b compares oedometer compression curves obtained on
samples taken by blocks and with an 85 mm sampler. Destructuration increases the recompression index

159
09031-02[2].qxd 18/Oct/02 12:12 AM Page 160

Figure 11.16. The effects of shear from sampling disturbance on strength, on the abruptness of yield and on yield
stress: (a) undrained triaxial tests (from Lefebvre, 1970); (b) oedometer tests (from Holtz et al., 1986); (c) limit state
(after La Rochelle et al., 1981).

Cs, decreases the compression index Cc and decreases the preconsolidation pressure, which is less well
defined. Figure 11.16c, which compares limit state curves obtained on high quality and low quality
samples, demonstrates that the entire limit state curve shrinks due to destructuration induced by sam-
pling and shearing.
Loss of microstructure may also result from the shearing accompanying deep compaction of sands
(Mitchell & Solymar, 1984).

11.3.3 Destructuration by swelling


In some materials in which the inter-particle bonds are not strong enough to resist the swelling forces
generated when the effective stresses are reduced, there can be destructuration by swelling as indicated
in Figure 11.14 for Todi clay. Figure 11.14a shows the strength envelopes obtained on the undisturbed
material and after free-swelling, i.e. after 12% expansion. It can be seen that the strength envelope of
the undisturbed soil is much higher than that of the same material destructured by swelling. As for
destructuration by drying, when the same results are normalized with respect to pe (Figure 11.14b), the
strength envelope of the undisturbed soil remains above the other envelopes but the difference, reflecting
the effect of microstructure only, is smaller. As shown by Leroueil & Vaughan (1990) and Hight et al.
(2002a*), London clay is also prone to destructuration by swelling.
Also indicated on Figure 11.7 is the fact that Go can decrease at small effective stresses in compari-
son with the Go–  relationship of the intact soil, if the material is subjected to destructuration by
swelling. Some destructuring can also be observed in Figure 11.6 for the cemented sand unloaded and
then reloaded.

160
09031-02[2].qxd 18/Oct/02 12:12 AM Page 161

Table 11.1. Strength characteristics of selected fresh and weathered mudrocks and
overconsolidated clays (after Taylor & Cripps, 1987).

Fresh Weathered

 (°) c (kPa)  (°) c (kPa)

London clay 20–29 31–252 17–23 1–18


Bearpaw shale 25–30 10–152 20–28 0–41
Lower Oxford clay 23–40 10–216 21–28 0–20
Upper Lias clay 25 27 18–25 1–17
Keuper Marl 40 30 25–42 2–80
Coal Measures Mudrock 46 131 26–39 0–25
Go (by ultrasonic pulse test)

Tuff Non-weathered
Progress of weathering
(x 10 kgf/cm 2 )

1.0

0.5 Weathered
4

0.1
0.1 0.5 1.0 5.0 10
Effective confining pressure σ c' (kgf/cm 2 )

Figure 11.17. Variation of Go with the confining stress at different degrees of weathering (after Nishi et al., 1989).

11.3.4 Destructuration by weathering


Weathering may also contribute to destructuration of geomaterials. It is generally associated with cli-
matic phenomena (wetting-drying cycles, freeze-thaw cycles) or chemical action. It may produce a
degradation of the bonds between particles or aggregates and thus a destructuration of the material
resulting in a reduction of strength, as indicated in Table 11.1 from Taylor & Cripps (1987). Nishi et al.
(1989) examined the variation of the small strain shear modulus, Go, of three different soft rocks pro-
gressively weathered by freeze-thaw cycles, drying-wetting cycles or natural outdoor weathering.
Figure 11.17 shows typical results. It can be seen that Go decreases as the weathering process prog-
resses, in a manner very similar to that shown in Figure 11.7.

11.3.5 Destructuration by fatigue


Yoshinaka & Osada (1995) examined four soft sedimentary rocks under static and cyclic loading. The
monotonic tests were performed in undrained conditions at an axial rate of 0.01 mm/min. In the fatigue
(or cyclic) tests the axial load was repeatedly applied at a frequency of 0.2 Hz until the specimen
reached failure. The normalized maximum deviatoric stress, defined as the ratio of the maximum devi-
atoric stress applied in cyclic tests to the deviatoric stress at failure in the monotonic tests, varied from
80% to 120%. Figure 11.18 shows the normalized maximum deviatoric stress applied in the fatigue test
as a function of the number of cycles necessary to reach failure for the Yokohama siltstone. It can be
seen that the number of cycles to failure increases as the applied stress level decreases and that failure
at the first loading cycle is reached at a normalized maximum deviatoric of about 1.2. This latter ratio
is explained by the fact that strain rates in cyclic loading tests are much larger than those in monotonic
tests. It thus appears that the strength under cyclic loading is influenced by strain rate, as evidenced
under the first loading cycle, and by fatigue during the subsequent loading cycles. Also, on the basis of
measured pore pressures, Yoshinaka & Osada (1995, 1998) concluded that the reduction in strength
results from a lowering of the strength envelope with the accumulation of cycles and time.

161
09031-02[2].qxd 18/Oct/02 12:12 AM Page 162

1.5

Normalized maximum deviator stress


Yokohama siltstone

1.0

s 3 = 0.5 MPa
0.5
s 3 = 2.0 MPa

0
2 3
1 10 10 10
Number of cycles to failure (N)

Figure 11.18. Number of cycles to failure as a function of the maximum deviator stress applied in undrained cyclic
traixial tests (from Yoshinada & Osada, 1995).

1.0

Cyclic strength
Monotonic strength (0.5%/h)
0.8 Extrapoled from monotonic test at
0.5%/h for a strain rate equivalent
σr ) max

to a frequency of 0.1 Hz
2 σ ' vc

0.6
Extrapolated strength
( σa −

for the first cycle


0.46
Rate effect
0.4
0.35

0.2
2 3
1 10 10 10
Number of cycles

Figure 11.19. Cyclic strength of normally consolidated B6 clay (modified after Lefebvre & Leboeuf, 1987).

Lefebvre & Leboeuf (1987) performed undrained triaxial tests on eastern Canada clays under monot-
onic conditions at different strain rates and under cyclic conditions (Figure 11.19) and came to conclu-
sions similar to those of Yoshinaka & Osada (1995): (a) cyclic behaviour is influenced by strain rate; (b)
due to fatigue, there is a lowering of the strength envelope of the overconsolidated material as the num-
ber of cycles increases. This latter aspect was confirmed by Santos et al. (1997), on a residual soil from
Rio de Janeiro.
As indicated by several authors, the destructuration process is probably associated with accumulation
of plastic strains (Gens & Nova, 1993; Kavvadas & Amarosi, 1998). This means that the behaviour
of structured soils should probably be examined in terms of strains rather than in terms of stresses.
Unfortunately, the strains involved in the pre-yield range are small, difficult to measure, and with an ori-
gin that is generally not clearly defined.

11.3.6 Conclusion on destructuration


From previous paragraphs, it appears that the effects of destructuration are similar whatever the mode
of destructuration: compression, shearing, swelling, weathering or fatigue. They can be summarized as

162
09031-02[2].qxd 18/Oct/02 12:12 AM Page 163

2.6

2.4
Void ratio e

2.2
120 d

2.0

(a) (b) α
1.8
4 5 6 7 8 9 2 34 5 6 7 8 9 2 3 4 5
3
10 10
σ 'v (kPa) Go (kPa)

Figure 11.20. Compression curve (a) and variation of the maximum shear modulus Go with void ratio (b) for
artificially sedimented Jonquière clay (from Perret, 1995 and Leroueil et al., 1996).

follows: (a) decrease in stiffness of the soil inside the limit state curve; (b) decrease in peak shear
strength and in preconsolidation pressure, as well as a shrinkage of the entire limit state curve; and
(c) decrease in compression index.

11.4 Structuration and restructuration with time (aging)


There is evidence that, in some materials, microstructure may develop or may recover, at least to some
extent, with time.
Leonards and Altschaeffl (1964), Perret (1995) and Leroueil et al. (1996) showed that, in young
reconstituted clays, the development of preconsolidation pressure during secondary consolidation can
be much higher than that related to the decrease in void ratio. For reconstituted Jonquière clay consoli-
dated to 10 kPa (Figure 11.20a) and left under this applied stress, the preconsolidation pressure was
equal to 18.5 kPa after 120 days whereas it should have been 11.5 kPa on the basis of the change in void
ratio only. It is worth noting that Burland (1990) reported void ratio-effective overburden stress rela-
tionships for normally consolidated deposits (sedimentation compression line) at void ratios higher than
the compression line obtained in the laboratory on reconstituted clay (see Section 18.3 and Figure 18.2),
indicating development of microstructure during the formation process of sedimentary clays.
Schmertmann (1991) presented data of a 20 cm diameter rigid plate loaded over a dry clean and uni-
form sand. Figure 11.21 shows the plate settlement versus plate load. It can be seen that after 844 min
of rest under an approximate constant load, the sand had acquired a resistance to further loading well in
excess of what could be expected on the basis of the settlement (or change in void ratio) observed dur-
ing that time. Other evidence of strength increases with time have been reported by Schmertmann
(1991), and by Mitchell & Solymar (1984) who observed sand restructuration with time after destruc-
turation by deep compaction.
Due to their viscous behaviour, soils have a tendency to rearrange their particles and generally
decrease their void ratio under constant effective stress. They may also be subjected to structuring phe-
nomena that have a tendency to fix particles into position. The rate of strain determines which of the
two opposing mechanisms would dominate. At relatively high strain rates the former mechanism will
dominate as described in Section 9. At relatively low strain rates, structuring effects may become impor-
tant and modify the strain rate effects. For example, on Batiscan clay, Leroueil et al. (1985a) observed
compression curves from CRS oedometer tests that were well organized, according to the stress-strain-
strain rate model (Equations 9.1 to 9.3), for strain rates larger than 107 s1 (Figure 11.22). However, as

163
09031-02[2].qxd 18/Oct/02 12:12 AM Page 164

can be seen on the figure, the test performed at the strain rate of 1.69  108 s1 does not fit with the
other results, the compression curve crossing the compression curves obtained at larger strain rates.
This is attributed to structuring phenomena that become dominant at this latter strain rate. Marques
et al. (2002) observed similar phenomena on another eastern Canada clay; they also observed that struc-
turing effects become more important as temperature increases from 10° to 50°C.

0.3
Dry medium sand
Plate settlement (in)

12 min
0.4
9
844 min 8
aging 9 9
0.5 7
UF K-box
K = active
8
0.6
250 300 350 400
Load on 8" diameter plate (lbs)

Figure 11.21. Aging-preconsolidation effect demonstrated by plate load test on sand (from Schmertmann, 1991).

Effective stress σ'v (kPa)


0 50 100 150 200 250
0
Batiscan
-1
CRS tests (s )
-5
5
ε v1 = 1.43 x 10 -6
ε v2 = 5.30 x 10
-6
ε v3 = 2.13 x 10
-7
ε v4 = 5.33 x 10
-7
10
ε v5 = 1.07 x 10
-8
Strain ε v (%)

ε v6 = 1.69 x 10
Excess pore pressure ²∆ub (kPa)

15 100

20 50

25 0
ε v4 , ε v5 and ε v6
Figure 11.22. Typical CRS oedometer tests on Batiscan clay (from Leroueil et al., 1985a).

164
09031-02[2].qxd 18/Oct/02 12:12 AM Page 165

12 INFLUENCE OF DISCONTINUITIES

This Section is mostly concerned with clayey soils and soft rocks in which the initially intact material
has been sheared into elements due to different processes: unloading, tectonic activity, ice-thrusting,
etc. These processes result in fissured stiff clays, stiff jointed clays, clay shales, tectonized clay shales,
etc. These materials are then made up of intact elements, often microstrutured, separated by disconti-
nuities such as fissures, joints or faults. These discontinuities influence their mechanical behaviour, not
so much in terms of compressibility but mostly in terms of stiffness and strength. The main influencing
factors are: mechanical properties of the intact material; density of discontinuities compare to the scale
of the considered problem (size of specimens in the laboratory or of a natural slope in situ); previous
relative displacement between adjacent elements; and direction of shearing compared to the orientation
of discontinuities. The general behaviour of these geomaterials has, in particular, been described by
Skempton & Petley (1967), Skempton (1977), D’Elia et al. (1998), and Picarelli et al. (2000, 2002*).

12.1 Stiff fissured and jointed clays


The study made by Calabresi & Manfredini (1973) on the stiff jointed Santa Barbara clay illustrates
well the influence of discontinuities on the strength of clays. In addition to well defined bedding planes,
the Santa Barbara stiff clay deposit is subdivided by several families of joints and by a series of faults
with a shear zone up to 20–30 cm thick.
Direct shear tests were performed with shear planes in the intact soil and along joints. Figure 12.1
shows shear stress-displacement curves for an effective normal stress of 100 kPa. The intact clay shows
very brittle behaviour. On the other hand, shearing along a joint shows almost no peak. It is worth not-
ing, however, that in this latter case, failure is reached at a very small displacement. Figure 12.2a shows
the corresponding strength envelopes. The peak strength envelope along joints has the same friction
angle as the intact clay, but a much smaller cohesion. Calabresi & Manfredini (1973) suggested that this
cohesion could be due to the roughness of the joint surface. No difference appears in the residual val-
ues of strength, but the displacement required to reach residual conditions is smaller along a joint than
for the intact material.
Along a fault, stress-displacement curves presented very small peak or no peak at all with a strength
very close to the residual strength of the clay (Figure 12.2b).
The behaviour of the Santa Barbara clay was related to the surface characteristics of the discontinu-
ities: large displacements occurred along faults, the surface appeared polished and slickensided, and the
residual strength had already been reached; along joints (and bedding planes), no displacement had

Figure 12.1. Stress displacement curves obtained in multiple stage direct shear tests on Santa Barbara clay,
n  100 kPa (after Calabresi & Manfredini, 1973).

165
09031-02[2].qxd 18/Oct/02 12:12 AM Page 166

Figure 12.2. Strength envelopes for the Santa Barbara clay: (a) in the intact material and along a joint at elevation
110 m; (b) along a fault at elevation 125 m (from Calabresi & Manfredini, 1973).

Figure 12.3. Strength of fissures and joints compared with critical state strength for London clay (after Skempton,
1977).

occurred previously and the surface appeared rough. As a consequence, the strength was characterized
by the same friction angle as the intact soil but almost no cohesion. It corresponds to the “post-rupture”
strength indicated in Section 8.3. Backanalyses of instabilities confirmed the strength parameters found
in the laboratory study.
These views were reinforced by studies made on London clay. Skempton (1977) found that the
strength of joints and fissures in London clay (c  0;   20°) is smaller than the peak strength of
the intact material (c  7 kPa;   20°), larger than the residual strength (c  0; r  13°) and very
similar to the “fully softened” or “critical state” strength obtained on remoulded, normally consolidated
clay (Figure 12.3).

12.2 Tectonized clay shales


Italian researchers, in particular Picarelli et al. (2000, 2002*), described the behaviour of so-called tec-
tonized clay shales, clay shales that have been subjected to large strains and displacements during major
tectonic events. The fissures materials are characterized by fine grained indurated soils crossed by a
network of small polished or slickensided fissures. These are spaced from a few millimetres to a few
centimetres and subdivide the material into very small fragments. The water content is lower than the
plastic limit, with a liquidity index typically between 0.2 and 0.4. When softened due to unloading
and swelling, or weathering, these fissured clay shales become a new material made up of small litho-
relicts of the original formation mixed with a softer clay matrix. In such materials, the overall water con-
tent results from that of the lithorelicts with negative liquidity index and that of the matrix with a
liquidity index which can be close to 1.0. The mechanical behaviour of these latter softened clay shales

166
09031-02[2].qxd 18/Oct/02 12:12 AM Page 167

Figure 12.4. Bisaccia clay: strength obtained by drained and undrained triaxial tests (from Olivares et al., 1993;
Picarelli et al., 2002*).

is also summarized by Picarelli et al. (2000, 2002*) and is not described here. Only aspects of the
behaviour of “intact” tectonized clay shales will be presented hereunder.
Olivares et al. (1993) presented results of drained and undrained triaxial tests performed on the
Bisaccia tectonized clay shale in both undisturbed and reconstituted conditions. For the reconstituted
clay, the soil was isotropically consolidated under a maximum effective stress of 500 kPa and then
unloaded before shearing. Figure 12.4 shows the strength envelopes, normalized with respect to the
equivalent consolidation pressure pe determined on the isotropic consolidation line of the reconstituted
clay. The strength envelope of the undisturbed material is much lower than the strength envelope of the
same soil reconstituted, which has been attributed to the fissures along which the shear strength is close
to residual.
Picarelli et al. (2000) indicate two main stages during shearing of tectonized clay shales: a first stage,
in which the local strength (i.e. the shear strength along fissures) is not completely mobilized. The spec-
imen then remains uniform and the stress path followed in an undrained test during this stage is essen-
tially at p  cst; a second stage, in which the local shear strength is attained and the mechanism of soil
deformation is characterized by important plastic “strains” due to slipping along fissures. D’Onofrio
et al. (1998) examined the behaviour of Bisaccia clay, both undisturbed and reconstituted, in the pre-
failure range. Performing cyclic and resonant column torsional shear tests at increasing strain levels,
they found the elastic threshold strain to be typically 3 times larger for the reconstituted material than
for the undisturbed one. In fact, in comparison with other experimental data (Figure 4.16), the reconsti-
tuted results are similar whereas the tectonized clay shales give smaller threshold strain values.
The presence of hard fragments in tectonized clay shales also influences the measured residual
strength. As shown in Figure 12.5, under a given normal effective stress, the undisturbed soil has a
larger residual friction angle than the reconstituted material, especially under small normal effective
stresses.

12.3 Fissured and jointed soft rocks


The stiffness of rock masses is strongly influenced by discontinuities (Cook, 1992). This was demon-
strated for chalk by Clayton et al. (1994, 2002*). These authors compared the mass and intact stiffness
of chalk from three sites with different porosities. The rock at each site was highly fractured with typi-
cal sub-horizontal discontinuity spacing of between 10 and 160 mm. The mass stiffness was determined
using surface wave geophysics; the intact stiffness was measured in the laboratory, in triaxial tests, on
locally instrumented specimens.
Figure 12.6 shows intact and mass stiffness for the three sites. The intact stiffness is strongly depen-
dent on porosity and density. The mass stiffnesses are much lower, but similar for the three sites, which
was explained as follows by Matthews et al. (2000): The high-density chalk displayed a loose fracture

167
09031-02[2].qxd 18/Oct/02 12:12 AM Page 168

Figure 12.5. Residual shear strength of Laviano tectonized clay shale (from Picarelli, 1991; Picarelli, 2000).

Figure 12.6. Mass and intact stiffness for chalks with different intact dry density (after Clayton et al., 1994;
Clayton et al., 2002*).

block system resulting in the rock mass stiffness being less than 20% of the intact rock stiffness. The
low-density chalk had a tight fracture block system resulting in a much smaller difference between
intact and mass stiffness.

13 INFLUENCE OF SOIL WATER CHEMISTRY

It is known that pore fluid composition influences the behaviour of active clays (Mesri & Olson, 1970,
1971; Sridharan, 1991, 2001; Mitchell et al., 1973; Barbour & Fredlund, 1989; Di Maio, 1996a, b;
Di Maio & Onorati, 2000). The following sections will briefly illustrate possible effects.

168
09031-02[2].qxd 18/Oct/02 12:12 AM Page 169

Figure 13.1. Influence of pore water chemistry on the behaviour of reconstituted Bisaccia clay: (a) Intrinsic
compression lines; (b) CIU triaxial test results on the normally consolidated material (after Di Maio & Onorati,
2000; Di Maio, 2002).

13.1 Isotropic and one-dimensional compression


Figure 13.1a shows isotropic and one-dimensional compression curves for the Bisaccia clay (a smectite
soil) reconstituted at a water content slightly larger than the liquid limit with distilled water and with a
1 M NaCl solution (Di Maio, 2002). The influence of water composition is important, the compression
curves with distilled water being at much higher void ratio. It is worth noting that the opposite holds for
some clays such as Norwegian and eastern Canada clays: for these materials, the sedimentation curve
obtained in fresh water is at void ratios that are smaller than those of the sedimentation curve obtained
in a marine environment (Bjerrum, 1967; Locat, 1982); (see Figures 11.2b and 18.8).
Picarelli et al. (2000, 2002*) illustrated the consequence of an active marine clay being exposed to
fresh water for the Bisaccia clay. The clay, reconstituted with 1 M NaCl was first one-dimensionally
compressed to high stresses and then unloaded to a vertical stress of 10 kPa. Finally, when exposed to
distilled water under the same vertical stress, the soil specimen swelled by 34% of its volume.

13.2 Peak and critical state strength envelopes


Di Maio & Onorati (2000) and Di Maio (2002) performed undrained and drained triaxial compression
tests as well as direct shear tests on the same Bisaccia clay reconstituted and immersed with distilled
water or with 1 M NaCl solution. Figure 13.1b compares the peak strength envelopes deduced from
undrained compression tests on normally consolidated specimens. It can be seen that the influence of
pore fluid composition is very important, the friction angle dropping from about 18° with NaCl solu-
tion to about 10° with distilled water.

13.3 Residual strength


Di Maio (1996a, b) showed that the effect of pore water chemistry can also be important on residual
friction angle. Figure 13.2 shows results obtained on three southern Italy clays and on Ponza bentonite

169
09031-02[2].qxd 18/Oct/02 12:12 AM Page 170

Figure 13.2. Residual shear strength against normal stress of specimens reconstituted with water and of specimens
reconstituted with saturated NaCl solution (from Di Maio, 1996a).

Figure 13.3. Residual friction angle of Ponza bentonite under normal stresses of 250–400 kPa for various
concentrations of NaCl solution (after Di Maio, 1996b).

prepared either with saturated NaCl solution or with distilled water. It can be seen that, with the excep-
tion of Marino clay, r changes from about 14° with the saturated NaCl solution to 6° with distilled
water, which may be extremely important from a practical point of view. Figure 13.3 shows the varia-
tion of r with concentration of NaCl solution for Ponza Bentonite; most of the change in r occurs for
concentrations between 0 and 35 g/l NaCl, which should be an encouragement for evaluating residual
friction angle in laboratory with pore water having a chemistry close to that existing in situ.

14 INFLUENCE OF PARTIAL SATURATION

14.1 Soil-water characteristic curve


The soil-water characteristic curve (SWCC), also called the water retention curve, is the relationship
between the amount of water in the soil (gravimetric or volumetric water content or degree of satura-
tion) and matric suction (ua  uw) in the soil, called herein suction. It gives an idea of the pore size
distribution of the considered soil and important information concerning the hydraulic as well as the

170
09031-02[2].qxd 18/Oct/02 12:12 AM Page 171

Figure 14.1. Soil-water characteristic curve illustrating the regions of desaturation for a soil (from Sillers et al.,
2001).

mechanical characteristics and behaviour of soils when unsaturated. Vanapalli (1994) and Sillers et al.
(2001) define three stages during the process of desaturation (increasing suction; see Figure 14.1):

– Capillary saturation zone: The pore water is in tension in this zone, however, the soil remains essen-
tially saturated due to capillary forces. This zone ends at the air entry value or bubbling pressure
(ua  uw)b where the air starts to enter in the largest pores (Figure 14.1).
– Desaturation or funicular zone: water within the pores is increasingly displaced by air in the desatu-
ration zone. The desaturation zone ends at the residual water content where the liquid phase becomes
discontinuous.
– Zone of residual saturation: the water is tightly held to the soil. Increasing soil suction does not result
in significant changes in water content and moisture movement occurs mainly as vapor flow. The suc-
tion at which all soils approach zero water content would be approximately 1 000 000 kPa (Croney &
Coleman, 1961; Fredlund & Xing, 1994).

Figure 14.2 shows SWCC for different soils. The SWCCs illustrate the influence of the pore size dis-
tribution of the soil, the air entry value being about three orders of magnitude larger for the Regina clay
than for the sand. From capillary theory, the radius of curvature of a water surface is related to the
matric suction as follows:

r  C/(ua  uw) (14.1)

where C is a constant equal to 2Ts cos /w ⬵ 150  103 N/m where Ts is the surface tension of
water,  is the angle of contact between soil particles and air-water interface and w is the unit weight
of water. The air-entry pressure values that can be deduced from Figure 14.2 would thus correspond to
the largest entrance pore radius in the order of 0.035 mm for the sand and 0.05 m for Regina clay.
The SWCC is defined by progressive drying of an initially saturated soil (ABC in Figure 14.3). Upon
wetting, there is a hydraulic hysteresis (CDA in Figure 14.3). If wetting is relatively rapid, some air
(residual air content) is trapped in the soil and the wetting curve goes towards a degree of saturation,
called submergence degree of saturation, Srf, that is smaller than 1.0 (see dashed line in Figure 14.3;
also see Rojas, 2002). It is only with time, by diffusion of air, that the degree of saturation will pro-
gressively increase from Srf to 1.0.

171
09031-02[2].qxd 18/Oct/02 12:12 AM Page 172

Figure 14.2. Typical soil-water characteristic curve for four Canadian soils (from Vanapalli et al., 1999).

Figure 14.3. Elasto-plastic representation of hydraulic hysteresis in a rigid soil.

The two curves ABC and CDA are limits of an accessible domain for any cycle of drying and
wetting (e.g., 12341 in Figure 14.3). Buisson and Wheeler (2000) explained hysteresis as follows
(Figure 14.4): On wetting, the radius of the meniscus separating air and water increases as suction
decreases (see Equation 14.1) and the meniscus migrates inwards from (1) to (2) in Figure 14.4a. As
suction decreases, further and the radius of the meniscus continues to increase, a limiting condition is
reached (3) when the radius is equal to that of the largest sphere that can fit within the void. Any further
reduction of suction would result in the void flooding with water. The value of suction at which this void
floods with water during a wetting path is therefore governed by r3. On drying, the radius of the menis-
cus separating air and water decreases and the meniscus migrates inwards from an initial position (4)
towards (5) in Figure 14.4b. A critical point is reached with the radius r5 for which any further increase
in suction would result in the void filling with air. Hydraulic hysteresis thus results from the fact that
the critical radius for emptying of the void, r5, is smaller than the critical radius for flooding the void, r3.
In situ, soils may be subjected to drying-wetting cycles such as 1–2–3–4–1 in Figure 14.3 and could be
at point E at a given time. Consequently, suction in soil cannot be defined on the basis of water content
or degree of saturation only.
Numerous equations have been proposed in the literature to represent SWCC. Reviews are provided
by Fredlund & Xing (1994), Leong & Rahardjo (1997) and Sillers et al. (2001).

172
09031-02[2].qxd 18/Oct/02 12:12 AM Page 173

Figure 14.4. Mechanism for flooding and emptying of a void (after Buisson & Wheeler, 2000).

Figure 14.5. Relation between air entry value and void ratio for a clayey soil (after Kawai et al., 2000).

The SWCC of a soil reflects its pore size distribution. It thus varies with the void ratio and the fabric
of the soil. This can be illustrated by the following examples:
– Kawai et al. (2000) determined SWCC for clay specimens statically compacted at different void
ratios. As shown in Figure 14.5, the air entry value increases from 5 kPa to 83 kPa when the void ratio
decreases from 2.1 to 0.9.
– Even at about the same void ratio, a given soil may have different fabrics and thus different SWCCs.
This is evident in Figure 14.6 showing very different SWCCs for a non-plastic glacial till compacted
at degrees of saturation smaller and larger than the corresponding value at the optimum condition. It
can be seen that the air entry value is about one order of magnitude larger when the soil is compacted
wet of optimum. Similar results were obtained by Vanapalli et al. (1999). Therefore, except for coarse
and clean materials, SWCC should not be seen as an intrinsic characteristic of a soil.

173
09031-02[2].qxd 18/Oct/02 12:12 AM Page 174

Figure 14.6. Soil-water characteristic curves of LG2 till compacted at different conditions (after Watabe et al.,
2000).

14.2 Strength of unsaturated soils


A number of researchers have attempted to extend the principle of effective stress to unsaturated soils,
with the objective to describe soil behaviour as a function of a unique stress variable. Several expres-
sions have been suggested for a generalized effective stress, the best known being that proposed by
Bishop (1959):
  (  ua)  (ua  uw) (14.2)

where  total stress, uw  pore-water pressure, ua  pore-air pressure, (  ua)  net total stress,
(ua  uw)  s  matric suction, and   effective stress parameter which has a value of 1 for saturated
soils and a value of 0 for dry soils.
When Equation 14.2 is applied to the strength of unsaturated soils, the Mohr-Coulomb criterion
becomes:
f  c  {( n  u a )  (u a  u w)}tan  (14.3)

where c  effective cohesion (same as for saturated soil) and   effective friction angle (same as for
saturated soil).
In general, unsaturated soils have a shear strength that is larger than that of the same soil saturated.
Fredlund et al. (1978) proposed that it could be expressed as a linear combination of the net normal
stress ( n  ua) and of the matric suction (ua  uw).

f  c  ( n  ua)tan   (ua  uw)tan b (14.4)

where tan b  strength parameter associated with matric suction.


Equation 14.4 is schematically represented in Figure 14.7. The term (ua  uw)tan b appears as a
cohesion due to suction.
Experimental evidence shows that Equation 14.4 is valid for a variety of unsaturated soils, but, most
of the time, within limited ranges of matric suction and net normal stress. Also, limitations on the use
of Equation 14.4 have been found:
a) Escario & Saez (1986), (Figure 14.8), Gan & Fredlund (1988), De Campos & Carillo (1995), and
Oloo & Fredlund (1996), have shown that at suctions smaller than the air entry value, whereas the
soil remains saturated, b is equal to . In these conditions, a change in suction is equivalent to a
change in effective stress.
b) Delage et al. (1987) and Maâtouk et al. (1995) observed strength envelopes for silts obtained at dif-
ferent suctions which were converging towards a unique point, indicating that there is no effect of
suction for net normal stress in excess of a given value.

174
09031-02[2].qxd 18/Oct/02 12:12 AM Page 175

Figure 14.7. Shear strength of unsaturated soils (after Fredlund et al., 1978).

Figure 14.8. Shear strength as a function of suction for different net normal stresses. Direct shear tests on Madrid
clayey sand (from Escario & Saez, 1986).

c) It is known that in some cohesionless soils shear strength first increases and then decreases as suc-
tion is progressively increased. Examples of such behaviour are reported by Donald (1956), Escario
& Juca (1989) and Fredlund et al. (1996). An example is shown in Figure 14.9.
Equation 14.4 and its previously mentioned limitations result from the observation of soil when
sheared at different net normal stresses and suctions. An apparently more rational approach has been
proposed by Fredlund et al. (1996) and Vanapalli et al. (1996). The main features are presented in the
following paragraphs.
As indicated by Fredlund et al. (1995), the contribution of matric suction to shear strength of an
unsaturated soil can be assumed to be proportional to the product of matric suction, (ua  uw), and the
normalized area of water, aw:

  aw (ua  uw)tan  (14.5)

where aw is the ratio of the area of water corresponding to any degree of saturation to the total area of
water at saturation.

175
09031-02[2].qxd 18/Oct/02 12:12 AM Page 176

Figure 14.9. Results of direct shear tests on fine Frankston sand (data from Donald, 1956; from Fredlund et al.,
1996).

The normalized area of water, aw, presents similarities with the normalized volumetric water content
, and Fredlund and co-workers proposed the following relationship:

aw  
(14.6)

where
  r
 (14.7)
s  r

in which , s and r are the volumetric water content respectively at the considered suction, at satura-
tion and at the residual condition. When r is equal to zero,  is also the degree of saturation.

is a soil parameter dependent upon soil type. According to Fredlund et al. (1996),
is close to 1 in
sandy soils, but can be greater than 1 when plasticity increases.
The strength of unsaturated soil can then be written as:

f  c  ( n  ua)tan   (ua  uw)


tan  (14.8)

Fredlund et al. (1996) consider that  is a function of matric suction and can be deduced from the
SWCC. As a result, the required parameters to determine the strength of an unsaturated soil would be
the strength parameters cand  of the saturated soil and the soil-water characteristic curve. It should
be emphasized that Equation 14.8 explains that b   at suctions smaller than the air entry value and
a possible decrease in shear strength when suction increases in excess of a given value, as shown in
Figure 14.9. An application of Equation 14.8 to a decomposed granite from Hong Kong is shown in
Figure 14.10. Figure 14.10a shows the soil-water characteristic curve for this soil and the calculated
variation of shear strength with suction is shown in Figure 14.10b, assuming a
value of 1.0.
When
 1 and r  0, 
becomes equal to Sr and Equation 14.8 becomes equivalent to Equation
14.3 with   Sr. Fleureau et al. (1995) and Öberg & Sällfors (1997) showed that Equation 14.3 gives
good agreement with the results for sands.
On the basis of shear strength data obtained on 14 different soils, Khalili and Khabbaz (1998)
observed that  in Equation 14.3 can be empirically related to the suction ratio {(ua  uw)/(ua  uw)b}
where (ua  uw)b is the air entry value (see Figure 14.11). Khalili & Khabbaz (1998) specified that the
lower bound values in Figure 14.11 correspond to coarse-grained materials whereas the upper bound
values correspond to fine-grained materials.
It is worth noting that the shear strength of saturated soils is characterized by three different levels:
the peak strength envelope, upper part of the limit state curve; the strength envelope of the normally
consolidated soil, or critical state line; and the residual strength envelope (see Figure 8.2). At the present
state of understanding of unsaturated soil mechanics, these differences do not clearly appear although
they certainly should.

176
09031-02[2].qxd 18/Oct/02 12:12 AM Page 177

Figure 14.10. Behaviour of a completely decomposed tuff from Hong Kong: (a) soil-water characteristic curve;
(b) shear strength versus suction (from Fredlund et al., 1996).

Figure 14.11. Relationships between the effective stress parameter  and suction ratio (ua  uw) /(ua  uw)b (from
Khalili & Khabbaz, 1998).

177
09031-02[2].qxd 18/Oct/02 12:12 AM Page 178

14.3 Compression curves of unsaturated soils


Bishop (1959) proposed the use of the generalized effective stress given by Equation 14.2 to describe
the behaviour of unsaturated soils in both compression and shear. The strength is then defined by
Equation 14.3. The possible use of this generalized effective stress for characterizing the compressibil-
ity of soils in general appears more difficult. Jennings & Burland (1962) illustrated this fact. These
authors tested a number of samples of silt that were air dried from slurry. The hatched zone in Figure 14.12
shows the domain occupied by the compression curves of the air-dried soil. The full line on the same
figure is for a sample soaked under zero stress and then compressed. The curve of the air-dried soil lies
above that of the soaked sample. Some of the air-dried samples were loaded to applied pressures of 200,

Figure 14.12. Oedometer compression curves on air dry and saturated silt showing the effects of soaking (after
Jennings & Burland, 1962, and Burland, 1965).

Figure 14.13. Isotropic compression curves at essentially constant suction performed on speswhite kaolin (after
Wheeler & Sivakumar, 1995).

178
09031-02[2].qxd 18/Oct/02 12:12 AM Page 179

400, 800 and 1600 kPa and then soaked. As seen in Figure 14.12, they collapsed and came to equilib-
rium (full triangles) at a void ratio very close to the compression curve of the initially soaked soil. Some
other air-dried samples were loaded to the same applied pressure values. They were then soaked with
the applied stress being reduced in an effort to maintain a constant void ratio. It can be seen that the
samples come to equilibrium (full circles) on the soaked line. Soaking corresponds to a decrease in suc-
tion and then, according to Equation 14.2, to a decrease in effective stress. On the basis of classical soil
mechanics, a decrease in effective stress should be associated with an increase in void ratio of the soil.
The reverse is observed, indicating that the behaviour of partly saturated soils cannot in general be
described by effective stress. From such results, it has been stated that the stress state of unsaturated
soils has to be described by two stress state variables, usually (  ua) and (ua  uw).
The effect of matric suction on the compression of unsaturated soils has been studied by many
researchers (Matyas & Radhakrishna, 1968; Wheeler & Sivakumar, 1995; Leroueil & Barbosa, 2000;
and many others). Figure 14.13 shows isotropic compression curves obtained in constant suction tests
performed on a compacted kaolin. Figure 14.14a also shows isotropic compression curves, obtained on
an “artificial loess”, a silt from Trois-Rivières, Québec, reconstituted at a high void ratio of 1.04.
Figure 14.15 shows compression curves obtained along stress paths with constant net stress ratio dif-
ferent from 1.0 on the same Trois-Rivières silt. Figure 14.16 shows compression curves deduced from
CRS tests performed at constant suction on a residual soil (50% sand, 45% silt and 5% clay) reconsti-
tuted (Figure 14.16a) and on the same soil reconstituted at the same void ratio, but with 2% cement

p - ua (kPa)
2 3
10 10 10
0 800
LC
eo ~ 1.04 e o = 1.04
0.1 600 eo ~ 0.8
u a - u w (kPa)

e initial ~ 1.04
eo ~ 0.7
0.2 u a - uw (kPa)
600 400
0.3
²∆e

400 Elastic
200 200
0.4 150
80 (b)
(a)
0.5 0
0 10 20 30 40 50 60 70 80
(p - ua ) y (kPa)

Figure 14.14. Isotropic compression tests on Trois-Riviéres silt subjected to different suctions: (a) Compression
curves when eo ⯝ 1.04; (b) Loading collapse curves (after Maâtouk et al., 1995).

p - u a (kPa)
2 3
10 10 10
1.1

1.0

0.9
σ3 - ua
e

0.8 K=
σ 1- ua
0.7 1.00
0.74
0.6 0.54
0.5

Figure 14.15. Compression tests at various stress ratios, for (ua  uw)  80 kPa. Trois-Rivières silt (from Maâtouk
et al., 1995).

179
09031-02[2].qxd 18/Oct/02 12:12 AM Page 180

(σ a - ua ) or (σ a - u w) (kPa) (σ a - ua ) or (σ a - u w) (kPa)
2 3 2 3
10 10 10 10 10 10
0 0

5 5
Vertical strain (%)

Vertical strain (%)


10 CRS tests
10
(ua - uw) (kPa)
15 (ua - uw) (kPa)
15 0
20 0 30
50 50
100 20
25 100
200 300
30 300 25 500
(a) 500 Non-cemented (b) Cemented (2%)
35 30

Figure 14.16. One-dimensional compression curves for non-cemented and cemented (2%) Viçosa residual soil
(from Leroueil & Barbosa, 2000).

σ y, e ∆σ y, st + suc
∆σ y, suc ∆σy, st
500
Cement
400 0%
2%
Suction (kPa)

300

200
LC a (0%)
LC a (2%)
100

0
0 40 80 120 160 200
(σ a - ua ) (kPa)

Figure 14.17. Loading-Collapse (LCa) curves for cemented and non-cemented Viçosa residual soil (from Leroueil
& Barbosa, 2000).

added (Figure 14.16b). In all cases, the compression of soil under a given net stress ( n  ua) decreases
when suction increases. In some of the tests performed on the reconstituted residual soi (Figure 14.16),
suction was decreased to zero after reaching a volumetric strain larger than 10%; it can be seen that soil
conditions then move on or close to the compression curve corresponding to zero suction.
The compression curves previously presented also show that the gross yield stress increases with
suction. Alonso et al. (1990) suggested plotting this yield stress versus suction and calling the corre-
sponding relationship the “loading-collapse” (LC) curve. Such curves are shown in Figures 14.14b for
the Trois-Rivières silt and 14.17 for the reconstituted residual soil non-cemented and cemented. In this
latter case, the axial yield stress is equal to 52 kPa when the soil is saturated and progressively increases
with suction to reach a value of 175 kPa at a suction of 500 kPa. In the case of the “artificial loess” that
was not standing on its own without suction, the LC curve does not cross the (p  ua) axis at a suction
of zero.
It has sometimes been observed that the slope of the compression curve in the e–ln(p  ua) dia-
gram and in the post-yield range, s, decreases as suction increases. However, Wheeler & Sivakumar
(1995 and Figure 14.13), Sharma (1998), Leroueil & Barbosa (2000, and Figure 14.16) and Machado
& Vilar (2002*) found an opposite tendency. As indicated by Wheeler & Karube (1996), this could be
related to the stress range investigated.

180
09031-02[2].qxd 18/Oct/02 12:12 AM Page 181

Fn u w< ua Fsuc

Tst Tsuc
Tn
(a) (b) (c)
Figure 14.18. Resistance to slippage at a contact between particles due to: (a) external force; (b) bonding; and
(c) partial saturation.

14.4 Combined influence of matric suction and microstructure


To understand the combined influence of matric suction and microstructure on unsaturated soil behav-
iour, and in particular yielding, it may be important to keep in mind the basic phenomena intervening at
the level of particles or aggregates, in particular that yielding mostly corresponds to inter-particle slip-
pages, even when loading is isotropic. Figure 14.18 shows a simple contact between two particles. If the
two particles are subjected to an external force with a component normal to the contact Fn, the resis-
tance to slippage, Tn, is equal to Fntan  for a cohesionless material (Figure 14.18a). If the same
contact is bonded (Figure 14.18b), the resistance to slippage is increased by the strength of the bond, Tst.
If the soil is partially saturated, there is some water surrounding the contact, a capillary force, Fsuc,
normal to the contact and a slippage force, Tsuc  Fsuc tan  (Figure 14.18c). The total resistance to
slippage could simply be the sum of Tn, Tst and Tsuc. However, as indicated by Alonso et al. (1994), the
resistance and thus the yield stress resulting from the combined effect of applied stress, partial satura-
tion and bonding could be larger than the sum of the contributions of the three factors. This can be due
to different distributions of water at the contact between particles or, as mentioned by Alonso et al.
(1994), to bond strength that is influenced by water content and indirectly suction. For the case pre-
sented in Figure 14.18, this would result in an additional component of the resistance to slippage Tstsuc.
The total resitance to slippage could then be written as follows:

T  Tn  Tst  Tsuc  Tstsuc (14.9)

This is for a single contact. For a soil specimen, yielding results from slippage of many contacts.
However, the yield stress should be described in the same form as Equation 14.9:

y  y,e   y,st   y,suc   y,stsuc (14.10)

in which y,e would be the gross yield stress associated with the void ratio;  y,st reflects the influence
of bonding alone;  y,suc reflects the influence of partial saturation alone; and  y,stsuc reflects the
combined effect of microstructure and partial saturation.
As previously indicated, Leroueil & Barbosa (2000) performed a series of CRS oedometer tests at
different suctions on a residual soil reconstituted and on the same soil reconstituted at the same void
ratio, but with 2% cement added. The results are presented in Figures 14.16a and b respectively. The
gross yield stresses deduced from these tests are reported in Figure 14.17 as a function of suction. In
spite of some variability, the Loading-Collapse curves, called here LCa curves to denote that yielding is
reached in the axial direction, are well defined. The LCa curve for the cemented material has been
drawn taking into account all the results obtained in these oedometer tests and triaxial tests that were
indicating that the yield stress obtained at a suction of 300 kPa was too small in comparison with other
results (Leroueil & Barbosa, 2000). This figure can be used to examine the influence of void ratio,
bonding and partial saturation on yielding. The yield stress y,e for the non-cemented material at zero
suction reflects the influence of void ratio; the difference between the yield stresses obtained in satu-
rated conditions for the soil cemented and non-cemented,  y,st, reflects the influence of bonding alone;
and the influence of suction alone is indicated by the Loading Collapse curve of the non-cemented
material,  y,suc. The results show, however, that, as predicted by Alonso et al. (1994) and Equation
14.10, the yield stress resulting from the combined effect of void ratio, partial saturation and bonding is
larger than the sum of the contributions of the three factors.

181
09031-02[2].qxd 18/Oct/02 12:12 AM Page 182

s
fs(s) SI

Elastic zone
I Y1

Y2

p*o p - ua
(a)
q

Increasing
M suction

Ysat

p*o p - ua
(b)

Figure 14.19. Limit state curves and elastic domains for unsaturated soils – BBM (after Alonso et al., 1990).

Carvalho et al. (2002) showed a similar combined effect of partial saturation and microstructure
when comparing the results of isotropic compression tests performed on a silica sand cemented at dif-
ferent contents of Portland cement.

14.5 Barcelona-type models for unsaturated soils


Alonso et al. (1987, 1990) extended the concepts of limit and critical states to unsaturated soils. The
proposed model, the Barcelona Basic Model (BBM), is essentially based on two considerations: first,
the behaviour of unsaturated soils is controlled by the net total stress (  ua) and the matric suction
(ua  uw); second, the proposed model is an extension of the Modified Cam-clay model (Roscoe &
Burland, 1968).
An important component of the approach is the Loading-Collapse (LC) yield locus and surface. The
concept is depicted in Figure 14.19 and can be explained as follows: if a saturated soil isotropically con-
solidated has an isotropic preconsolidation pressure such as po*, its limit state curve is Ysat in
Figure 14.19b. If suction is increased, the isotropic yield stress increases along the LC curve as seen in
Section 14.3 (Figure 14.19a) and the entire limit state curve increases in size, as shown in Figure 14.19b.
It results in an enlargement of the elastic zone. In BBM, there would also be an elastic limit, the “Suction
Increase (SI)” line corresponding to the maximum suction experienced by the soil (Figure 14.19a). In a
(p  ua)–q–s diagram (Figure 14.20), the limit state curves generate a limit state surface. As shown
in Figure 14.19b, the critical state line moves upwards when suction increases. In the model proposed
by Alonso et al. (1990), it is considered that the slope (thus the friction angle) is constant and inde-
pendent of suction.
Figures 14.12 to 14.17 can be understood by using Figure 14.19a and considering initial conditions
at a point such as I. Wetting of unsaturated specimens under constant net total stress (Figure 14.12) is
associated with a decrease in suction and collapse at Y2 (Figure 14.19a); increase in net total stress
under constant suction first shows an elastic behaviour and then development of plastic volumetric
strains after yielding at a point such as Y1 (Figure 14.19a). The volumetric strains generated in both
cases cause a movement of the LC curve towards larger net total stresses.

182
09031-02[2].qxd 18/Oct/02 12:12 AM Page 183

M s

SI
p*o

p - ua (po , s)
LC
Figure 14.20. Three-dimensional view of the yield surface in BBM (after Alonso et al., 1990).

As it is an extension of the Cam Clay model, BBM has the same weaknesses. In particular it does not
consider the possible effects of anisotropy, viscosity and microstructure. It also considers that the gross
isotropic yield stress increases as soon as there is some matric suction (Figure 14.19), and this is prob-
ably not true when suction does not exceed the air entry value and the soil remains saturated.
Several other elasto-plastic models, similar to the BBM model, have been proposed to describe and
predict the behaviour of unsaturated soils: Karube (1988) and Karube & Kato (1989) proposed one in
which the following equation is used for the critical state line:

q  M((p  ua)  f1(ua  uw)) (14.11)

This equation is similar to the one obtained by Fredlund et al. (1995, 1996) and Vanapalli et al.
(1996) when they relate the shear strength and the soil-water characteristic curve (Equation 14.8); the
model proposed by Wheeler & Sivakumar (1995) adds one state parameter, the water content; Cui &
Delage (1996) introduce anisotropy; Geiser et al. (1999, 2000) use ( n  uw) and (ua  uw) as stress
state variables; Buisson & Wheeler (2000) and Wheeler (2002) are working on the development of a
model that would incorporate the influence of hydraulic hysteresis.
Existence of the elastic limit SI at constant suction (Figure 14.19) has been subject to discussions, in
particular because it does not seem appropriate for unsaturated expansive clays (Gens & Alonso, 1992).
These latter authors considered BBM appropriate for describing the volumetric behaviour of unsatu-
rated soils of low plasticity and suggested an enhanced version, called BExM (Barcelona Expansive
Model), for expansive clays. The BExM is clearly explained by Alonso et al. (2001) and only general
aspects are described here. Two levels are distinguished:
• The microstructural level corresponding to the arrangements of clay particles. These arrangements
can exist in a homogeneous matrix or in aggregates. It is likely that this level remains saturated and
that effective stress concepts hold. The behaviour would be essentially elastic and volumetric.
• The macroscopic level accounts for larger scale features such as aggregates. Coupling between
microstructure and macrostructure may result in a build-up of macrostructural elastoplastic strains
when elastic microstructural strains take place.
The behaviour of the macrostructure follows the BBM and is in particular characterized by the LC
curve (Figure 14.21). As microstructure is considered saturated and elastic, volumetric strains depend
only on p and are thus equal to zero along the Neutral Line (NL at 45° in s vs (p  ua) diagram;
Figure 14.21) corresponding to p  p  uw  p  s  cst. As microstructural swelling and microstruc-
tural shrinkage may affect the macrostructure and induce irreversible changes in void ratio, two addi-
tional yield curves parallel to NL were introduced in BExM. These two yield curves SI (Suction Increase)
and SD (Suction Decrease) are associated with shrinkage and swelling respectively (Figure 14.21).
More details on BExM can be found in Gens & Alonso (1992) and Alonso et al. (1994, 2001).

183
09031-02[2].qxd 18/Oct/02 12:12 AM Page 184

Figure 14.21. BExM: yield loci for microstructure and macrostructure in (p  ua)–s plane (from Gens & Alonso,
1992, and Alonso et al., 1994).

400
Saturated
)
=0

s = 100 kPa
(s

s = 200 kPa
L
CS

300

200
GF
q (kPa)

Y,
=2 s
s=

00

100
10

kP
0k

a
Pa
s=
0k
Pa

AL
0

-100
0 100 200 300 400
p -ua (kPa)

Figure 14.22. Yielding and GFY curves for a natural weathered colluvium soil from Brazil (data from Machodo &
Vilar, 2002*; curves drawn by Leroueil & Barbosa, 2000).

14.6 Limit state curves of unsaturated soils


As for saturated soils (Section 3), limit states of unsaturated soils can be defined by a variety of tests
and, up to now, limit state curves have been defined for about 10 different unsaturated soils. Most of
these soils were reconstituted and Machado & Vilar (2002*) are presenting an exceptional set of limit
state curves determined on a naturally unsaturated colluvium.
Figure 14.22 shows the limit state curves defined by Machado & Vilar (2002*) on Sao Carlos collu-
vium at suctions of 0, 100 and 200 kPa. The extension of the yield curve with suction is quite clear.

184
09031-02[2].qxd 18/Oct/02 12:12 AM Page 185

1000
s = (u a - u w )
s = 1500 kPa
800

s = 800 kPa
600 s = 400 kPa
q (kPa)

s = 200 kPa
400

200

K0w

0
0 200 400 600 800 1000
p -u a (kPa)

Figure 14.23. Limit state curves of the compacted Jossigny silt at various suctions (from Cui & Delage, 1996).

400
Normal compression points
Critical state points 2.2
Normal compression
Deviatoric stress q (kPa)

300 line, s = 200 kPa


Critical state line
s = 200 kPa 2.1
Specific volume v

200
2.0
Critical state line
100 s = 200 kPa
1.9

(a) (b)
0 1.8
0 100 200 300 50 100 200 400
Mean net stress (p - ua ) (kPa) Mean net stress (p - u a) (kPa)

Figure 14.24. Test paths for constant-suction shear tests conducted as s  200 kPa (from Wheeler & Sivakumar,
1995).

From these results, it appears that the limit state curves obtained on this soil are essentially isotropic.
Limit state curves obtained by Zakaria et al. (1995) on a compacted kaolin after isotropic consolidation
also seem to be isotropic (These results are presented later in Figure 14.27). Figure 14.23 presents the
limit state curves of the compacted Jossigny silt for suctions of 200, 400, 800 and 1500 kPa (Cui &
Delage, 1996). These curves, as well as those obtained by Maâtouk et al. (1995), are inclined to the hor-
izontal, which probably reflects an anisotropy of particle arrangement due to the mode of preparation of
the soil specimens. This implies that, as for saturated clays, the particle arrangement and the shape of
the limit state curve will vary depending on the stress path followed, i.e., there will be kinematic hard-
ening. This is confirmed by Romero et al. (2002).
Results obtained by Wheeler & Sivakumar (1995) on a compacted kaolin also indicate that the con-
cept of critical state applies to unsaturated soils. As shown in Figure 14.24, a variety of tests performed
at a suction of 200 kPa give ultimate points which are on a single critical state line in both q–(p  ua)

185
09031-02[2].qxd 18/Oct/02 12:12 AM Page 186

q
pe
velo
en
h
gt
tren
rs
ea
Sh

Yield
curve
1 2

(a)
p - ua

800

150 600

σ a- σ r (kPa)
σa- σ r (kPa)

100 400
Type 1 behaviour Type 2 behaviour
50 u a - u w= 150 kPa 200 u a - u w= 150 kPa
σ 3 - u a = 10 kPa σ 3 - u a = 200 kPa
(b) (c)
0 0
0 5 10 15 20 25 0 5 10 15 20 25 30 35
ε a (%) εa (%)

Figure 14.25. Drained shear (CID) of unsaturated Trois-Rivières silt specimens: (a) Schematic stress paths; (b)
specimen consolidated in the pre-yield range; (c) specimen consolidated in post-yield range (after Maâtouk et al.,
1995).

and v–(p  ua) diagrams. However, it appears from other data shown by Wheeler & Sivakumar (1995)
that the critical state lines obtained at different suctions are not parallel in an e–log (p  ua) diagram,
and tend to converge at large net pressures.
An application of CSSM to unsaturated soils is illustrated in Figure 14.25. Maâtouk et al. (1995) per-
formed drained triaxial compression tests ( 3  ua  cst) at different suctions on the unsaturated silt
from Trois-Rivières after isotropic consolidation within the limit state curve and outside, as schema-
tized in Figure 14.25a. For tests starting from post-yield conditions, plastic strains develop progres-
sively and the maximum deviatoric stress and critical state conditions are reached at axial strains of
25–30% (Figure 14.25c). For tests starting from pre-yield conditions, the behaviour is first elastic with
a rapid increase in deviatoric stress, and plastic strains develop only when the stress path has reached
the limit state curve; critical state conditions are obtained at axial strains of 12–15%. These simple
patterns of behaviour and many others reported in the literature (Wheeler & Sivakumar, 1995; Cui &
delage, 1996; Geiser et al., 2000; etc.) confirm the applicability of CSSM concepts to unsaturated soils.

14.7 A simple mechanistic model for unsaturated soils


The simple yielding model initially proposed by Larsson (1977) for saturated soft clays and discussed
in Section 3.6 has been extended to unsaturated soils by Leroueil & Barbosa (2000). Starting from
Larsson’s idea that “a soft clay cannot be submitted to stresses in any direction higher than the previous
stresses in that direction without yielding”, the yield curve of soft clays can be approximately defined
by 4 segments, as shown in Figure 3.14. Leroueil & Barbosa (2000) added the following considerations:
– When a soil (orthotropic, with the same properties in all horizontal directions) is formed, either nat-
urally or artificially, it develops a particle or aggregate arrangement so that the soil skeleton has yield
stresses in saturated conditions yv and yh in the vertical and horizontal directions respectively. In a
stress diagram, the line reflecting this induced anisotropy is called the anisotropy line (AL) and is
characterized by KAL  yh/ yv.

186
09031-02[2].qxd 18/Oct/02 12:12 AM Page 187

0)
(s = on

( σ a − σ r )/2
L
Cs CS ressi
p
com
B'
Bs suc. = s
Bo AL
suc. = o As
Ao
Fs O σ *ryo (a)
σ *ayo σ *rys σ *ays
Do
( σ a + σ r )/2 - ua
Ds
D'
Es
CS
L
ext ( s = 0
ens
ion )
Suction

LC r LCa
s

(b)
o
* σ *rys
σ *ryo σ ayo σ *ays

Figure 14.26. Description of the Given-Fabric-Yield (GFY) model (from Leroueil & Barbosa, 2000).

– As indicated in Section 14.4 (Figure 14.18 and Equations 14.9 and 14.10), matric suction generates
a resistance to slippage at the contacts between particles (or aggregates). Its global effect should thus
reflect the distribution of these contacts, and the anisotropy line (AL) should be the same as for the
same soil when saturated. This is confirmed by the studies referred to in Section 14.6, which show
that yield curves obtained at different suctions keep the same shape (see Figures 14.22 and 14.23).
– The previous statement implies that, contrary to what has often been stated in the literature, the effect
of matric suction on soil behaviour is not automatically isotropic when the air phase is continuous as
capillary forces then act at the contacts, with the resulting effect reflecting the distribution of these
contacts.
The model proposed by Leroueil & Barbosa (2000) called GFY (for Given Fabric Yield) model can
be explained with the help of Figure 14.26. If a soil in saturated conditions has a limit state curve such
as OBoAoDo in Figure 14.26a, a given suction should extend the cap BoAoDo to BsAsDs, with an
increase in axial yield stress from *ayo to *ays and an increase in radial yield stress from *ryo to *rys.
The variation of the axial and radial yield stresses with suction defines two Loading-Collapse curves,
LCa and LCr respectively (Figure 14.26b). However, the anisotropy line, AL, remains the same inde-
pendently of suction, the ratio *rys / *ays at any suction is constant and equal to KAL.
By increasing the axial yield stress from *ayo to *ays, matric suction increases the shear strength of
the soil in compression from point Bo to point Bs, or from Bo to B, in Figure 14.26a. It can be easily
demonstrated that this compression strength increase is equivalent to a cohesion in compression due to
suction, cc,suc, defined as follows:

sin  ⋅ cos 
cc,suc  ( *ays  *ayo ) (14.12)
1  sin 

187
09031-02[2].qxd 18/Oct/02 12:12 AM Page 188

This equivalent cohesion depends only on suction and on the distribution and number of contacts. So,
as long as this distribution remains the same, which should be the case in non-expansive soils for
net stresses smaller than yield stresses *ays and *rys, cc,suc should be a constant, as indicated on
Figure 14.26a. According to Equation 14.12, this cohesion is typically 27% of the increase in axial yield
stress due to suction.
In extension, the shear strength increase due to a given suction is associated with an increase in radial
yield stress from *ryo to *rys (Figure 14.26a). This corresponds to an equivalent cohesion in extension,
ce,suc, equal to:

sin  ⋅ cos 
ce,suc  ( *rys  *ryo) (14.13)
1  sin 

As in compression, if fabric remains the same for net stresses smaller than yield stresses, the cohe-
sion in extension due to suction should be constant, as shown in Figure 14.26a.
Equations 14.12 and 14.13 indicate that cohesions in compression and extension due to suction are
generally different. The ratio ce,suc /cc,suc is equal to KAL, and takes a value of 1.0 only for soils having an
isotropic fabric.
The GFY model previously described depends on the hypothesis that fabric remains the same when
there are changes in suction or net stresses. If there is a reduction in the number of contacts between par-
ticles when decreasing net stresses, the cohesion due to suction should decrease; also, if fabric is mod-
ified by loading the soil outside the gross yield curve or by shearing, the effect of suction should change
accordingly. The GFY model has been described for cohesionless materials. It can, however, be easily
extended to microstructured materials, the reference being in all cases the gross yield curve in saturated
conditions.
The GFY model has been applied to the data sets of Machado & Vilar ( 2002*), Zakaria et al. (1995)
and Cui & Delage (1996). The comparisons of the experimental results with the GFY model are briefly
described hereunder.
– As previously indicated, Machado & Vilar (2002*) tested a natural colluvium from Brazil. The gross
yield points are shown in Figure 14.22. They are perfectly fitted by the GFY model, assuming an
isotropic AL, which is in agreement with previous experience with residual soils (see Leroueil &
Vaughan, 1990).
– Zakaria et al. (1995) tested a kaolin, compacted and then subjected to an isotropic stress history.
Figure 14.27 shows points obtained at gross yielding under suctions of 0, 100 and 200 kPa as well as
the critical state lines obtained for the same suctions. It can be seen that the proposed model fits the
data points well when assuming an isotropic AL, reflecting the stress history of the material. The
strength appears however slightly overestimated.

400 a
CSLs indicated by kP Pa
Zakaria et al. 200 00 k )
s= =1 =0
s (s
L
300 CS
Pa
0 k kPa
q (kPa)

0
2 0
200 s = = 10
s GFY
s = 200 kPa
100
GFY, s = 0
GFY AL
s = 100 kPa
0
0 100 200 300 400 500
p - u a (kPa)

Figure 14.27. Yielding and GFY curves for compacted kaolin after isotropic consolidation (data from Zakaria et al.,
1995; curves drawn by Leroueil & Barbosa, 2000).

188
09031-02[2].qxd 18/Oct/02 12:12 AM Page 189

– Maâtouk et al. (1995) and Cui & Delage (1996) tested compacted silty soils and both observed limit
state curves reflecting some anisotropy. This is demonstrated on Figure 14.23 showing the results
obtained by Cui & Delage (1996) on the compacted Jossigny silt at suctions from 200 to 1500 kPa.
Trying to apply the GFY model to Cui & Delage results, Leroueil & Barbosa (2000) concluded that:
(a) KAL for this compacted soil is in the order of 0.45; (b) the model overestimates the increase in
strength with suction. However, Cui & Delage (1996) were estimating yielding in shear tests per-
formed in the pre-yield range well before failure, which is not in accordance with the GFY model
which assumes full mobilization of the capillary forces. By correcting Figure 14.23 accordingly in
the pre-yield (overconsolidated) domain, Figure 14.28 is obtained. It can be seen that the GFY model
captures relatively well the overall behaviour of the soil, but has a tendency to overestimate the
strength.
Figures 14.22, 14.27 and 14.28 indicate that the simple GFY model that incorporates anisotropy and
the effect of suction captures relatively well the behaviour of unsaturated soil, but could overestimate
the strength increase due to suction. Some explanations for such behaviour are presently being investi-
gated at Université Laval and could justify the application of a correction factor to Equations 14.12 and
14.13. In any case, it is thought that not only the strength increase due to suction can be related to the
SWCC of the soil (Equation 14.8; Fredlund et al., 1996), but the entire limit state curve.
The GFY model also implies that, if the soil is compressed to stresses in excess of the limit state
curve and void ratio decreases, there could be a change in SWCC and a change in cohesion associated
with a given suction. On the other hand, if there is a decrease in the number of interparticle contacts
when the soil is unloaded under a given suction, the suction component of cohesion should decrease.

14.8 Influence of matric suction on small strain stiffness


A comprehensive study on the effect of partial saturation on small strain stiffness has been presented by
Vinale et al. (2001). The material tested was the compacted Metrano silty sand. Figure 14.29 shows the
Loading-Collapse curves deduced from isotropic compression tests performed on the soil compacted at
the optimum and wet (wopt + 2.5%) of optimum. The results confirm that the variation of yield stress
with matric suction depends on the fabric of the soil, which is different at the optimum and on the wet
side of optimum (see Figure 14.6 and associated text).

1200

1000
)
=0
a
kP

(s
00

L
CS
15

800 AL
s =
Y,
GF

600
40 = 80
q (kPa)

20 s = s
0

400
0

s (kPa)
s=

1500
200 800
400
200
0
0 200 400 600 800 1000
p -u a (kPa)

Figure 14.28. Yielding and GFY curves for compacted Jossigny silt (data from Cui & Delage, 1996).

189
09031-02[2].qxd 18/Oct/02 12:12 AM Page 190

Figure 14.29. Loading-Collapse curve of the optimum and wet compacted Metramo silty sand (after Vinale et al.,
2001).

Figure 14.30. Initial shear modulus versus matric suction in suction controlled torsion shear tests on Metramo silty
sand: (a) optimum compacted; (b) wet compacted (after Vinale et al., 2001). Note that the curves have been redrawn,
by the authors.

Figure 14.30 shows how the initial shear modulus obtained from torsional shear tests varies with
matric suction. As can be seen, the curves (as drawn by the authors) indicate an effect of suction on Go.
The Go values are greater than the Go value for the saturated soil, by an amount that is about the same
for the three applied stresses at a given compaction condition. The effect of suction is however different
for the two compaction conditions. At least as a first approximation, the small strain shear modulus of
the unsaturated soil can be expressed as follows:

Go,unsat  Go,sat  f(ua  uw) (14.14)

This equation is logical with the model shown in Figure 14.18 and described by Equation 14.10 in
which the effect of suction on a soil with a given fabric is additional to the effect of applied stress. This
is also consistent with the strength that is described by the addition of a component associated with
applied stress and a component associated with suction (Equations 14.4 and 14.8). Equation 14.14 is
however different from the corresponding equations proposed by Alonso (1998) and Vinale et al. (2001)
in which Go,unsat is described as the product of Go,sat and a function of suction.
Vinale et al. (2001) stated that the mechanical response observed is in conformity with the typical
S-shape of the SWCC, which is probably different at the optimum and on the wet side of optimum. At
low suction values, the soil is almost saturated and variations in (ua  uw) are equivalent to changes in
mean effective stress. At larger suctions, the increase in Go reflects capillary phenomena. It is worth not-
ing that, as for the gross yield stress (LC curves shown in Figure 14.29), the effect of matric suction on
Go is small at values larger than 150 kPa. Other test results presented by Vinale et al. (2001) indicate that
the shape of the G/Go versus  curves is not significantly influenced by suction.

190
09031-02[2].qxd 18/Oct/02 12:12 AM Page 191

Figure 14.31. Low-amplitude shear modulus versus degree of saturation for Glacier Way Silt (from Wu et al.,
1984).

Figure 14.32. Maximum shear moduli as a function of confining stress and suction. Compacted soil from
Vallfomes (data from Pintado (1993) presented by Alonso, 1998).

Wu et al. (1984) performed resonant column tests on compacted Glacier Way silt. Figure 14.31
shows Go versus the degree of saturation for a void ratio of 0.65 and different applied stresses. Several
remarks can be made: (a) Go is essentially the same in completely saturated (Sr  100%) and dry
(Sr  0%), (Go,dry  Go,sat); (b) the component of Go in excess of Go,dry is due to the effect of suction.
It can be seen in this case that Go does not vary significantly for Sr values between 100% and 70%,
increases as Sr decreases to reach a peak at a degree of saturation of about 15% and finally drops to
Go,dry as Sr decreases towards 0%; (c) (Go  Go,dry) is essentially the same for the three confining pres-
sures. As suction is primarily related to the degree of saturation and not to the applied stress, these
results support Equation 14.14 and the idea that at given void ratio and fabric, the effect of suction is
additional to the effect of applied stresses.
Pintado (1993; in Alonso, 1998) reports resonant column tests performed on the “Vallformés” silty
sand. Figure 14.32 shows Go as a function of suction for different confining stresses. Except for the
unconfined test, the increase in Go with suction in excess of Go under zero suction is about the same,
independently of the applied stress, confirming the validity of Equation 14.14.
Vinale et al. (2001) also investigated the effect of strain rate on Go measured during resonant column
and torsional shear tests performed under constant matric suction. The observations were very similar

191
09031-02[2].qxd 18/Oct/02 12:13 AM Page 192

Figure 14.33. Strain rate effect on normalized shear modulus of unsaturated Metrano siltys and compacted wet of
optimum (after Vinale et al., 2001).

for both optimum and wet compacted specimens. Figure 14.33 shows the normalized shear modulus
Go/Go(0.1%/min) as a function of shear strain rate for the wet of optimum compacted material under a
constant mean stress of 200 kPa. The strain rate effect seems to be independent of suction.

15 GENERALITIES ON SOIL PROPERTIES

It has been shown in Part I of this paper that all soils and soft rocks behave in a consistent manner. The
framework considered uses CSSM as a sound foundation but, for understanding the behaviour of dif-
ferent types of natural soils and soft rocks, it has been necessary to build on this foundation in order to
take into account the influence of anisotropy, viscosity, microstructure, discontinuities and partial satu-
ration, which are not considered in CSSM. This is shown in Sections 2 to 14. Part II of this paper
(Sections 15 to 20) is more oriented towards soil properties.
With its mineralogy, geological history and microstructure, a soil constitutes a specific entity. Each
parameter obtained from any test is like a photograph of this entity, and all parameters, all photographs
are thus linked together. In other words, the soil properties measured at a given depth in a soil deposit
are not independent.
When compiling the geotechnical characteristics of eastern Canada clays, Leroueil et al. (1983a)
divided soil characteristics in five different categories:
– Geology and mineralogy, which determine the nature of the considered material and its geological
history.
– Properties of the remoulded or reconstituted soil.
– Properties of the intact soil, which reflect the microstructure of the soil and, indirectly, its geological
history.
– Properties associated with the passage from intact to destructured or remoulded conditions, which
reflect the behaviour of the soil when the initial microstructure is broken.
– Hydraulic properties.
Geology and mineralogy are discussed in the Overview Paper by Locat et al. (2002*). The four other
parts are respectively in the Sections 16, 17, 18 and 19. Part II of this paper is completed by Section 20
on consolidation, which combines deformation characteristics and hydraulic conductivity.
Concerning the properties of intact soils, several remarks can be made:
(1) The shape of the limit state curve depends mainly on nc (and indirectly on plasticity index in most
cases).
(2) Strengths measured in different types of tests are on the limit state curve, but intercept it at different
points, depending on the stress path followed. This means that the strengths obtained by different

192
09031-02[2].qxd 18/Oct/02 12:13 AM Page 193

Figure 15.1. Undrained shear strength profiles from SBP, vane and laboratory tests on Drammen clay (from
Lacasse et al., 1981).

tests at a given depth may differ. This is illustrated by the undrained shear strength profiles obtained
on Drammen clay using the vane, SBP and triaxial compression and extension (Figure 15.1); simi-
lar results were also observed on Fucino clay (Soccodato, 2002*) and on many other sites.
(3) For a given nc value, the ratio between two undrained shear strengths determined by different tests
or between an undrained shear strength and the preconsolidation pressure should be constant. This,
in particular, implies that strength ratios should be approximately constant on a homogeneous site.
However, because of (1), these ratios may vary with nc or Ip.
(4) What has been said in (2) and (3) also applies to deformation characteristics.
Correlations in soil mechanics are thus justified. Empirical relationships must, however, be used with
extreme care: (a) the chosen parameters for the relationships may not be the most appropriate and the
relationship may not be general; (b) they may not apply in types of soil in which they have not been
developed or checked; (c) they are approximate and should not be used for final design. On the other
hand, they can be extremely useful for checking the consistency of test results, establishing preliminary
designs, and interpolating soil characteristic profiles on the basis of a limited number of test results.
These practical aspects are developed in the Overview Paper by Hight & Leroueil (2002*).

16 PROPERTIES OF SOILS IN REMOULDED OR RECONSTITUTED CONDITIONS

16.1 Generalities
The geotechnical properties of soils in remoulded or reconstituted conditions are the physical proper-
ties such as grain size distribution, particle shape, water content, plastic and liquid limits, plasticity
index, liquidity index and void index; they can also include compression and recompression indices of

193
09031-02[2].qxd 18/Oct/02 12:13 AM Page 194

the reconstituted soil as well as the critical state friction angle and the stress ratio under one-
dimensional compression in the normally consolidated range, Konc. Only the properties of remoulded
or reconstituted soils that characterize or can be used to assess the mechanical behaviour of natural soils
are described in this section.
The properties of soils in remoulded or reconstituted conditions are sometimes described as “intrin-
sic” properties. It is, however, important to realize that these properties can be influenced by several
factors and are not so “intrinsic” for the following reasons.
– The consistency limits and the compression curve depend on the fact that fine-grained soils may or
may not be dried before being reconstituted. Almeida & Marques (2002*) indicate that plasticity
index decreases by 10 to 30% with oven drying of the organic Sarapui clay. Several standards in the
world still recommend drying the soil before determining the consistency limits. As it may modify
the results, it is strongly recommended that this practice be abandoned.
– Several properties may be influenced by the conditions of sample preparation. This is true for plastic
soils as well as for cohesionless soils (See Section 5.4).
– Most properties of reconstituted active clays may be influenced by soil water chemistry (Section 13),
viscosity (Sections 9 and 10) and aging effects (Section 11).
– Natural soils are often laminated even at a microscale and reconstitution involves mixing of the layers.

16.2 Liquidity index


Houston & Mitchell (1969) suggested the existence of a unique relation between liquidity index, deter-
mined using the Casagrande apparatus or the fall-cone penetrometer, and remoulded shear strength of
clays. The range of values obtained was rather large, which can be explained by the fact that the liquid
limit obtained using the Casagrande apparatus is not associated with a unique value of the strength of
the remoulded clay. On the other hand, Wroth & Wood (1978) and Leroueil et al. (1983a) have shown
that, if only values of liquid limit obtained using the fall-cone penetrometer are considered, the relation
between IL and cur is very well defined (Figure 16.1). For values of liquidity index between 0.4 and 3.0,
Leroueil et al. (1983a) proposed:

cur(kPa)  1/(IL  0.21)2 (16.1)

Equation 16.1 shows that cur and IL are parameters that reflect the same soil characteristic, namely,
the consistency of remoulded clays.

16.3 Intrinsic compression line and void index


To characterize soil microstructure in terms of compressibility, Burland (1990) referred to the intrinsic
compression line (ICL) defined for clay reconstituted in the laboratory at a water content between wL
and 1.5 wL (preferably 1.25 wL ) and then consolidated under one-dimensional conditions. Burland also
defined the void index Iv as follows:

e  e*100 e  e*100
Iv   (16.2)
e*100  e*1000 C*c

where e *100 and e *1000 are the void ratios corresponding to v  100 kPa and v  1000 kPa respec-
tively, on the intrinsic compression line. C*c is the intrinsic compression index. When e  e *100, Iv  0
and when e  e *1000, Iv  1. The void index is thus defined by the two variables e *100 and C*c.
According to Burland (1990), the ICL can be represented with sufficient accuracy by the following
equation:

Iv  2.45  1.285x  0.015x3 (16.3)

where x  log v in kPa.


The ICL can be easily obtained experimentally. However, Burland (1990) showed that e*100 and C *c
could be reasonably related to the void ratio at the liquid limit, eL, as follows:

e *100  0.109  0.679eL  0.089eL2  0.016eL3 (16.4)

194
09031-02[2].qxd 18/Oct/02 12:13 AM Page 195

Figure 16.1. Relationship between liquidity index and remoulded shear strength (from Leroueil et al., 1983a).

and
C *c  0.256eL  0.04 (16.5)

16.4 Compression and swelling indices of non-microstructured clays


Numerous correlations between the compression index of non-microstructured clays and basic soil
properties have been proposed. Skempton (1944) and Terzaghi & Peck (1948) respectively proposed:

Cc  0.007(wL  10) (16.6)

and

Cc  0.009(wL  10) (16.7)

Nagaraj & Srinivasa Murthy (1986) established a relationship between the compression index and
the liquid limit based on considerations of physical chemistry. They came with Cc  0.2343eL, which
is very close to Equation 16.5.
Nagaraj & Srinivasa Murthy (1986) also suggested the following equation, which is also close to
equations proposed by other authors:

Cc  0.39e (16.8)

195
09031-02[2].qxd 18/Oct/02 12:13 AM Page 196

It is worth noting that Tanaka (2000) found for a variety of natural and sometimes microstructured
clays that Equation 16.7 is also valid under large applied stresses. It is also worth mentioning that Cc has
to decrease as the applied stress increases since the void ratio cannot become negative. For this reason,
Butterfield (1979) suggested representing compression curves in a log v–log e diagram.
Terzaghi et al. (1996) indicated that the swelling or recompression index Cs of clayey soils mostly
depends on the mineralogy of the clay, reflected by wL, and clay fraction, CF (see Figure 18.6). For non
microstructured clays, the ratio Cc/Cs is typically between 4 and 6.

16.5 Critical state friction angle


The strength of normally consolidated clays or loose cohesionless soils is characterized by the critical
state friction angle cs (or nc for clayey soils). In clayey soils, cs reflects the mineralogy of the parti-
cles and has often been related to the plasticity index of the soil, as shown in Figure 16.2. Although there
is significant scatter, the data show a general tendency for the friction angle to decrease as plasticity
index increases. However, as also shown on the figure, attapulgite and Mexico City clay are exceptions.
They can be explained by the fibrous shape of attapulgite particles and by the fact that Mexico City clay
contains glass particles and diatom shell fragments, which in both cases generate friction angles in
excess of 40°. In fact, it has to be recognized that plasticity index is not a property that reflects mechan-
ical behaviour. cs should rather be referred to the angularity or roughness of particles. This was
demonstrated by Tanaka (2000) and Shiwakoti et al. (2002) and is illustrated in Figure 16.3. These
researchers examined the effect of percentage of diatoms in clay on the friction angle. Singapore clay
and kaolin were considered. Without diatoms, the friction angle was slightly below the cs–IP relation-
ship shown in Figure 16.2, with values close to 23°, but increased to about 38°with 75% diatoms,
although the change in IP was small (Figure 16.3).
The friction angle at critical state of cohesionless soils depends on the mineralogy, shape and angu-
larity of the particles. It also depends on the grain size distribution of the soil. On quartz sand, Koerner
(1970) observed an increase in  from 30° to 38° as particle shape changed from sub-rounded to angu-
lar. Compiling cs of 17 well documented sands, Bolton (1986) reported values between 30° and 37°
and indicated typical values of 33° for a quartz sand and 40° for a felspathic sand. Coop & Airey
(2002*) indicate a typical value of 40° for carbonate sands. Miura & Yagi (2002*) show cs between
40° and 50°, with a tendency to decrease as the confining stress increases, for volcanic soils.

16.6 Residual strength friction angle


The residual strength is reached along a surface of discontinuity after large shear displacements. It is
characterized by a residual friction angle, r, without any cohesion. The residual friction angle depends

Figure 16.2. Values of friction angle nc for clays of various compositions as reflected in plasticity index (from
Terzaghi et al., 1996).

196
09031-02[2].qxd 18/Oct/02 12:13 AM Page 197

on the percentage of clay particles that can be reoriented during shearing and the ability of these parti-
cles to be reoriented. As plate-shaped particles are found among clay minerals most of the time, the
residual friction angle has often been related to the clay fraction, CF, of soils.
There have been several attempts to correlate residual friction angle with other soil index properties
such as the plasticty index Ip ( Ac  CF) that reflects the mineralogy through the activity and the clay
fraction. Several of these relationships are shown in Figure 16.4. Stark & Eid (1994) related r to the
liquid limit, clay fraction and effective normal stress (Figure 16.5).

Figure 16.3. Influence of diatoms on the friction angle of clays (data from Tanaka, 2000).

Figure 16.4. Residual strength: correlations with plasticity index (after Lupini et al., 1981, and Wesley, 2002*).

197
09031-02[2].qxd 18/Oct/02 12:13 AM Page 198

Figure 16.5. Relationship between drained residual friction angle and liquid limit (after Stark & Eid, 1994).

Figure 16.6. Horizontal stress coefficient for normally consolidated soils (modified after Mayne & Kulhawy,
1982).

It is worth noting that these relationships do not directly take into account the shape of the particles
and may consequently be misleading. For example, carrying out ring shear tests on highly plastic trop-
ical clays containing allophane and halloysite clay minerals, Wesley (1977) found r values that are
10 to 15° larger than the maximum values given in Figure 16.4 and 25 to 30° larger than the values
deduced from Figure 16.5. Wesley (2002*) also shows residual friction angles typically between 30 and 39°,
apparently independently of the plasticity index (see Figure 16.4). Also, for soils for which the particles

198
09031-02[2].qxd 18/Oct/02 12:13 AM Page 199

cannot be reoriented, for example eastern Canada clays which are mainly made up of rock flour, the
residual friction angle is not significantly lower than the critical state friction angle (Kenney, 1967).

16.7 Stress ratio under one-dimensional compression in the normally consolidated range, Konc
Numerous experimental studies have shown that the at-rest earth pressure coefficient during virgin
(normally consolidated) loading Konc may be expressed using the equation proposed by Jaky (1944) and
which can be approximated by:

Konc  1  sin cs (16.9)

Data compiled by Mayne & Kulhawy (1982) confirmed the general validity of this equation
(Figure 16.6). In particular, the Konc value of 0.31 obtained by Diaz-Rodriguez et al. (1992) and Diaz-
Rodriguez (2002*) on Mexico City clay is consistent with the friction angle of 43° obtained on this
material. However, Coop & Airey (2002*) found Konc values consistently higher (by 0.05–0.12) than
those given by Equation 16.9 for carbonate sands.
It is worth mentioning that, testing 20 natural marine clays from different parts of the world with the
same personnel and equipment, Watabe et al. (2002) found the following equations

Konc  1.05  sin  (16.10)

when  is defined on normally consolidated samples at large strains (( 1/ 3)max), and

Konc  0.95  sin  (16.11)

when  is defined on normally consolidated samples at the peak (( 1  3)max).

17 MECHANICAL PROPERTIES OF INTACT SOILS

17.1 Introduction
The properties of intact soils are those which reflect the mineralogy and the grain size distribution of
the soil, its geological history and its microstructure. They are thus existing in in situ conditions, where
and when they have not been disturbed by any stress release, loading, shearing, sampling, etc. They
are of paramount importance since they generally control the soil response to any change in boundary
conditions.
Numerous tools and interpretation methods have been developed for the determination of the prop-
erties of intact soils since the birth of soil mechanics, about one century ago. In this paper, however, the
authors do not aim at establishing a state-of-the-art on soil characterization. They rather will limit the
presentation to some usual tests and properties.

17.2 Soil characterization


Soil characterization is a difficult task. Numerous techniques, sometimes sophisticated, exist. There are
still, however, basic parameters that are difficult to obtain. This is particularly the case for the porosity
of sands. Lunne et al. (2002a*) present, an interesting discussion on this topic. In the present section,
almost only characterization with cone penetration testing will be considered.
For soils in which it can be pushed, the piezocone is probably the best tool for soil characterization.
When the cone penetrometer (CPT) or piezocone (CPTU) probe is pushed down at a constant rate of
penetration, several parameters are measured. The three main parameters are:
• The tip resistance qc; however, to get the true (generally termed the “corrected”) tip resistance qt, qc
has to be corrected for pore water pressure acting just behind the cone (u2).
• The pore pressure, generally u2 measured just behind the tip and used for correcting the tip resistance.
However, pore pressure can also be measured at other locations.
• The friction measured along a sleeve located behind the tip, fs.
Generally, cone penetrometers have a 10 cm2 base area, an apex angle of 60° and a surface area of the
friction sleeve of 150 cm2; the rate of penetration is normally 120 cm/min (ISSMFE, 1989). Cone pene-
trometers can also be equipped with one (or several) geophone(s) allowing measurement of shear wave

199
09031-02[2].qxd 18/Oct/02 12:13 AM Page 200

velocity at depth intervals or in a continuous manner. More details on cone penetration testing can be
obtained in the book by Lunne et al. (1997).
For soil characterization, Wroth (1984, 1988) suggested the use of the following normalized
parameters:
• Normalized tip resistance

q t  v0
Qt 
′v0 (17.1)

• Pore pressure ratio

u
Bq 
q t  v0 (17.2)

where u  u2  uo, with uo, pore water pressure in the ground at the considered depth.
• Normalized friction ratio

fs
Fr  (17.3)
qt  v 0

Several soil classification charts have been presented using cone penetration test data. Robertson
(1990) proposed the classification chart shown in Figure 17.1, using Qt, Fr and Bq. With the set of

Figure 17.1. Soil behaviour type classification chart based on normalized CPT/CPTU data (after Robertson,
1990).

200
09031-02[2].qxd 18/Oct/02 12:13 AM Page 201

parameters that can be measured at depth intervals of 1 cm or less, and the use of a classification chart,
the piezocone provides detailed information on soil stratigraphy. It is however important to realize that:
(a) the zones shown in classification charts are approximate; (b) in stratified deposits, the measured
parameters may not be representative of the penetrated layer; and (c) classification charts may not be
valid in soils for which they have not been checked. This seems to be the case for unsaturated colluvium
(Machado & Vilar, 2002*) and residual soils (Danziger et al., 1998; Mayne & Brown, 2002*; Viana da
Fonseca, 2002*).
Robertson et al. (1995) and Lunne et al. (1997) also suggested the classification chart shown in
Figure 17.2 and based on Qt and Go/qt.
It is seen in the following Sections that the piezocone can be used for determining design parameters
such as friction angle of cohesionless soils, preconsolidation pressure of clayey soils, small strain shear
modulus, etc. It has also been shown that piezocone data can be related to the relative density Dr of
sands. From well-documented results of tests performed in calibration chambers on five normally con-
solidated sands, Lancellotta (1983) obtained the relationship shown in Figure 17.3. As indicated by
Jamiolkowski et al. (1985), the correlation is not directly applicable to overconsolidated sands for which
no unique relationship exists between Dr and qc.

17.3 Preconsolidation pressure, apparent preconsolidation pressure and OCR


The preconsolidation pressure, or apparent preconsolidation pressure, is usually determined by conven-
tional 24-hrs oedometer tests, and most of our practical experience is based on this test. Because they
can be automated, are rapid to perform and give continuous compression curves, CRS oedometer tests
are very attractive. It must be recalled, however, that they are often performed at strain rates much
higher than those of 5  108 to 107 s1 associated with the conventional 24-hrs compression curve
and, consequently (Section 9.1), they may give larger preconsolidation pressure values.
The preconsolidation pressure obtained in conventional oedometer tests has often been correlated
with the net tip resistance qt obtained from piezocone test through the factor N t , as follows:

qt  vo  N t p (17.4)

Figure 17.2. Soil classification chart on normalized cone resistance and small strain shear modulus (after Lunne
et al., 1997).

201
09031-02[2].qxd 18/Oct/02 12:13 AM Page 202

Figure 17.4 shows the p vs (qt  vo) relationship obtained for 31 sites from the Champlain Sea
basin, Québec, with preconsolidation pressures varying from 38 to 940 kPa. The correlation is excellent
with an average N t value of 3.4, independently of the plasticity index (Demers & Leroueil, 2002).
Other studies (Mayne & Holtz, 1988; Larsson & Mulabdic, 1991; Mayne, 2001; Almeida & Marques,

100

qc
Dr(%) = 68 [log( )1]
80 povo
Relative Density, Dr (%)

60 2

2
Sand
40 Ticino
Ottawa
Edgar
Hokksund
Hllion mines
20

10 50 100 500 1000

qc/po
Cone Tip Resistance,
(vo/po)0.5

Figure 17.3. Correlation Relative density – Normalized tip resistance for normally consolidated sands (after
Lancellotta, 1983).

Figure 17.4. Relationship between preconsolidation pressure and net tip resistance (from Demers & Leroueil,
2002).

202
09031-02[2].qxd 18/Oct/02 12:13 AM Page 203

2002*; De Groot & Lutenegger, 2002*) indicate similar N t values generally between 3.25 and 3.5.
Mayne (2001) mentioned, however, that such N t values could underestimate the preconsolidation pres-
sure in fissured geomaterials.
It can also be mentioned that numerous approaches have been proposed for estimating OCR from
piezocone test data. The following studies can be mentioned: Konrad & Law (1987); Mayne (1991);
Chang et al. (2001).

17.4 Strength parameters


17.4.1 Drained strength parameters of sandy soils
The strength of cohesionless soils is characterized by a friction angle. Numerous indirect methods have
been proposed for estimating this parameter. Two are briefly described hereunder.
Hatanaka & Uchida (1996) established empirical correlations between the peak friction angle of
sandy soils obtained on high quality undisturbed samples, retrieved by means of in situ freezing in
Japanese sites, and SPT N values. The results were for an SPT efficiency of 78%. Mayne (2001)
adjusted the equation for an efficiency of 60% and obtained the equation:

  [15.4(N1)60]0.5 + 20° (17.5)

in which (N1)60 is the N60 value corrected to a reference stress of one atmosphere (patm):

(N1)60  N60/( vo /patm)0.5 (17.6)

The database used was characterized by: 0.15 mm  D50  7.5 mm, 1.6  Cu  22.3 and
34  Dr  81%. Hatanaka & Uchida (1996) also observed that available correlations are usually too
conservative. Figure 17.5a shows the relationship between  and (N1)60 described by Equation 17.5,
together with data obtained on silty sand from Atlanta and reported by Mayne (2001). It can be seen that
the agreement is very good.
Relationships have also been suggested for evaluating  from piezocone tests. Figure 17.5b shows
the relationship proposed by Robertson & Campanella (1983) for unaged, uncemented quartz sands.
Applying the relationship shown in Figure 17.5b to Piedmont silty sand, Mayne & Brown (2002*)
obtained representative results.

17.4.2 Strength parameters of clayey soils


Effective stress strength parameters of clayey soils have been discussed in Section 3.2 (Equation 3.1) for
overconsolidated clays, Section 16.5 (Figures 16.2 and 16.3) for normally consolidated clays, and in
Section 16.6 for the residual friction angle. It can be added that the mobilized friction angle also

Figure 17.5. Peak friction angle from in situ tests; (a) from SPT data (after Hatanaka & Uchida, 1996; from Mayne,
2001); (b) from CPT data (after Robertson & Campanella, 1983; from Mayne, 2001).

203
09031-02[2].qxd 18/Oct/02 12:13 AM Page 204

Figure 17.6. Approximate undrained strength ratios.

depends on the mode of failure, compression or extension, triaxial or plane strain and more generally
on anisotropy. Relative  values are indicated in Table 2.1.
In undrained conditions, the strength depends on the stress path followed and thus on the type of test
performed (Figure 15.1). Figure 3.5 shows that the normalized triaxial undrained compression strength
can be approximated by suTC/ p  sin nc/(1  sin nc), which is confirmed by data presented by
Kulhawy & Mayne (1990). This relationship is shown in Figure 17.6. Figure 3.7 shows that the ratio of
the undrained shear strengths in extension and in compression typically increases from 0.30 for non-
plastic soils to about 0.78 at a plasticity index of 60%. These results have been used to define an approx-
imate relationship between normalized triaxial undrained extension strength suTE/ p and the friction
angle. This relationship is also shown in Figure 17.6. As shown by Ladd (1991), the normalized
undrained shear strength obtained in direct simple shear (DSS) test is between the compression and the
extension undrained shear strengths. For normally consolidated clays, Ladd (1991) found suDSS/ p
equal to 0.225. This relationship has also been drawn in Figure 17.6. It is thought that these strength
ratios established for clays could also be valid for cohesionless materials.
The vane shear strength suv is often considered as a reference for undrained shear strength and has
often been used in correlations with other strength parameters. One of these correlations is the ratio
suv/ p versus plasticity index. Figure 17.7 shows a relationship proposed by Bjerrum (1973), data
obtained on eastern Canada clays (Leroueil et al., 1983a) and data obtained on Japanese clays (Tanaka,
2002). The relationships proposed by Bjerrum (1973) and Leroueil et al. (1983a) are slightly different
but show the same tendency. It can be added that most non-organic clays in the world more or less fol-
low these relationships. However, for organic clays (Leroueil et al., 1985c) and also for Japanese clays
(Tanaka, 2002), strength ratios are generally higher and apparently not related to plasticity index (see
Figure 17.7).
Lacasse et al. (1978) and Jamiolkowski et al. (1985) suggested that field vane strength should be
related to the in situ vertical effective stress and OCR using Equation 3.1. An m value of 1.0 in this
equation corresponds to a constant su/ p ratio. Compiling data from different authors, Tavenas &
Leroueil (1987) concluded that m is close to 1.0 for most natural clays and that the ratio su/ p does not
seem to be significantly dependent on OCR.
The tip resistance in a piezocone test is related to the undrained shear strength of clays as follows:

qt  vo  Nktsu (17.7)

Nkt depends on the undrained shear strengh considered. It is generally between 10 and 20, without
any clear tendency to vary with plasticity index, when vane shear strength is considered; it is lower, with
values between 8 and 14, when undrained triaxial compression strength is considered (Aas et al., 1986).
For the time being, Nkt has to be considered as site and shearing mode specific.

204
09031-02[2].qxd 18/Oct/02 12:13 AM Page 205

Figure 17.7. Variation of suv / p with plasticity index, for eastern Canada clays (from Leroueil et al., 1983a) and
Japanese clays (Tanaka, 2002).

3
Self-boring pressuremeter strength / vane strength ratio

Champlain sea clays (after Hamouche, 1995)


Other eastern Canada clays (after Konrad et al., 1985)

0
0 2 4 6 8
OCR

Figure 17.8. Correlation of self-boring pressuremeter/vane strength ratio with the overconsolidation ratio (from
Leroueil, 1997).

Figure 17.8 shows the ratio of the undrained shear strengths deduced from self-boring pressuremeter
and field vane tests as a function of the overconsolidation ratio for eastern Canada clay deposits. The
relationship is quite good and shows a clear decrease of the ratio suSBP/suv when the OCR increases. The
fact that SBP generally gives higher undrained shear strength than other usual tests (Figure 15.1) has
been analysed by Hamouche (1995) and Leroueil et al. (2002*); it seems to be associated with the vis-
cous behaviour of clays.

17.5 Deformation parameters


17.5.1 Small strain shear modulus
The small strain shear moduli can be directly measured in the laboratory or deduced from shear wave
velocity Vs as indicated by Equation 4.1. The shear wave velocity can be measured in different ways: shear

205
09031-02[2].qxd 18/Oct/02 12:13 AM Page 206

Figure 17.9. Correlation between Go and qc for cohesionless soils (modified after Rix & Stokoe, 1991).

waves propagating vertically, with particle motion in the horizontal direction (Vs-VH); shear waves prop-
agating horizontally, with particle motion in the vertical direction (Vs-HV); shear waves propagating hor-
izontally, with particle motion in the horizontal direction (Vs-HH). Vs-VH  Vs-HV is used to define
Gvh  Ghv  Go. Vs-HH is used to define Ghh (Section 4.2; Equations 4.8 to 4.10).
Several geophysical methods are used for determining shear wave velocities in soil deposits: down-
hole (DH) tests; cross-hole (CH) tests; seismic penetration cone (SCPT or SCPTU); Spectral Analysis
of Surface Waves (SASW), (Stokoe & Santamarina, 2000). The Vs-VH and Vs-HV values deduced from
DH, CH and SCPT tests are generally very similar to the values obtained by SASW testing technique
(e.g. Lo Presti et al., 2002*; Soccodato, 2002*). However, Butcher & Powell (1995) showed that the
shear wave velocities measured with the different techniques can be significantly different in heavy
overconsolidated clays or layered soils.
Baldi et al. (1989) proposed a relationship between Go and qc. After testing uncemented quartz sands,
Rix & Stokoe (1991) found data in the hatched area of Figure 17.9, with an average relationship
between Go/qc and a normalized qc value close to that proposed by Baldi et al. (1989). The correlation
found by Jamiolkowski & Lo Presti (1998) for sands and gravels of Pleistocene age from Messina
Straits is also shown in Figure 17.9. It can be seen that this latter correlation is close to the average
curve. However, as indicated by Rix & Stokoe (1991), the Go vs qc relationship for a given soil may
depend on factors such as fines content, particle angularity, etc. that are not considered in Figure 17.9.
For clays, Tanaka et al. (1994) proposed the following relationship:

Go  50(qt  vo) (17.8)

The Go/(qt  vo) ratios obtained by Shibuya & Tamrakar (2002*), and Leroueil et al. (2002*) are
also not far from the value of 50 given in Equation 17.8, but slightly lower. Data compiled by Mayne
(2001) also indicate values lower than 50 for very soft to soft clays.
The self-boring pressuremeter test (SBPT), which minimizes soil disturbance around the proble,
may be used for determining the following shear moduli: (a) Go from the initial slope of the expanion
curve; (b) Gur from small unload-reload cycles that can be performed during the expansion curve. Gur is
equivalent to Geq in Figure 4.11.

206
09031-02[2].qxd 18/Oct/02 12:13 AM Page 207

Figure 17.10. An approach for estimating the in situ stiffness curve on the basis of the laboratory stiffness curve
and in situ Go.

Reliable measurement of Go with SBPT is difficult, while the measurement of Gur from small
unload-reload cycles is easier. Lo Presti et al. (2001) discuss the realization and interpretation of this
part of SBPT; they also compare Gur with Go values obtained from other tests.

17.5.2 Estimation of the in situ non-linear stress-strain curve


As indicated in Section 11.3 and in Hight & Leroueil (2002*), Go measured in laboratory is often dif-
ferent, generally smaller, than that deduced in situ from seismic tests. Also, the G/Go  f() curve is
available only from laboratory tests. A practical question of major importance is thus to know how a
representative in situ stress-strain curve could be obtained. Ishihara (1996) proposed the following gen-
eral equation:

Gs (l)
Gs (f)  Cr Go (f)
Go (l) (17.9)

where: (a) Gs(f) and Gs(l) are secant shear moduli at a given shear strain, , respectively in the field and
in the laboratory; (b) Go(f) and Go(l) are initial shear moduli determined respectively in the field, from
in situ seismic tests, and from laboratory tests; (c) Cr is a correction factor.
If it is assumed that the G/Go  f() curves have the same shape in situ and in the laboratory, then Cr
is taken equal to 1.0. Figure 17.10 shows such an example; if the continuous line is the laboratory line,
the in situ line becomes the dashed line. Ishihara (1996) proposed Cr values that are dependent on the
quality of samples and shear strain.

17.6 In situ stresses and Ko


Ko, the ratio of the horizontal and vertical effective stresses in the ground, is an important parameter as
it defines the initial stress conditions for any construction project. It is also one of the most difficult
parameters to determine. Several methods have been proposed for determining Ko. Most of the meth-
ods are based on in situ testing but some are based on laboratory testing. Quite often, also, Ko is estab-
lished on the basis of empirical correlations. Soils being inelastic and with non-linear stress-strain
behaviour (see Section 4.3 and Hight & Leroueil, 2002*), it seems most unlikely that a soil specimen,
which has been subjected to total and effective stress changes due to sampling, could reflect in any reli-
able way the stresses it was subjected to prior to sampling. In the following paragraphs, only some in
situ measurement methods and empirical correlations are indicated.
Several methods have been proposed for determining Ko in situ: hydraulic fracturing (Bjerrum &
Andersen, 1972); pushed-in spade-shaped stress cell (Massarch et al., 1975); dilatometer test (Marchetti,
1980); Iowa stepped blade (Handy et al., 1982); self-boring pressuremeter (Baguelin & Jezekel, 1973;

207
09031-02[2].qxd 18/Oct/02 12:13 AM Page 208

Wroth & Hugues, 1973); shear wave velocities (Stokoe et al., 1991). Except for the last one, all these
methods were discussed in detail by Jamiolkowski et al. (1985). Fahey (1998) stated that: “of the in situ
methods, the self-boring pressuremeter test still appears to be the most accurate “direct” method of
determining ho”, even if he questioned the reliability and the accuracy of this approach in stiff clays
and sands.
The method using shear wave velocities is very attractive (see Fioravante et al., 1998). Ko would be
given by the following equation:

1/n
V ⋅ Cs-HV 
K o   s-HH 
 Vs-HV ⋅ Cs-HH  (17.10)

where Cs-HH and Cs-HV are dimensional material constants that reflect the fabric anisotropy of the soil
and n is a stress parameter generally close to 0.125.
Equation 17.10 was established assuming that shear wave velocity is a power function of stresses,
similar to Equation 4.10 for shear modulus. Such relationships, which imply zero shear wave velocity
or shear modulus under zero stress, may be valid for soils that are not microstructured, as the reconsti-
tuted sands tested by Fioravante et al. (1998). For soils that are microstructured, i.e. most natural soils,
Vs or Go may not depend on applied stresses only (see Section 11.2.3), and Equation 17.10 is then not
applicable.
Empirical correlations have been proposed for the estimation of Ko. The most common is:

Ko  Konc OCR (17.11)

where Konc is the Ko value for the soil normally consolidated, OCR is the overconsolidation ratio and
 is a parameter.
Mayne & Kulhawy (1982) found good correlations with measured Ko values with   sin tc. On
the other hand, Lefebvre et al. (1991) and Hamouche et al. (1995a) found  values close to 1.0 in east-
ern Canada clays. There is no clear explanation for that, but other high  values seem to exist in other
sensitive clay deposits, Bothkennar and Fucino in particular (Hamouche et al., 1995a).

18 PROPERTIES ASSOCIATED WITH THE PASSAGE FROM INTACT TO DESTRUCTURED,


POSSIBLY REMOULDED CONDITIONS

18.1 Generalities
It has been shown in Sections 16.2 and 16.3 that the undrained shear strength of remoulded clays and
the compression curve of reconstituted clays can respectively be characterized by the IL–cur relationship
and the intrinsic compression line (ICL). These two lines can be used as references for assessing the
degree of microstructuration of natural clayey materials.
Figure 18.1 shows the cur–IL relationship for remoulded clays (Equation 16.1 and Figure 16.1). Also
shown on the figure is an approximate IL– v relationship assuming a cur/ v ratio equal to 0.3. These
two lines represent the behaviour of the remoulded or completely destructured clay. If the undrained
shear strength cu and the preconsolidation pressure p of the intact soil are plotted at the
liquidity index of the soil in its natural conditions, the degree of microstructuration of the clay can
be assessed. The larger the distance between intact conditions and non-microstructured ones at the
natural liquidity index, the more important are the bonds between particles or aggregates.

18.2 Strength sensitivity


In undrained conditions, i.e. at constant liquidity index, the distance between log cu and log cur at the
natural liquidity index is the logarithm of the sensitivity or strength sensitivity, St (Figure 18.1). As cur
is related to the liquidity index (Equation 16.1), it follows:

St  cu/cur  cu  (IL  0.21)2, with cu in kPa (18.1)

Equation 18.1 means that St, cu and cur are interrelated. If two of these parameters are defined, the
third one is defined too.

208
09031-02[2].qxd 18/Oct/02 12:13 AM Page 209

2.4
Intact clay
lg St cu σp'
2.0
Ar A B

Liquidity index, I L
1.6

1.2 I LD

0.8
c ur (kPa) = 1
0.4 (I L - 0.21)2 c ur
(Eq. 16.1) Remoulded clay σ'v
0
2 3
10-1 1 10 10 10
c u or σ'v (kPa)

Figure 18.1. Structure and destructuration of clays (from Leroueil et al., 1983a).

It is worth mentioning that the sensitivity determined on the basis of the fall-cone penetrometer
(Equation 18.1) is generally larger than the value obtained with field vane test, probably because full
remoulding cannot be obtained with this latter test.

18.3 Void index in situ, yield stress ratio and stress sensitivity
Using the void index Iv as a normalizing parameter, it is possible to compare the sedimentation com-
pression curve followed in normally consolidated clay deposit with the ICL defined in Section 16.3. The
in situ void ratio eo is then used in Equation 16.2 and the corresponding void index is termed Ivo. For a
normally consolidated clay, the ratio of the vertical effective stress vo and the vertical effective stress
*ve on the ICL at Ivo reflects the microstructure of the soil developed during its deposition. For most
normally consolidated clay deposits (but not all), Burland (1990) found that the sedimentation com-
pression curves lie on the same line that he called the sedimentation compression line (SCL), (Figure 18.2).
The ratio of the vertical stresses on SCL and ICL at a given void index, vo/ *ve is often equal to 4–5.
On an Iv– v diagram, it is also possible to draw the compression curve of a natural clayey soil as
obtained in an oedometer test. An example from Sault-Ste-Marie, near Chicago, is shown in
Figure 18.3. In addition to vo/ *ve it is also possible to define the vy/ *ve ratio (where vy is the
vertical yield stress or preconsolidation pressure), termed yield stress ratio by Burland (1990), which is
a measure of the increase of resistance to compression caused by the natural microstructure. Its mean-
ing is very similar to that of sensitivity expressed in terms of undrained shear strength and the yield
stress ratio has been termed stress sensitivity S by Cotecchia & Chandler (2000). For the case shown
in Figure 18.3, S is equal to about 15. Evaluating strength and stress sensitivities for Louiseville clay,
Leroueil et al. (2002*) found St  20 and S  11. In fact, it is logical to find St  S as sedimented or
reconstituted clays generally show strength sensitivities larger than 1.0 when on their compression line.
This is illustrated in Figure 18.4 where are shown as a function of IL the remoulded strength cur, the
compression curves obtained on the clay sedimented in water with salinity of 0.2 g/l and 35 g/l, and, at
the natural liquidity index ILo, cu and p of the intact clay. It can be seen that, at the in situ void ratio (at
ILo  2.4), St  140 whereas S  10 or 40, depending if the clay sedimented in 0.2 g/l or 35 g/l water
salinity.

18.4 Metastability and stiffness sensitivity


Other ways for assessing the degree of microstructure have been considered. Soga & Mitchell (1996)
proposed the degree of metastability (or metastability index), defined as the difference between the
liquidity index of the soil at the preconsolidation pressure and the liquidity index of the reconstituted
soil at the same effective stress.

209
09031-02[2].qxd 18/Oct/02 12:13 AM Page 210

LL LL LL
Oslo fjord 98 Älvängen 95 S.Joaquin 64
A-33 80 Shellhaven 82 Milazzo 62
5 A-31 63 Avonmouth 71 Baku 40
B-87 58 Drammen 54
C-18 46 Grangemouth 41
Drammen 38
x Detroit 28
4

e*100 – e*1000
e – e*100
3
Sedimentation
compression
Void index I vo =

2 line (SCL)

1
xx
x
xx
0

Intrinsic compression line


–1

–2
–1 2 3 4
10 1 10 10 10 10
σ'vo (kPa)
Figure 18.2. Comparison between the normalized one-dimensional compression curve obtained in laboratory on
reconstituted clays and in situ conditions for many normally consolidated clays (from Burland, 1990).

Figure 18.3. Example of oedometer compression curve in Iv– log v diagram. Sault-Ste-Marie clay (after Burland,
1990).

Shibuya (2000) proposed a similar metastability index based on Go. The Go  IL relationship being
defined under one-dimensional loading for both the reconstituted and the intact soil, the metastability
index is the difference between the liquidity index on the undisturbed line, close to gross yield, and the
liquidity index of the reconstituted soil at the same Go value.
Alternatively, the Go–IL (or Go–e) relationship established for the reconstituted and the intact soil
could be used to define a “stiffness sensitivity”, Ss. Stiffness sensitivity would be the ratio of Go value
of the intact soil close to gross yield and Go of the reconstituted soil at the same void ratio. Referring to

210
09031-02[2].qxd 18/Oct/02 12:13 AM Page 211

3.0

Sedimented clay

Intact clay
2.5
C ur I
Cu σ'p
Destructured
clay

2.0 Cu A
Cur A
σ'vA
Remolded clay 0.2 g/ l D
A c = 0.22
I Lc

1.5

Cu B
1.0
Cur B σ'vB
35 g/ l
A c = 0.4

0.5

0
-1 -2
10 1 10 10
Cu, Cu , σ'v (kPa)
r

Figure 18.4. Undrained shear strength-liquidity index and effective stress-liquidity index relations for Grande-
Baleine clay (after Locat & Lefebvre, 1982; from Leroueil et al., 1985b).

Figure 11.20, the stress sensitivity and the stiffness sensitivity of the Jonquière clay would be approxi-
mately equal, around 1.6. For Louiseville clay, strength and stress sensitivities were found equal to 20
and 11 relatively at a depth of 9.2 m (Leroueil et al., 2002*). Stiffness and stress sensitivities could also
have been evaluated at a depth of 11.5 m (Shibuya, 2000; Leroueil et al., 2002*); their value would be
approximately 4 at that depth.

18.5 Compression and swelling indices


Many empirical relationships have been established in the past between compression index and physi-
cal properties of soils, mainly liquid limit and void ratio. Some, defined for non- or slightly microstruc-
tured soils, have been presented in Section 16.4. The relation Cc  0.39e (Equation 16.8) proposed by
Nagaraj & Srinivasa Murthy (1986) for such soils is shown in Figure 18.5a (dashed line NS). For
microstructured soils however, as shown in Figure 11.2 for Mexico City and Grande Baleine clays, and
also in Figure 18.3 for Sault-Ste-Marie clay, the compression index observed just after gross yielding is
larger than that of the reconstituted soil. Examining data from eastern Canada and from other parts of
the world, Leroueil et al. (1983a) came to the conclusion that Cc could be approximately defined as a
function of void ratio and strength sensitivity, as indicated in Figure 18.5a for the range of consolida-
tion pressures between p and 1.4 p. The constant St curves shown in Figure 18.5a could probably be
graduated in term of S or Ss as well.
Figure 18.5b shows the Cc–eo relationship obtained for Juturnaiba organic clays which have void
ratios varying from 1.2 to 9 (Coutinho & Lacerda, 1987). In comparison with Figure 18.5a, this rela-
tionship would correspond to a sensitivity of 3 to 6, which is reasonable for this type of soil. For Mexico
City clay, Mesri et al. (1975) found Cc  1.1e (Figure 18.5b). Figure 18.5c shows relationships obtained
for several residual soils.
Figure 18.6 shows the relationship between the swelling strain index (Css   v/log v ) and CF.wL
displayed by Terzaghi et al. (1996) for reconstituted clayey soils. It can be seen that it typically increases

211
09031-02[2].qxd 18/Oct/02 12:13 AM Page 212

Figure 18.5. Relations between Cc and eo: (a) clays in general (from Leroueil et al., 1983a); (b) Mexico City clays
(Mesri et al., 1975) and Juturnaiba organic clays (Coutinho & Lacerda, 1987); (c) Residual soils (Leroueil &
Vaughan, 1990).

Figure 18.6. Swelling strain index Css as a function of clay fraction, CF, and liquid limit wL.

212
09031-02[2].qxd 18/Oct/02 12:13 AM Page 213

from 0.01 for a CF.wL value of 5 to 0.3 for a CF.wL value of 1000. The swelling index of microstruc-
tured soils may be much smaller, as indicated by the arrows on the figure. This is demonstrated in
Figure 11.3 which shows that the swelling index of Laviano clay shale (CF  34%; wL  48%; car-
bonate content of 19%) increases as destructuration by compression progresses. The corresponding
evolution of the swelling strain index with the loading-unloading cycles is indicated on Figure 18.6; it
increases from about 0.008 for the undisturbed material (U on the figure) to 0.05 after 6 cycles of load-
ing-unloading, and to 0.08 after remoulding (R on the figure). The swelling index (e/log v))
defined on the reconstituted soil was termed the intrinsic swelling index by Burland (1990) and
expressed by C*.s The ratio C*/C
s s is also an indication of the degree of microstructuration and was
defined by Schmertmann (1969) as the swell sensitivity.

18.6 Effective strength parameters


Burland et al. (1996) extended the concepts of Hvorslev (1937) and critical state soil mechanics
described in Section 2 and the concepts of intrinsic behaviour proposed by Burland (1990) to assess the
influence of microstructure on the strength parameters of natural clays. For that purpose, they suggested
normalizing strength envelopes with respect to the vertical stress on the intrinsic compression line
(ICL) at the void ratio of interest, *ve. The failure envelope for reconstituted clay can then be described
by the normalized Mohr-Coulomb equation:

 
 *  tan e*
(18.2)
*ve *ve

where * is the intrinsic Hvorslev cohesion and *e is the Hvorslev true friction angle.
Burland et al. (1996) studied four different natural stiff clays. Figure 18.7 shows the normalized
Hvorslev strength envelopes for the intact and reconstituted Vallerica clay. The influence of microstruc-
ture can be estimated by the distance between the two envelopes. An interesting conclusion of Burland
et al. (1996) is that e value for intact clay is similar to the reconstituted value and that microstructure
mostly influences effective cohesion.

18.7 Influence of leaching on the characteristics of marine clays


Due to the combined effect of glacio-isostatic rebound and change in sea level, some clay deposits that
were sedimented in a marine environment are now above the sea level. This is particularly the case in
Canada, Russia and Scandinavia. Associated with this new situation is a flow of fresh water through the
sediments and progressive leaching of the marine clay. As indicated, in particular by Bjerrum (1954,
1967), leaching has several effects on clay characteristics: the plasticity of the clay decreases;
the undrained shear strength may slightly decrease, but at an essentially constant water content; as the
cur–e and v–e curves of the leached soil, remoulded or destructured, are at void ratios smaller than the

Figure 18.7. Comparison of intact and intrinsic Hvorslev strength envelopes for Vallericca clay (after Burland
et al., 1996).

213
09031-02[2].qxd 18/Oct/02 12:13 AM Page 214

Figure 18.8. Influence of leaching on the characteristics of marine clays.

corresponding curves for the clay sedimented in marine environment, it follows that the sensitivity and
compression index of the leached clay are larger than those of the unleached material. Refering to
Figure 18.8, St2 and Cc2, associated with leached clay, are respectively larger than St1 and Cc1 associated
with unleached clay.

19 HYDRAULIC PROPERTIES

19.1 Generalities
The flow rate of a liquid, q, through an area A of a porous media, due to a hydraulic gradient i, is con-
trolled by Darcy’s law:

q  kiA (19.1)

where k is the hydraulic conductivity of the soil.


According to Poiseuille’s law, the flow through capillaries would be characterized by a hydraulic con-
ductivity proportional to the square of the radius of the tubes, or an equivalent hydraulic radius reflect-
ing the flow channels in the soil. According to Kozeny-Carman equation, the hydraulic conductivity of
a soil would be proportional to Ds2, where Ds is a characteristic grain size, and e3/(1  e) (see, for exam-
ple, Mitchell (1993) for details). Though Poiseuille’s law and Kozeny-Carman equation are better able
to represent the flow through cohesionless soils than through clayey soils, they indicate the main factors
controlling the flow of liquid through porous materials.
For loose and uniform sands, Hazen (1911) proposed the following equation relating the hydraulic
conductivity and d10, the particle diameter corresponding to 10% passing:

k (m/s)  0.01d10 (mm)2 (19.2)

The validity of Equation 19.2 has been confirmed by Loudon (1952). Sherard et al. (1984) and
Kenney et al. (1985) proposed similar relationships, but referring to d15 and d5 respectively.

214
09031-02[2].qxd 18/Oct/02 12:13 AM Page 215

Figure 19.1. Tentative hydraulic conductivity assignment to main classes of soil (modified from Manassero, 1994
and personal communication).

Figure 19.2. Hydraulic conductivity of LG-2 till as a function of the compaction degree of saturation at a void ratio
of 0.25 (from Watabe et al., 2000).

The hydraulic conductivity of soils thus varies tremendously from soil to soil. Figure 19.1 from
Manassero (1994) provides an indication of the hydraulic conductivity of soils; it typically varies by 10
orders of magnitude from gravel to clays. In fact, the hydraulic conductivity of clayey materials can
even be smaller than 1010 m/s: Mesri et al. (1975) and Diaz-Rodriguez (2002*) showed a decrease in
the hydraulic conductivity of Mexico City clay from 108 m/s at a void ratio of 15 to 1011 at a void
ratio of 2; Chandler et al. (1990) found k values between 1010 and 1011 m/s for London clay at
Bradwell; Picarelli et al. (2000; 2002*) found k values between 1010 and 1011 m/s for Bisaccia clay
shale; Neuzil (1993) estimated the hydraulic conductivity of Pierre shale (South Dakota) to be from
1013 to 1014 m/s.
For a given soil at a given void ratio, the hydraulic conductivity depends on the size of the flow chan-
nels and may thus be influenced by soil fabric. This can be illustrated by test results presented by
Watabe et al. (2000). Figure 14.6 shows soil water characteristic curves for a non-plastic glacial till
compacted wet or dry of optimum. It can be seen that the air entry pressure, which is in the order 30 kPa
when the soil is compacted on the wet side of optimum, is approximately ten times larger than when the
soil is compacted on the dry side of optimum. This indicates that the pore radii of the wet compacted
soil are typically 10 times smaller than the pore radii of the dry compacted soil (see Equation 14.1).
Twelve permeability tests have been performed in saturated conditions on this till, after compaction
under different conditions. Figure 19.2 presents the measured saturated hydraulic conductivity as a
function of the compaction degree of saturation at a void ratio of 0.25. It can be seen that the hydraulic
conductivity changes by a factor 100 when the compaction degree of saturation decreases from values
higher than the value at the optimum (75%) to values that are smaller (i.e. typically the square of
the ratio of pore radii of the dry compacted and wet compacted soil (102)). Viana da Fonseca (2002*)
also indicates that the hydraulic conductivity of undisturbed Porto residual soil can be several orders of

215
09031-02[2].qxd 18/Oct/02 12:13 AM Page 216

magnitude larger than the hydraulic conductivity of the same soil reconstituted at the same unit weight
and water content.
The hydraulic conductivity also depends on the characteristics of the permeant liquid:

p
kK (19.3)


where K is the intrinsic permeability of the soil, and p the unit weight and  the viscosity of the permeant.
The hydraulic conductivity of rock masses is mostly controlled by discontinuities. In fissured clayey
soils, the hydraulic conductivity of the mass depends if the fissures are open or not.

19.2 Vertical and horizontal hydraulic conductivities in natural conditions


19.2.1 Permeability anisotropy
Anisotropy in the permeability of soils is characterized by rk  kh/kv, where kh and kv are the hydraulic
conductivities in the horizontal and vertical directions respectively.
Figure 19.3 shows the horizontal and vertical hydraulic conductivities of soft clays at their natural
void ratio, kho and kvo. Several remarks can be made: (a) for most natural soft clays, the vertical
hydraulic conductivity is between 3  1010 and 5  109 m/s; (b) in marine homogeneous clays, rk is
generally smaller than 1.2; (c) rk is larger in stratified deposits or varved clay deposits but generally
seems to be lower than 10.
In stiff geologically overconsolidated clays of marine origin, the anisotropy is larger than in soft
marine clays, probably due to a preferential orientation of clay particles in the horizontal direction when
the overburden pressure was increasing. Chandler et al. (1990) found a rk value in the order of 2 in
London clay; Cotecchia (2002*) indicates a rk value of about 3 in Pappadai clay.
Permeability anisotropy is generally small in granular soils, although this may depend on the stratig-
raphy and also on the shape of the particles. Comparing horizontal and vertical hydraulic conductivities
measured in pairs of undisturbed samples of gravelly soils retrieved by freezing technique, Hatanaka
et al. (2001) found a rk value varying from 1.1 and 1.7. On the other hand, Jamiolkowski & Lo Presti
(2002*) report rk values deduced from pumping tests in Messina sand and gravel of at least 5. This high
value could be associated with the presence of layers with different hydraulic characteristics, but also
with gravel grains that are mainly disk-shaped and could be preferentially oriented.

-8
10
Marine clays 1 - Bothkennar
Eastern Canada clays (Leroueil et al., 1992 ;
Hong Kong Hight et al., 2002b)
Swedish clays
2 - New Liskeard varned clay
(Chan & Kenney, 1973)
k vo (m/s)

1
10 4 3 - New-Jersey varved clay
-9
10 (Saxena et al., 1978)
2
4 - Connecticut varved clay
(DeGroot & Lutenegger, 2002*)
3
5 - Atchafalaya
5 (Authors' files)

-10
10 -10 -9 -8
10 10 10
k ho (m/s)
Figure 19.3. Permeability anisotropy in soft marine clays (from Leroueil et al., 1990).

216
09031-02[2].qxd 18/Oct/02 12:13 AM Page 217

For considerations on the mass hydraulic conductivity of layered soils, the reader is referred to
Kenney (1963).

19.2.2 Hydraulic conductivity measurement


In the laboratory, the hydraulic conductivity can be measured in flexible wall permeameter (e.g. triax-
ial cell) or in rigid wall permeameter (e.g. oedometer). It can also be measured in special apparatuses
(radial-flow cell (Leroueil et al., 1990); cubic cell (Chan & Kenney, 1973). The hydraulic conductivity
can be directly measured by application of a constant hydraulic head or a variable hydraulic head
through a soil specimen. It can also be obtained from continuous Controlled Gradient (CG) or CRS
oedometer tests by using the following equation (Lowe et al., 1969; Smith & Wahls, 1969):

∂H Hw
k (19.4)
∂t 2b

where H is the current height of the specimen and ub is the excess pore pressure at the base of the spec-
imen. This equation is based on the assumptions that k is constant within the specimen and that the pore
pressure isochrone is parabolic in shape. These assumptions are not truly representative, in particular
for stresses close to the preconsolidation pressure. However, experience indicates that reasonable k val-
ues can be obtained from continuous oedometer tests provided the values near p are disregarded
(Tavenas & Leroueil, 1987). Figure 19.11 shows some e–log kv relationships obtained from CRS tests.
Perrmeability tests can be performed in situ, in boreholes, in pushed-in piezometers (including the
BAT system proposed by Torstensson, 1984), or in self-boring permeameter. Tests in boreholes are
unreliable mainly because of the uncertainties in the shape of the cavity through which water is injected.
A pushed-in piezometer displaces and remould the soil around it, which then consolidates to a lower
void ratio. In addition, there can be clogging of the porous element by fine clay particles. As a result,
pushed-in piezometers generally underestimate the k values and are not recommended for measuring
the hydraulic conductivity of clays (Tavenas et al., 1986). The self-boring permeameter, which essen-
tially eliminates the difficulties associated with pushed-in piezometers, seems to give reliable results.
Figure 19.4 compares hydraulic conductivity values measured by pushed-in piezometers and the self-
boring permeameter at 5 different sites. On relatively homogeneous sites such as Louiseville (see
Leroueil et al., 2002*), the ratio between the 2 values is about 2; on the other hand, in the more strati-
fied St-Alban clay deposit, the ratio is larger than 10. With special tests, Tavenas et al. (1986) were also
able to define the relative contribution to the reduction in k due to remoulding of the clay and to clog-
ging of the porous element (Figure 19.4).
The influence of the geometry of the porous element on the measured hydraulic conductivity was
examined by Diène (1989) and Bouclin (1990). These authors performed permeability tests with porous
elements with different length over diameter (l/d) ratios. Typical results are shown in Figure 19.5. It can
be seen that pushed-in piezometer gives a hydraulic conductivity that regularly increases with (l/d)
whereas the self-boring permeameter gives higher values than the pushed-in piezometer.

Figure 19.4. Results of permeability tests with different probes illustrating the effects of remoulding and clogging
(after Tavenas et al., 1986).

217
09031-02[2].qxd 18/Oct/02 12:13 AM Page 218

Figure 19.5. Hydraulic conductivity of Saint-Esprit clay as a function of the geometry of the porous element (after
Bouclin, 1990).

Diène (1989) analysed the problem and showed numerically that the radial zone of soil influenced by
permeability test carried out in situ was largely dependent on the (l/d) ratio of the porous element. The
diameter of the influenced zone typically increases by a factor of 4 when (l/d) increases from 1 to 15.
As a result, for small (l/d) ratios (possibly 1 or less), the region of soil affected by the permeability test
is mostly in the region that has been disturbed by the penetration of the piezometer and has reconsoli-
dated. Underestimation of the hydraulic conductivity is thus expected from pushed-in-place piezome-
ters with such geometry. As (l/d) increases, the influence of the disturbed zone decreases and the
obtained hydraulic conductivity becomes closer to that of the intact soil. It is thought that the effect of
the disturbed zone becomes negligible for (l/d) values larger than about 12.
The BAT permeameter developed by Torstensson (1984) and Torstensson & Petsonk (1986) is a
pushed-in-place piezometer with a 30 mm diameter and 40 mm high filter tip. Considering its geome-
try (l/d  1.33), it is expected that values of k will be underestimated with this equipment. This is the
conclusion that was reached at Bothkennar (Leroueil et al., 1992) and at Onsoy (Lunne et al., 2002b*)
where BAT system gave hydraulic conductivities generally smaller than other permeability tests.
From Equation 20.3, it is theoretically possible to deduce k values from coefficients of consolidation.
However, considering the uncertainty on this latter parameter (see section 20.2), this approach is not
recommended.

19.3 Hydraulic conductivity-void ratio relationships


The relation between hydraulic conductivity and void ratio (or porosity) has been the object of numer-
ous investigations. Figure 19.6 shows a relationship for chalk. For clays, it appears that there is no
general relationship between the two parameters. However, Tavenas et al. (1983a) showed that e–log kv
relationships for clays are approximately linear for most clays for vertical strains up to about 20%, i.e.
in the range usually encountered in practice (Figure 19.7). These relationships can then be characterized
by the hydraulic conductivity change index Ck defined as:
Ck  e/log k (19.5)

Tavenas et al. (1983a) also observed that Ck is simply related to the initial void ratio, eo, of the clay
(Figure 19.8). As a first approximation, they proposed:
Ck  0.5eo (19.6)

Since 1983, Equation 19.6 has generally been confirmed for soft clays and also for London clay
(Chandler et al., 1990). However, as indicated on Figure 19.8 and by Cotecchia (2002*) who found a Ck
value of 0.29 for a natural void ratio of 0.83 for Pappadai clay, the ratio Ck/eo can be slightly smaller
than 0.5 for low void ratios.

218
09031-02[2].qxd 18/Oct/02 12:13 AM Page 219

Figure 19.6. Permeability of chalk (after Scholle, 1977; ref. by Addis, 1987).

Figure 19.7. Variation of the vertical hydraulic conductivity with void ratio (after Tavenas et al., 1983a).

Al-Tabbaa & Wood (1987) and Leroueil et al. (1990) investigated the variation of both vertical and
horizontal hydraulic conductivities with void ratio when the soil is vertically compressed. Al-Tabbaa &
Wood (1987) observed on kaolin an increase in rk from 1.0 at liquidity index of 1.8 to 2.7 at liquidity
index of zero. On natural clays from Sweden and eastern Canada, which are mostly made up of rock
flour, Leroueil et al. (1990) observed a smaller increase in rk as the liquidity index was decreasing
(Figure 19.9). rk values obtained in natural London clay and Pappadai clay are also indicated on the fig-
ure, at a liquidity index close to zero; they are between 1.4 and 2.7 for London clay and about 3 for
Pappadai clay.

219
09031-02[2].qxd 18/Oct/02 12:13 AM Page 220

Figure 19.8. Relationship between the permeability change indices Ckv and Ckh and the initial void ratio (from
Leroueil et al., 1990).

Figure 19.9. Permeability anisotropy of marine clays as a function of the liquidity index (after Leroueil et al.,
1990).

220
09031-02[2].qxd 18/Oct/02 12:13 AM Page 221

Figure 19.10. Matagami varved clay (8.8 m): (a) water content profile; (b) variation of hydraulic conductivity with
void ratio during vertical compression (after Leroueil et al., 1990).

Leroueil et al. (1990) also investigated the variation of rk with void ratio in varved clays loaded ver-
tically. The samples were from Matagami, about 800 km north of Montreal, in an area that was covered
about 8000 years ago by the Barlow-Ojibway Lake. The variation of the vertical hydraulic conductivity
with void ratio was studied in a triaxial cell where the specimen was compressed in steps under essen-
tially one-dimensional conditions. The variation of the horizontal hydraulic conductivity with void ratio
was studied in a radial-flow permeameter. Figure 19.10 shows a typical result. Figure 19.10a shows the
detailed water-content profile of the intact specimen tested in the radial-flow cell. Figure 19.10b shows
the e–log k relationships deduced from the 2 tests. kv decreases with void ratio with a Ck value close to
0.5 eo. The reduction of horizontal hydraulic conductivity with vertical compression is much smaller.
This is mostly due to the fact that clayey laminae are more compressible than silty laminae, reducing
vertical hydraulic conductivity more significantly than horizontal hydraulic conductivity. A practical
implication is that anisotropy in the hydraulic conductivity of varved clays (and probably of laminated
clays in general) significantly increases when they are vertically compressed.

19.4 Other factors influencing hydraulic conductivity


When considering hydraulic conductivity, other important factors are the pore fluid chemistry and
hydraulic characteristics. There are three aspects: (a) the pore fluid chemistry may modify soil fabric;
(b) the pore fluid may influence the characteristics of the “electric double layer”, which may obstruct
the flow channels when it increases in thickness; and (c) as indicated in Equation 19.3, the hydraulic
conductivity is influenced by the viscosity of the pore fluid.
The influence of viscosity on hydraulic conductivity can be observed when permeability tests are
performed at different temperatures. Figure 19.11 shows an example where kv has been determined in
an oedometer cell at different void ratios at a temperature of 20°C and in CRS tests at temperatures of
10 and 50°C. The ratios between the hydraulic conductivities measured at the different temperatures at
the same void ratio are essentially constant. In particular, kv(50°C)/kv(10°C)  2.2, which corresponds
to the change in water viscosity between 10 and 50°C.
The use of distilled water in permeability tests may have an adverse effect on the measured hydraulic
conductivity. Figure 19.12 which shows 3 examples of clays from Sweden and Canada subjected to a
flow of distilled water, confirms the decrease in k as the salinity of the outflow water decreases and the
thickness of the “electric double layer” increases. The results, however, suggest that the effect becomes
significant only when the volume of permeated distilled water exceeds about half the pore volume of
the specimen.

221
09031-02[2].qxd 18/Oct/02 12:13 AM Page 222

Figure 19.11. Variation of hydraulic conductivity with void ratio and temperature of Saint-Roch-de-l’Achigan clay
(after Marques et al., 2002).

Figure 19.12. Effect of distilled water on the results of permeability tests on three different clays (from Tavenas &
Leroueil, 1987).

222
09031-02[2].qxd 18/Oct/02 12:13 AM Page 223

Figure 19.13. Effect of chemistry of the permeant on the hydraulic conductivity (from Fernandez & Quigley,
1985).

Figure 19.13 shows the very dramatic change in hydraulic conductivity that may be observed when
chemical fluids and in particular organic compounds are permeated through clays.

19.5 Hydraulic conductivity of unsaturated soils


The hydraulic conductivity of unsaturated soils is generally characterized by the relative hydraulic con-
ductivity, kwr, defined as the ratio of the hydraulic conductivity in the unsaturated condition, kwus, to
that in the saturated condition, kws. The relative conductivity is often evaluated by a power function of
the effective degree of saturation Sre (Brooks & Corey, 1964; Fredlund & Rahardjo, 1993):

k wus
k wr   Srek (19.7)
k ws

where Sre  (Sr  Srr)/(1  Srr), with Srr being the residual degree of saturation (see Section 14.1 and
Figure 14.1), and k is a soil parameter.
Fredlund et al. (2001) examined approximately 300 sets of permeability data obtained on a variety of
soils. They found k values ranging from less than 1.0 to more than 6.0. This is partly due to difficul-
ties associated with the measurement of the parameters involved, but there are also other factors.
Finding that the mean value for k is 2.37 for sands and 4.34 for clays, Fredlund et al. (2001) concluded
that “there was a definite trend towards a larger k value for soils with higher plasticity.” The parame-
ter k can also be deduced from the soil-water characteristic curve (SWCC; see Section 14.1) of the soil
(Marshall, 1958; Fredlund et al., 1994). Applying this approach to the compacted LG-2 non-plastic gla-
cial till (see SWCCs in Figure 14.6), Le Bihan & Leroueil (2001) found k values between 5 and 7,
values that were also confirmed experimentally. In general, k increases from a value of about 3 for uni-
form soils to larger values in well-graded materials.

20 CHARACTERISTICS OF CONSOLIDATION

When a saturated soil mass is subjected to changes in boundary conditions, generally it has a tendency
to change volume, and as a consequence, expel or absorb water. This process takes time before the soil
mass can reach a steady state. Two stages are generally considered: (a) primary consolidation during
which the rate of settlement or strain is controlled by the ability of the soil mass to expel water, which

223
09031-02[2].qxd 18/Oct/02 12:13 AM Page 224

is governed by Darcy’s law and hydraulic conductivity; (b) secondary consolidation during which the
rate of strain is controlled only by the viscosity of the soil skeleton.

20.1 Coefficient of consolidation


Before discussing the coefficient of consolidation, its measurement and how the measured values can
be used in in situ conditions, it is of interest to recall that the basic and most general equation for one-
dimensional consolidation can be written (Berry & Poskitt, 1972) as:

∂ v 1  eo ∂  k ∂u 
   (20.1)
∂t w ∂z  1  e ∂z 

where v  vertical strain, z  depth related to initial thickness, eo  initial void ratio, e  present
void ratio, k  hydraulic conductivity, and u  excess pore-water pressure.
The assumptions required to establish Equation 20.1 are that Darcy’s law is valid and the soil is sat-
urated. From Equation 20.1, it is worth noting that at a given time, the strain rate v   v/t depends
essentially on the pore pressure distribution and the vertical hydraulic conductivity at that time, and
very little on soil compressibility (through (1  eo) and (1  e)).
In the simple case where the applied total stress, the vertical hydraulic conductivity kv, and the defor-
mation modulus M  d v/d v, are all constant, the equation controlling consolidation takes the well-
known form:

∂u ∂2 u
 cv 2 (20.2)
∂t ∂z

with

kM k v (1  eo )
cv   (20.3)
w 0.434wCc

where M is the deformation modulus defined as d v/d v in one-dimensional compression.


The coefficient of consolidation cv is thus not a soil parameter as such but a function of the defor-
mation modulus M and the vertical hydraulic conductivity kv. From Equation 20.3, it can be seen that
the coefficient of consolidation will be larger in the overconsolidated domain than in the normally con-
solidated domain, where the hydraulic conductivity decreases with void ratio and the deformation mod-
ulus is lower. Even in this overconsolidated domain, as the bulk modulus is generally not constant, cv is
not constant either.
When cv is constant during consolidation, the degree of consolidation expressed in terms of pore
pressure dissipation is equal to the degree of consolidation expressed in terms of settlement, and
Terzaghi’s theory perfectly applies for this one-dimensional consolidation case.
In the special case of soil treatment with vertical drains under an embankment, deformation is still
mainly vertical but water flow is mainly horizontal. The coefficient of consolidation then combines ver-
tical deformation modulus M and horizontal hydraulic conductivity kh in Equation 20.3.
Among the hypotheses of Terzaghi’s theory, those of constant hydraulic conductivity and of constant
deformation modulus are questionable. k varies with void ratio (Section 19.3); constant M means that
the effective stress-strain relationship for the soil is linear and not affected by time or strain rate, which
is seldom the case. As a consequence, Terzaghi’s theory should generally be seen as an approximation
of real behaviour.
The change in coefficient of consolidation when the soil becomes normally consolidated has conse-
quences on the consolidation process when the soil is loaded from an overconsolidated state to a nor-
mally consolidated state. Mesri & Rokhsar (1974) and Tavenas et al. (1979) had simulated such cases
(see Figure 20.1) and showed that under this condition, a step is present in the plot of the degree of con-
solidation in terms of pore pressure versus logarithm of time corresponding to the passage of the pre-
consolidation pressure. On the other hand, the plot of degree of settlement versus logarithm of time
shows a typical S-shape close to the Terzaghi curve associated with the final value of the coefficient of
consolidation, i.e. the lowest value (Figure 20.1).

224
09031-02[2].qxd 18/Oct/02 12:13 AM Page 225

Figure 20.1. Results of CONMULT program for a stress increment spanning the preconsolidation pressure, and
comparison with Terzaghi theory (from Tavenas et al., 1979).

Figure 20.2. Consolidation, experimental (1) and theoretical (2 to 4) curves.

20.2 Evaluation of coefficients of consolidation


20.2.1 In the laboratory using oedometer test results
– Conventional incremental-loading oedometer tests. Several methods have been proposed for deter-
mining coefficient of consolidation from the consolidation curve obtained from one-stage loading:
Casagrande (log t) method; Taylor (兹t) method; rectangular hyperbolic fitting method (Sridharan &
Sreepada Rao, 1981); etc. They are all based on the assumption that the consolidation curve perfectly
fits Terzaghi’s theory. As previously indicated, this is seldom the case. Le Bihan (2002) examined the
consequences on a typical case (curve (1) in Figure 20.2). The Casagrande (log t) method is based on
t50, the time necessary for reaching a degree of consolidation of 50%. If the deduced cv value, equal
to 5.0  109 m2/s in this case, is used to establish the theoretical degree of consolidation-time

225
09031-02[2].qxd 18/Oct/02 12:13 AM Page 226

relationship, curve (2) is obtained, which coincides with the experimental curve at a degree of con-
solidation of 50% but not anywhere else. The cv value deduced from Taylor’s method was found equal
to 8.3  109 m2/s. The corresponding theoretical degree of consolidation-time relationship is curve
(3), which gives good agreement with the observed behaviour in the early stages of consolidation, but
diverges considerably afterwards. Finally, cv (calc.) has been calculated on the basis of Equation 20.3,
with M equal to the average modulus of deformation over the loading step and directly measured
hydraulic conductivity. That gave a cv value equal to 1.3  108 m2/s. The corresponding theoretical
degree of consolidation-time relationship is curve (4). This trend of behaviour is typical of what is
generally observed (Leroueil, 1985b; Olson, 1985), with:

cv (log t)  cv (兹苵t )  cv (calc) (20.4)

The main reason for this set of differences seems to come from the effect of strain rate on the com-
pression curve followed during the test. If we refer to Figure 9.6a, the effective stress-strain curve fol-
lowed a-b-c during one stage-loading, with high deformation modulus between a and b (Mab) and a
lower one between b and c (Mbc). It is different from the curve assumed by joining the end-of-loading
points, a-d-c, and characterized by an average deformation modulus Madc. As previously indicated
(Section 20.1, Figure 20.1), the consolidation curve, defined in term of settlement, for a soil with a
bi-linear compression curve is mostly associated with the coefficient of consolidation corresponding
to the lowest deformation modulus. Referring again to Figure 9.6a, the consolidation curve followed
during the loading stage a to c is thus mostly associated with Mbc, with the result that the coefficient
of consolidation best representing the observed behaviour (cv (log t)) is smaller than cv calculated
assuming an average deformation modulus Madc.
– Continuous loading oedometer tests. CRS and controlled Gradient oedometer tests can also be used
to determine cv as a function of strain. The hydraulic conductivity given by Equation 19.4 at a given
time is then combined with the deformation modulus M calculated at the same time over a time incre-
ment t for calculating cv from Equation 20.3. The obtained cv values are close to the cv (calc) indi-
cated in the previous paragraphs and Equation 20.4.
– Laboratory vs in situ coefficients of consolidation. Coefficients of consolidation have also been
deduced from observations of settlements under embankments and compared with values deduced
from laboratory tests, generally with the Casagrande (log t) method. Leroueil (1988) compiled such
data for 16 sites and obtained cv (in situ)/cv (lab) ratios varying between 3 and 200. When the two
extreme high values of 200 and 70 are neglected, the average is 20. Almeida and Marques (2002*)
report similar ratio values ranging between 20 and 30 for Sarapui clay deposit.
There are several reasons for explaining this discrepancy:
• The cv (lab) values considered by Leroueil (1988) had generally been determined by the Casagrande
(log t) method that underestimates cv in comparison with the value calculated by the Taylor (兹苵t)
method.
• As explained in Section 19.2, the hydraulic conductivity measured in laboratory using small speci-
mens may underestimate the hydraulic conductivity of the soil mass, particularly in heterogeneous or
stratified deposits.
• The neglect of the two- or three-dimensional aspect of field problems and of permeability anisotropy.
In agreement with Simons (1975), the authors thus recommend for practical purposes to calculate cv
on the basis of permeability tests carried out in the laboratory on large specimens, or in situ.

20.2.2 In situ, from piezocone dissipation tests and SBP holding tests
In the field, coefficient of consolidation can be determined from SBP holding tests, from piezocone
dissipation tests, or from calculations based on small strain shear modulus and in situ hydraulic
conductivity.
The holding test proposed by Clarke et al. (1979) consists of inflating the self-boring pressuremeter
up to about 10% radial strain and then maintaining a constant deformation while measuring the decay
of the pore-water pressure and the variation of cell pressure, i.e. the change in effective stress. The
analysis of these observations provides a coefficient of consolidation.
Many approaches have been developed to model the change in pore pressure with time after piezo-
cone penetration in order to obtain a coefficient of consolidation (Levadoux & Baligh, 1986; Teh &
Houlsby, 1991; Burns & Mayne, 1998). Most of the proposed solutions indicate a monotonic decay of
pore-water pressure with time. However, an initial increase in pore pressure before the decrease towards

226
09031-02[2].qxd 18/Oct/02 12:13 AM Page 227

Figure 20.3. Coefficient of consolidation profiles (from Hamouche et al., 1995b; Leroueil et al., 2002*).

equilibrium conditions is often observed, in particular in overconsolidated materials (see examples in


Burns & Mayne (1998) and Leroueil et al. (2002*)). The solution proposed by Burns & Mayne (1998)
allows the simulation of both cases, depending on the overconsolidation ratio.
A very good practical question is: How representative is the coefficient of consolidation determined
from in situ tests? Numerous papers have been written on this topic, and the authors do not think that
there is a definite answer yet. As they do not have the intention of reviewing this literature, they will
rather show some results that reflect their own experience.
On Louiseville site, Hamouche et al. (1995b) and Leroueil et al. (2002*) determined cv or ch in both
the laboratory and in situ. In the laboratory, cv has been determined from oedometer consolidation tests,
using the Casagrande (log t) method. The values obtained in the normally consolidated range were
between 7  109 and 1.4  108 m2/s, thus typically equal to 108 m2/s. In the overconsolidation
range, due to creep, it was generally difficult to define the end of primary consolidation. In situ, the hor-
izontal coefficient of consolidation was determined from SBP holding tests and from piezocone dissi-
pation tests, using the interpretation method proposed by Teh & Houlsby (1991). The results are shown
in Figure 20.3. Both types of test gave essentially the same results, with ch values decreasing slightly
with depth from 2  106 m2/s at a depth of 3 m to 1  106 m2/s at 14 m. The coefficient of consoli-
dation has also been calculated on the basis of Go obtained from cross-hole tests and ko measured both
in the laboratory and in situ, at depths between 5 and 9 m, where these two parameters do not vary sig-
nificantly. The calculated values were between 4 and 7  106 m2/s, depending mostly on the Poisson’s
ratio considered. As can be seen in Figure 20.3, these values are 3 to 5 times the values deduced from
SBP holding tests and piezocone dissipation tests.
Leroueil et al. (1995) evaluated ch with the piezocone at eight Champlain Sea clay sites where coef-
ficients of consolidation were also determined by other methods. At five of these sites, an in situ value
for the normally consolidated range was deduced from settlement observations of embankments, using
Asaoka (1978)’s method. On three of them, it was also possible to deduce an in situ value in the over-
consolidated range from pore pressure observations during the early stages of construction. Figure 20.4
summarizes the results. The in situ coefficients of consolidation in the normally consolidated range are
typically 10 times larger than the values deduced from oedometer test results, using the Casagrande
method. This is consistent with observations reported elsewhere (see Section 20.2.1). The in situ coef-
ficients of consolidation in the normally consolidated range are also about two orders of magnitude

227
09031-02[2].qxd 18/Oct/02 12:13 AM Page 228

Figure 20.4. Coefficients of consolidation (from Leroueil et al., 1995).

smaller than the values in the overconsolidated range deduced from in situ observations. The ch values
deduced from piezocone dissipation tests appear to be somewhere in between the field values obtained
for the overconsolidated and normally consolidated ranges. This could possibly be due to the fact that
pore pressure dissipation around a piezocone probe involves a volume of soil that is partly intact and
overconsolidated, and partly remoulded and normally consolidated.
The coefficient of consolidation thus remains one of the most difficult geotechnical parameters to
determine.

20.3 Hypothesis A vs Hypothesis B


One important practical question concerning consolidation is to know how to define the relevant com-
pression curve for in situ conditions. As indicated by Ladd et al. (1977) and Jamiolkowski et al. (1985),
there are two extreme possibilities (Figure 20.5). “Hypothesis A assumes that creep occurs only after the
end of primary consolidation … ” and consequently that the stress-strain curve followed in situ is the same
as the one obtained in the laboratory at the end of primary consolidation (instant compression line in
Figure 20.5a). At the end of primary consolidation, the strain in situ is thus the same as in the laboratory.
“Hypothesis B assumes that some sort of “structural viscosity” is responsible for creep, that this phe-
nomenon occurs during pore pressure dissipation, and therefore that the strain at the end of primary con-
solidation increases” with sample thickness. The compression line followed would thus be as the dashed
line in Figure 20.5a, i.e. below the compression line defined in the laboratory. The question of which one
of the two hypotheses is representative of the true behaviour has been the topic of many discussions
(Mesri & Choi, 1985; Kabbaj et al., 1988; Leroueil, 1988; Mesri et al., 1994, 1995; Leroueil, 1996).
There are many indications that the effective stress-strain curve followed by the soil during primary
consolidation is not unique but depends on strain rate thus on drainage path and loading conditions. As
a result of these indications, a number of pertinent issues can be raised:
– During secondary consolidation, there are settlements and thus hydraulic gradients and excess pore
pressures. How does the soil know the threshold gradient below which the soil is in secondary con-
solidation?
– Hypothesis A would mean that soil behaviour is controlled by excess pore pressures and not by effec-
tive stresses, which is contrary to the principles put forward by Terzaghi.
– Hypothesis A would mean that soil has one constitutive equation when there are pore pressures in
excess of a reference value and another when pore pressures are close to this reference value. This
would mean that an element in a soil deposit has to know this reference.

228
09031-02[2].qxd 18/Oct/02 12:13 AM Page 229

Figure 20.5. Consolidation of clay layers of different thickness according to hypotheses A and B.

– There are evidences in the laboratory that effective stress-strain curves followed by soil specimens or
sub-specimens depend on the strain rate history (Figures 9.3 to 9.7).
– Most of the well documented case histories show that, in accordance with the effects of strain rate and
temperature described in Sections 9 and 10, the compression curves followed by soil sub-layers under
embankments (which are at low strain rates) are below compression curves obtained in the laboratory
using high quality samples, indicating that Hypothesis A is not valid (Kabbaj et al., 1988; Leroueil,
1996, 2000; Hight & Leroueil, 2002*).
Thus in the opinion of the authors, Hypothesis A is wrong.

21 CONCLUSION

This paper is an overview of the mechanical and hydraulic behaviour and properties of natural soils and
soft rocks, from soft clay to chalk. It proposes a general framework into which most geomaterials can
be qualitatively fitted whatever their geological origin. This framework is based on the concepts of limit
and critical states (CSSM) that are extremely powerful for understanding soil and soft rock behaviour.
However, as CSSM does not take into account important features of natural geomaterials, such as
anisotropy, viscosity, microstructure, presence of discontinuities and partial saturation, it is necessary to
extend the basic concepts to incorporate the influence of these factors.
The writing of the paper has been guided by three main objectives: (a) summarize the current knowl-
edge on soil behaviour; (b) try to reduce the gap between evermore specialized researchers and the large

229
09031-02[2].qxd 18/Oct/02 12:13 AM Page 230

majority of practicing engineers who may not be aware of all the recent developments, or may not know
where to find information on specific aspects (e.g. small strain behaviour, localization, behaviour of
unsaturated soils, effects of principal stress rotation, etc.); (c) provide a document which, hopefully,
could be useful in the teaching of advanced soil mechanics.
It is obviously impossible to summarize the mechanical and hydraulic behaviour of soils and soft
rocks in a short conclusion. However, key aspects of this behaviour, or the recent and apparently inter-
esting developments appear to be:

1) Even though it has been demonstrated for almost 30 years, it is still not well recognized that Mohr-
Coulomb failure criterion has a tendency to underestimate the strength of soils for shear modes
other than compression.
2) Anisotropy plays a major role in soil behaviour as it may influence small strain soil behaviour, the
limit state curve and the consequences of stress axis rotation, in particular in undrained loading.
3) The concept of limit state, initially developed for reconstituted clays, captures well the yielding of
most natural soils. A simple model based on the idea that the shape of the limit state curve reflects
the distribution of contacts between soil particles or aggregates is advocated.
4) With the development of both accurate instruments for measuring strains in the pre-failure stage
and of geophysical methods, the understanding of the behaviour of soils and soft rocks at small
strains has progressed considerably in the last 20 years. In particular, it has been recognised that
dynamic and static moduli are the same when examined at the same very small strains.
5) The uniqueness of critical state or the steady state line in sands, as well as the usefulness of the state
parameter, are still the subject of debate in the profession. In particular, soil crushing is known to
influence the critical state line. The concept of collapse and static liquefaction of loose sands has
been fully accepted, although there are still discussions on the characteristics of the instability line.
6) It has been demonstrated that the preparation of sandy soil specimens influences strongly their
static behaviour during shearing. In particular, it has been shown that moist tamping can lead to
contractant behaviour while water pluviation can lead to dilatant behaviour at the same global void
ratio. A difference in hydraulic characteristics could also be seen, and all this has to be attributed to
differences in fabric.
7) The presence of clay or platey particles has a very significant effect on the behaviour of granular soils.
8) The importance of microstructure on the behaviour of soils is now fully recognized. This overview
confirms that microstructure is as important as void ratio and stress history for understanding of the
behaviour of natural soils. Microstructure influences yielding of soils and modifies the small strain
moduli that are then not dependent on applied stresses alone. Consequently, destructuration decreases
stiffness and strength of soils. It is also worth mentioning that, in addition to (strength) sensitivity,
a number of other approaches and parameters (stress and stiffness sensitivities) have been recently
proposed for characterizing microstructure.
9) Conditions for localization and its consequences have been examined recently; they give an inter-
esting insight into the development of post-failure deformations and its consequences on strength
mobilization.
10) The concept of brittleness, which is essential to the understanding of post-failure behaviour, has
been generalized to be applicable to a wide range of geomaterials and for conditions that can be
drained or undrained.
11) Viscosity, i.e. the influence of strain rate and temperature, seems to exist in most geomaterials. The
effects are more important in clayey soils but also exist in cohesionless soils. They are more impor-
tant at large strains but also exist at strains as small as 105. The effect of viscosity on the limit state
curve and the undrained shear strength appears to be about the same for all clays. On the other hand,
the critical state and the residual friction angles are almost not affected by changes in strain rate and
temperature.
12) Pore water chemistry may influence compressibility and strength characteristics of clayey soils in a
significant manner. This should be seen as an encouragement to use in situ pore water in laboratory
testing when appropriate.
13) Understanding of the behaviour of unsaturated soils has also progressed considerably in the last two
decades. It has been shown that the concept of limit state also applies to unsaturated soils. It has also
become clear that the soil-water characteristic curve controls to a large extent the influence of
matric suction on the entire limit state curve, including strength, and the hydraulic conductivity.
More research is however warranted, sometimes on some apparently simple aspects. For example,
the three levels of strength (peak, critical state and residual) that have been clearly established in
saturated soil mechanics do not seem to be so well defined in unsaturated soil mechanics.

230
09031-02[2].qxd 18/Oct/02 12:13 AM Page 231

There are also a number of remarks on soil properties that follow from this overview:
– The piezocone that was developed about two decades ago has become essential equipment for
detailed soil investigation in soft to stiff clays and sands. Many classifications and relationships with
a variety of soil parameters have been developed.
– Care should be taken over the use of the word “intrinsic” as, often, the considered characteristics can
be influenced by the mode of preparation, pore water chemistry, viscosity, etc.
– Plasticity index is generally not an appropriate parameter for characterizing the mechanical behaviour
of soils. Where possible, it seems more logical to use the critical state friction angle as a reference.
– Correlations in general certainly have to be used with care in geological environments where they
have not been developed or checked. Types of geomaterials that do not fit classical correlations seem
to be clayey soils containing diatoms and tropical, often unsaturated, soils.
One problem that can be found reading this paper is the lack of consistency in the symbols used. In
fact, the symbols used in geotechnical engineering vary with time and also with the authors, and, to
some extent, this problem is in itself an overview of the present situation in our profession. However,
this is not satisfactory and the authors thought that all this should be homogenized. Unfortunately, they
did not find the necessary time to do so.

ACKNOWLEDGEMENTS

The authors would first like to thank the Organising Committee of the International Workshop on
Characterisation and Engineering Properties of Natural Soils for giving them the unique opportunity to
prepare this Overview Paper. They in particular thank Professor Tan Thiam Soon (chair) and Dr. Phoon
Kok Kwang (Secretary) for their continuous encouragement. They also thank colleagues from different
parts of the world with whom, through the years and sometimes specifically for this Workshop, they
have exchanged ideas on various aspects of soil behaviour. They finally acknowledge the help of Cédric
Allenou, Mingjing Jiang, Jean Parent and Julie Turgeon in the preparation of the paper and of members
of the Organising Committee who kindly reviewed it.

LIST OF SYMBOLS

aw normalized area of water


A area; soil parameter
AL anisotropy line
AP air pluviation
b ( 2  3)/( 1  3)
Bq pore pressure ratio
c soil constant
cc.suc cohesion due to suction in compression
ce.suc cohesion due to suction in extension
cu undrained shear strength
cur remoulded undrained shear strength
c effective cohesion
ce Hvorslev cohesion
cpe Hvorslev cohesion parameter
cv coefficient of consolidation
C constant relating matric suction and radius of curvature of air-water interface
CAD anisotropically consolidated drained triaxial test
CAU anisotropically consolidated undrained triaxial test
CG controlled gradient (test)
CH cross hole (test)
CID isotropically consolidated drained triaxial compression test
CIU isotropicallly consolidated undrained triaxial compression test
CPT cone penetrometer
CPTU piezocone
CRS constant rate of strain oedemeter test
CS critical state

231
09031-02[2].qxd 18/Oct/02 12:13 AM Page 232

CSD constant shear drained (test)


CSL critical state line
CSSM critical state soil mechanics
CVDSS constant volume direct simple shear
Cc compression index
C*c intrinsic compression index
Cs recompression index
Css swelling strain index
Cs-HV, Cs-HH dimensional material constants
Ck hydraulic conductivity change index
Ckx hydraulic conductivity change index in the x direction
Cu uniformity coefficient; undrained shear strength
Ce secondary compression index
CF clay fraction
d diameter
dx particle diameter corresponding to x% passing
DH down hole (test)
DSC directional shear cell
DSS direct simple shear
Dr relative density
Ds equivalent hydraulic radius
D50 mean particle diameter
e void ratio
e*100 void ratio corresponding to v  100 kPa on the intrinsic compression line
e*1000 void ratio corresponding to v  1000 kPa on the intrinsic compression line
e rate of void ratio change
ecs void ratio on the critical state line
eg granular void ratio
emax maximum void ratio
emin minimum void ratio
eo initial void ratio
ess void ratio on the steady state line
(Ev)o, (Eh)o Young’s moduli under reference stress o
Evmo vertical Young’s modulus of microstructured soil under zero stress
Ehmo horizontal Young’s modulus of microstructured soil under zero stress
Esec secant Young’s modulus
Eu undrained secant modulus
Ex Young’s modulus in the direction x
fs friction measured
Fr normalized friction ratio
GFY name of Leroueil & Barbosa (2000)’s model for given fabric yielding
Geq shear modulus defined in small unload-reload cycle
Go small strain shear modulus
Gmax maximum shear modulus
Gs secant shear modulus
Gt tangent shear modulus
Gvh, Ghv, Ghh small strain shear moduli referring to vertical (v) and horizontal (h) axes
H height
HCA hollow cylinder apparatus
i hydraulic gradient
ICL intrinsic compression line
IB brittleness index
IGB generalized brittleness index
IL liquidity index
Ip plasticity index
IR relative dilatancy index
Iv void index
I1, I2, I3 first, second and third stress invariants
k hydraulic conductivity or coefficient of permeability

232
09031-02[2].qxd 18/Oct/02 12:13 AM Page 233

kx hydraulic conductivity in the x direction


ko initial hydraulic conductivity
kwr relative hydraulic conductivity
kwus hydraulic conductivity of the unsaturated soil
kws hydraulic conductivity of the saturated soil
K bulk modulus; stress ratio ( 3/ 1); intrinsic permeability
K*, KAL stress ratio representing the distribution of inter-particle or inter-aggregate contacts in soil
Ko coefficient of earth pressure at rest
Konc coefficient of earth pressure at rest of the normally consolidated soil
Ktan tangent bulk modulus
l length
LC loading collapse curve
m stress function parameter
m strain rate parameter
M slope of critical state line; deformation modulus
MT moist tamping
Mc slope of critical state line in compression
Me slope of critical state line in extension
Mce 6 sin e/(3  sin e)
Mee 6 sin e /(3  sin e)
n stress function parameter
N standard penetration test N-value (blows per 30 cm)
Nkt cone factor related to the undrained shear strength
N t cone factor related to the preconsolidation pressure
N60 N value for a SPT efficiency of 60%
(N1)60 N60 value corrected to a reference stress of one atmosphere
NCL normal compression line
NSP normalized soil parameter
NG soil parameter describing the change in Go with time due to aging
OCR overconsolidation ratio
PSC plane strain cell
p mean total stress
p mean effective stress
pcs mean effective stress on the critical state line at a given void ratio
pe mean effective stress on the normal compression line
pins mean effective stress on instability line
po initial mean effective stress
pp stress at the intersection of the swelling line with the isotropic normal consolidation line
pr reference pressure
pc crushing stress of soil particles
p*o isotropic preconsolidation pressure of the saturated soil
ps stress describing the brittleness of soils
pa or patm atmospheric pressure
q deviatoric stress ( 1  3); flow rate
qc measured piezocone tip resitance
qt corrected piezocone tip resistance
qcs deviatoric stress on the critical state line
Q soil parameter
Qt normalized tip resistance
r radius
rk anisotropy permeability ratio
R soil parameter
Ro overconsolidation ratio defined in isotropic conditions
RC resonant column (test)
s matric suction (us  uw)
su undrained shear strength
su TC undrained shear strength in compression
su TE undrained shear strength in extension
su DSS undrained shear strength measured in DSS test

233
09031-02[2].qxd 18/Oct/02 12:13 AM Page 234

SBP self-boring pressuremeter


SBS state boundary surface
SC seismic cone (test)
SPT standard penetration test
SSL steady state line
SWCC soil water characteristic curve
Sr degree of saturation
Sre effective degree of saturation
Srf submergence degree of saturation
Ss stiffness sensitivity
St Sensitivity or strength sensitivity
S stress sensitivity
t time
tp peak resistance at intermediate strain; time at the end of primary consolidation
TS torsional shear (test)
TTA true triaxial apparatus
T temperature
Tn inter-particle strength due to external stress
Ts surface tension of water
Tst inter-particle strength due to bonding
Tsuc inter-particle strength due to suction
Tstsuc additional inter-particle component of strength due to the combined effect of suction and
bonding
u pore pressure
u2 pore pressure measured just behind the tip of the piezocone probe
strain rate
u excess pore pressure
ua air pore pressure
ub pore water pressure at the base of the specimen
uw water pore pressure
(u*)
s res shear deformation of shear band necessary after peak to reach residual or steady state
conditions
v specific volume (1  e)
V volume
Vo initial volume
Vs shear wave velocity
Vs-HH shear wave velocity with shear wave propagating horizontally, with particle motion in the
horizontal direction
Vs-HV shear wave velocity with shear wave propagating horizontally, with particle motion in the
vertical direction
w water content
wL liquid limit
wp plastic limit
WT water pluviation
Yi yield curve
Ysat yield curve of the saturated material
 angle of contact between soil particles and air-water interface; angle of the axis of the
major principal stress to the vertical; exponent used for defining Ko
k soil parameter used for defining the hydraulic conductivity of unsaturated soils
p direction of the major principal stress to the vertical
 angle of the axis of the major principal stress to the vertical
p increment of plastic shear strain
pv increment of plastic volumetric strain
u excess pore pressure
V volume change
 y.st increase in gross yield stress due to bonding
 ysuc increase in gross yield stress due to suction
 y.stsuc increase in gross yield stress due to the combined effect of bonding and suction
1 axial strain

234
09031-02[2].qxd 18/Oct/02 12:13 AM Page 235

x axial strain in the x direction


e, e elastic strain
h horizontal strain
p, p plastic strain
s shear strain
v vertical strain
vol volumetric strain
x axial strain in the direction x
strain rate
1 axial strain rate
24 h axial strain rate after 24 hours in conventional oedometer test
x strain rate in the x direction
 friction angle
cs critical state friction angle
crit critical state friction angle
e Hvorslev angle of shearing resistance
mob mobilized angle of friction
nc friction angle of the normally consolidated soil
r residual friction angle
tc friction angle in triaxial compression
te friction angle in triaxial extension
b unsaturated soil strength parameter
 unit weight; shear strain
p unit weight of permeant liquid
vh, hv, hh shear strains referring to vertical (v) and horizontal (h) axes
w unit weight of water
1 linear threshold shear strain

slope of recompression curve in e–ln p diagram; soil parameter
slope of the normal compression line in v-ln p or e–ln p diagram; slope of the steady
state or critical state line in e–log p diagram
c slope of the steady state or critical state line in e–log p diagram, at stresses larger than
the crushing stress
s slope of the normal compression line in e–ln (p–ua) diagram for unsaturated soils
 viscosity of permeant liquid
 Poisson’s ratio
 volumetric water content; angle of failure surface developed in laboratory test to the
horizontal
s volumetric water content at saturation
r volumetric water content at residual conditions
 mass density
 effective normal stress
1 major principal effective stress; axial effective stress
2 intermediate principal effective stress
3 minor principal effective stress; radial effective stress
a axial effective stress
ac axial consolidation stress
c consolidation effective stress
o reference effective stress
p preconsolidation pressure or vertical gross yield stress in one-dimensional
compression test
p iso isotropic gross yield stress
n normal effective stress
r radial effective stress
*ve vertical effective stress on the intrinsic compression line
vm maximum vertical effective stress
vy vertical yield stress or preconsolidation pressure
x effective stress in the direction x
y gross yield stress
y.e gross yield stress associated with void ratio

235
09031-02[2].qxd 18/Oct/02 12:13 AM Page 236

1 major principal total stress; axial total stress


2 intermediate principal total stress
3 minor principal total stress; radial stress
 shear stress
f shear strength
mob mobilized shear stress
p peak strength
r residual strength
vh, hv, hh shear stresses referring to vertical (v) and horizontal (h) axes
 state parameter (eo – ess)
N normalized state parameter
苵 modified state parameter
 effective stress parameter
 normalized volumetric water content
(ua – uw) matric suction or suction
(ua – uw)b air entry value or bubbling pressure
( – ua) net total stress

REFERENCES

Aas, G., Lacasse, S., Lunne, T. & Hoeg, K. 1986. Use of in situ test for foundation design on clay. Proc. ASCE
Specialty Conf. In Situ’86: Use of in situ Tests in Geotechnical Engineering, Blacksburg, pp. 1–30.
Adachi, T., Oka, F. & Zhang, F. 1998. An elasto-viscoplastic constitutive model with strain softening. Soils &
Foundations, 38(2): 27–35.
Addis, M.A. 1987. Mechanisms of sediment compaction responsible for field subsidence. Ph.D. Thesis, University
of London, U.K.
Addis, M.A. & Jones, M.E. 1990. Mechanical behaviour and strain rate dependence of high porosity chalk. Proc. of
the 1st Chalk Symp., Brighton: 111–116.
Afifi, S.E.A. & Richart, F.E.Jr. 1973. Stress-history effects on shear modulus of soils. Soils & Foundations, 13(1):
77–95.
Akai, K., Adachi, T. & Ando, N. 1975. Existence of a unique stress-strain-time relation of clays. Soils & Foundations,
15(1), 1–16.
Alarcon-Guzman, A., Leonards, G.A. & Chameau, J.L. 1988. Undrained monotonic and cyclic strength of sands.
J. Geotech. Engng, ASCE, 114(10): 1089–1109.
Almeida, M.S.S. & Marques, M.E.S. 2002. The behaviour of Sarapuí soft clay. Proc. Int. Workshop on
Characterisation and Engineering Properties of Natural Soils, Singapore.
Alonso, E.E. 1998. Suction and moisture regimes in roadway bases and subgrades. Proc. Int. Symp. on Subdrainage
in Roadway Pavements and Subgrades, Grenade, 57–104.
Alonso, E.E., Gens, A. & Hight, D.W. 1987. Special problem soils. General report. Proc. 9th European Int. Conf. on
Soil Mech. and Found. Engrg., Dublin, 3: 1087–1146.
Alonso, E.E., Gens, A. & Josa, A. 1990. A constitutive model for partially saturated soils. Géotechnique, 40(3):
405–430.
Alonso, E.E., Gens, A. & Gehling, W.Y.Y. 1994. Elasto-plastic model for unsaturated expansive soils. Proc. 3rd Eur.
Conf. Num. Methods Geotech. Eng., Manchester: 11–18.
Alonso, E.E., Vaunat, J. & Gens, A. 2001. Modelling the mechanical behaviour of expansive clays. Engineering
Geology, 54(1–2):173–183.
Al-Tabbaa, A. & Wood, D.M. 1987. Some measurements of the permeability of kaolin. Géotechnique, 37(4):
499–503.
Amarosi, A. & Rampello, S. 1998. The influence of natural soil structure on the mechanical behaviour of a stiff clay.
Proc. 2nd Int. Symp. on the Geotechnics of Hard Soils – Soft Rocks, Naples, Vol. 1: 395–402.
Anderson, D.G. & Richart, F.E. 1974. Temperature effect on shear wave velocity in clays. J. Geotech. Engng. Div.,
ASCE, 100: 1316–1320.
Anderson, D.G. & Stokoe, K.H. 1978. Shear modulus, a time dependent soil property. In Dynamics Geotechnical
Testing, ASTM, STP 654: 66–90.
Anderson, S.A. & Riemer, M.F. 1995. Collapse of saturated soil due to reduction in confinement. J. of Geotechnical
Engineering, ASCE, 121(2): 216–220.
Anderson, S.A. & Sitar, N. 1995. Analysis of rainfall-induced debris flows. J. of Geotechnical Engineering, ASCE,
121(7): 544–552.
Arthur, J. R. F., Bekenstein, S., Germaine, J. T. & Ladd, C. C. 1981. Stress path tests with controlled rotation of prin-
cipal stress directions. ASTM STP740, “Laboratory shear strength of soils”, 516–540.
Arthur, J. R. F., Chua, K. S. & Dunstan, T. 1977. Induced anisotropy in a sand. Géotechnique, 27(1): 13–30.

236
09031-02[2].qxd 18/Oct/02 12:13 AM Page 237

Arulanandan, K., Shen, C.K., & Young, R.B. 1971. Undrained creep behaviour of a coastal organic silty clay.
Géotechnique, 21(4): 359–375.
Asaoka, A. 1978. Observational procedure of settlement prediction. Soils & Foundations, 18(4): 87–101.
Atkinson, J.H., Richardson, D. & Stallebrass, S.E. 1990. Effects of stress history on the stiffness of overconsolidated
soil. Géotechnique, 40(4): 531–540.
Atterberg, A. 1911. Über die physikalishe Bodenuntersuchung und über die Plastizität der Tone. Int. Mitt. Für
Bodenkunde, 1, pp. 10–43.
Aubertin, M. & Simon, R.A. 1996. A multiaxial failure criterion that combines two quadratic surfaces. Proc. Conf.
on Rock Mechanics: Tools and Techniques, pp. 1729–1736.
Aubertin, M. & Simon, R.A. 2000. A multiaxial stress criterion for short- and long-term strength of isotropic rock
media. Int. J. of Rock Mech. and Mining Sciences, 37: 1169–1193.
Aubertin, M., Julien, M.R., Servant, S. & Gill, D.E. 1999. A rate-dependent model for the ductile behavior of salt
rocks. Canadian Geotech J., 36(): 660–674.
Aversa, S., Evangelista, A. & Ramondini, M. 1991. Snervamento e resistenza a rottura di un fufo a grana fine. Atti
del II Convegno dei Ricercatori del G.N.C.S.I.G. del C.N.R., Ravello, 1: 3–22.
Aversa, S., Evangelista, A., Leroueil, S. & Picarelli, L. 1993. Some aspects of the mechanical behaviour of “struc-
tured” soils and soft rocks. Proc. Int. Symp. on Hard Soils – Soft Rocks, Athens, Vol. 1: 359–366.
Badra-Blanchet, P. 1981. Effet de la vitesse de déformation sur le comportement de l’argile normalement consolidée.
M.Sc. Thesis, Université Laval, Québec.
Baldi, G., Bellotti, R., Ghionna, V.N., Jamiolkowski, M. & Lo Presti, D.C.F. 1989. Modulus of sands from CPTs and
DMTs. Proc. 12th Int. Conf. on Soil Mech. and Found. Engng., Rio de Janeiro, Vol. 1: 165–170.
Baligh, M.M. 1985. Strain path method. J. Geotech. Engng Div., ASCE, 111, GT9, 1108–1136.
Banks, D.C., Strohm, W.E., De Angulo, M. & Lutton, R.J. 1975. Study of clay-shale slopes along the Panama Canal.
Report No.3, Engineering Analyses of slides and strength properties of clay shales along the Gaillard Cut.
Technical report S-70-9. US Army Engineers Waterways Experiment station, Vicksburg, Miss.
Barbour, S.L. & Fredlund, D.G. 1989. Mechanisms of osmotic flow and volume change in clay soils. Canadian
Geotech. J. 26: 551–562.
Baguelin, F. & Jézéquel, J.F. 1973. Le pressiomètre autoforeur. Annales de l’Institut Technique du Bâtiment et des
Travaux Publics. Supplément n°307-308, série Sols et Fondations n°97, pp. 133–160.
Been, K. & Jefferies, M.G. 1985. A state parameter for sands. Géotechnique, 35(2): 99–112.
Been, K., Jefferies, M.G. & Hachey, J. 1991. The critical state of sands. Géotechnique, 41(3): 365–381.
Bellotti,. R., Jamiolkowski, M., Lo Presti, D.C.F. & O’Neill, D.A. 1996. Anisotropy of small strain stiffness in Ticino
sand. Géotechnique, 46(1): 115–131.
Berre, T. & Iversen, K. 1972. Oedometer tests with different specimen heights on a clay exhibiting large secondary
compression. Géotechnique, 22(1): 53–70.
Berry, P.L. & Poskitt, T.J. 1972. The consolidation of peat. Géotechnique, 22(1): 27–52.
Bishop, A.W. 1959. The principle of effective stress. Tek. Ukeblad, Vol. 39: 859–863.
Bishop, A.W. 1967. Progressive failure – with special reference to the mechanism causing it. Proc. Geotechnical
Conf., Oslo, Vol.2:142–150.
Bishop, A.W. & Bjerrum, L. 1960. The relevance of the triaxial test to the solution of stability problems. Proc. ASCE
Research Conference on Shear Strength of Cohesive Soils, Boulder, pp. 437–501.
Bishop, A.W., Webb, D.L. & Lewin, P.I. 1965. Undisturbed samples of London clay from the Ashford Common
shaft; strength-effective stress relationships. Géotechnique, 15, No. 1, 1–13.
Bishop, A.W. & Green, G.E. 1965. The influence of end restraint on the compression strength of a cohesionless soil.
Géotechnique, 15(3): 243–266.
Bishop, A.W. & Lovenbury, H.T. 1969. Creep characteristics of two undisturbed clays. Proc. 7th Int. Conf. on Soil
Mech. and Found. Engng, Mexico City, 1: 29–37.
Bjerrum, L. 1954. Geotechnical properties of Norwegian Marine clays. Géotechnique, 4(2): 49–69.
Bjerrum, L. 1967. Engineering geology of normally consolidated marine clays as related to the settlement of build-
ings. Géotechnique, 17(2): 83–119.
Bjerrum, L. 1973. Problems of soil mechanics and construction on soft clays. Proc. 8th Int. Conf. on Soil Mech. and
Found. Engrg., Moscow, Vol. 3, pp. 111–159.
Bjerrum, L., Simons, N. & Torblaa, I. 1958. The effect of time on the shear strength of a soft marine clay. Proc. Conf.
on Earth Pressure Problems, Brussels, 1(1): 48–158.
Bjerrum L. & Landva A. 1966. Direct simple shear testing on a Norwegian quick clay. Géotechnique, 16(1):
1–20.
Bjerrum, L. & Simons, N.E. 1960. Comparison of shear strength characteristics of normally consolidated clays.
Proc. ASCE Research Conf. on Shear Strength of Cohesive Soils, Boulder, pp. 711–726.
Bjerrum, L. & Andersen, K.H. 1972. In situ measurement of lateral pressures in clay. Proc. 5th European Conf. On
Soil Mech. and Found. Eng., Madrid, Vol. 1, pp. 11–20.
Bolton, M.D. 1986. The strength and dilatancy of sands. Géotechnique, 36(1): 65–78.
Bouclin, G. 1990. Perméabilité de l’argile de St-Esprit et anisotropie de perméabilité. M.Sc. Thesis, Université
Laval, Québec, Canada.
Boudali, M. 1995. Comportement tridimensionnel et visqueux des argiles naturelles. Ph.D. Thesis, Université Laval,
Québec.

237
09031-02[2].qxd 18/Oct/02 12:13 AM Page 238

Boudali, M., Leroueil, S. & Murthy, B.R.S. 1994. Viscous behaviour of natural soft clays. Proc. 13th Int. Conf. on
Soil Mech. and Found. Engrg., New Delhi, 1: 411–416.
Bressani, L.A. & Vaughan, P.R. 1989. Damage to soil structure duringtriaxial testing. Proc. 12th Int. Conf. on Soil
Mech. and Found. Engrg., Rio de Janeiro, Vol. 1:17–20.
Brignoli, E. 1997. Qualifica dei riporti di rilevati artificiali mediante tecniche di misura della velocità di
propagazione delle onde di taglio. RAT-STA-1657/97, ISMES, Bergamo (ref. by Jamiolkowski and Lo Presti,
1998).
Brooks, H. & Corey, A.T. 1964. Hydraulic properties of porous media. Colorado State University, Hydrology
Paper 3.
Buisson, M.S.R. & Wheeler, S.J. 2000. Inclusion of hydraulic hysteresis in a new elasto-plastic framework for
unsaturated soils. Proc. Int. Workshop on Experimental Evidence and Theoretical Approaches in Unsaturated
Soils, Trento, pp.: 109–119.
Burghignoli, A. & Desideri, A. 1988. Influenza della temperatura sulla compressibilità delle argille. Gruppo
Nazionale di Coodinamento per gli Studi di Ingegneria Geotecnica, Convegno di Monselice: 193–206 (Ref. by
Burghignoli et al. 1992).
Burghignoli, A., Desideri, A. & Miliziano, S. 1992. Deformability of clays under non isothermal conditions. Rivista
Italiana di Geotecnica, Anno XXVI, 4: 227–236.
Burland, J.B. 1965. Some aspects of the mechanical behaviour of partially saturated soils. Moisture Equilibria and
Moisture Changes Beneath Covered Areas, Butterworths, Sydney, pp. 270–278.
Burland, J.B. 1989. Ninth Laurits Bjerrum Memorial Lecture: “Small is beautiful” – the stiffness of soils at small
strains. Canadian Geotech. J., 26(4): 499–516.
Burland, J.B. 1990. On the compressibility and shear strength of natural clays. Géotechnique 40(3): 329–378.
Burland, J.B., Rampello, S., Georgiannou, V.N. & Calabresi, G. 1996. A laboratory study of the strength of four stiff
clays. Géotechnique, 46(3): 491–514.
Burns, S.E. & Mayne, P.W. 1998. Monotonic and dilatory pore-pressure decay during piezocone tests in clay.
Canadian Geotech. J., 35(6): 1063–1073.
Bucher, F. 1975. Ref by Lupini et al., 1981.
Butcher, A.P. & Powell, J.J.M. 1995. Practical considerations for field geophysical techniques used to assess ground
stiffness. Proc. Int. Conf. on Advances in Site Investigation Practice, London, pp. 701–714,
Butterfield, R. 1979. A natural compression law for soils (An advance on e – log p). Géotechnique, 29(4): 469–480.
Calabresi, G. 1980. The effect of sample size on strength parameters for intact and fissured stiff clays. Proc. of
Euromechanical Colloquium no. 134, Copenhagen: Technical University of Denmark.
Calabresi, G. & Manfredini, G. 1973. Shear strength characteristics of the jointed clay of S. Barbara. Géotechnique,
23(2): 233–244.
Calabresi, G. & Scarpelli, G. 1985. Effects of swelling caused by unloading in overconsolidated clays. Proc. 11th Int.
Conf. on Soil Mech. and Found. Engrg., San Francisco, Vol. 2: 411–414.
Callisto, L. & Calabresi, G. 1998. Mechanical behaviour of a natural soft clay. Géotechnique, 48(4): 495–513.
Campanella, R.G., Robertson, P.K. & Gillepsie, D. 1986. Seismic cone penetration test. Proc. ASCE Spec. Conf.
In Situ’86 on Use of In Situ Tests in Geotech. Engng., Blacksburg, pp. 116–130.
Carvalho, L.C., Loret, A. & Josa, A. 2002. Isotropic compression tests on an artificially cemented unsaturated soil.
Proc. 3rd Int. Conf. on Unsaturated Soils, Recife, Vol. 2: 529–534.
Casagrande, A. 1932. The structure of clay and its importance in foundation engineering. Journal of the Boston
Society of Civil Engineers: 168–209.
Casagrande, A. 1936a. Characteristics of cohesion-less soils affecting the stability of earth fills. J. Boston Society of
Civil Engineers. Reprinted in “Contributions to Soil Mechanics 1925–1940”, Boston Society of Civil Engineers.
Casagrande, A. 1936b. The determination of the preconsolidation load and its practical significance. Proc. 1st. Int.
Conf. on Soil Mech. and Found. Engrg., 3: 60–64.
Casagrande, A. 1975. Liquefaction and cyclic deformation of sands, a critical review. Proc. 5th Panamerican Conf.
on Soil Mech. and Found. Engng., Buenos Aires, Vol. 5: 79–133.
Casagrande, A. & Watson, J.D. 1938. Compression tests and critical density investigations of cohesionless materials
for Franklin Falls Dam, Merrimack Valley flood control. Corps of Engineers, U.S. Army Engineering Office,
BII-7 (Ref. by Desrues et al., 1996).
Castro, G., Enos, J.L., France, J.W. & Poulos, S.J. 1982. Liquefaction induced by cyclic loading. Report to National
Science Foundation, Washington, D.C., No NSF/CEE-82018.
Chan, H.T. & Kenney, T.C. 1973. Laboratory investigation of permeability ratio of New Liskeard varved soil.
Canadian Geotechnical J., 10(3): 453–471.
Chandler, R.J. 1984. Recent European experience of landslides in over-consolidated clays and soft rocks. Proc. 4th
International Symposium on Landslides, Toronto, Vol. 1: 61–81.
Chandler, R., Leroueil, S. & Trenter, N. 1990. Measurement of the permeability of London clay using a self-boring
permeameter. Géotechnique, 40(1): 113–124.
Chang, M.F., Teh, C.I. & Cao, L.F. 2001. Undrained cavity expansion in modified Cam Clay II: Application to the
interpretation of the piezocone test. Géotechnique, 51(4): 335–350.
Chu, J. 2002. Personal communication.
Chu, J. & Lo, S.C.R. 1994. Asymptotic behaviour of a granular soil in strain path testing. Géotechnique, 44(1):
65–82.

238
09031-02[2].qxd 18/Oct/02 12:13 AM Page 239

Chu, J., Leroueil, S. & Leong, W.K. 2001. A framework for interpretation of instability of slope. Canadian
Geotechnical Conf., Calgary.
Chu, J., Leroueil, S. & Leong, W.K. 2002. Unstable behaviour of sand and its implications for slope instability.
Canadian Geotechnical J. Submitted for publication.
Chung, S.G., Giao, P.H. & Tanaka, H. 2002. Geotechnical characteristics and engineering problems of Pusan Clays.
Proc. Int. Workshop on Characterisation and Engineering Properties of Natural Soils, Singapore.
Clarke, B.G., Carter, J.P. & Wroth, C.P. 1979. In situ determination of the consolidation characteristics of saturated
clays. Proc. 7th European Conf. on Soil Mech. and Found. Engng., Brighton, Vol. 2: 207–213.
Clayton, C.R.I. & Serratrice, J.F. 1993. General Report Session 2: The mechanical properties and behavior of hard
soils and soft rocks. Proc. 1st Int. Symp. on the Geotechnics of Hard Soils – Soft Rocks, Athens, Vol. 3:
1839–1877.
Clayton, C.R.I., Gordon, M.A. & Matthews, M.C. 1994. Measurement of stiffness of soils and weak rocks using
small strain laboratory testing and field geophysics. Proc. 1st Int. Conf. on Pre-Failure Deformation Characteristics
of Geomaterials, Sapporo, Vol. 1: 229–234.
Clayton, C.R.I. & Heymann, C. 2001. The stiffness of geomaterials at very small strains. Géotechnique, 51(3):
245–256.
Clayton, C.R.I., Matthews, M.C. & Heymann, G. 2002. The Chalk. Proc. Int. Workshop on Characterisation and
Engineering Properties of Natural Soils, Singapore.
Colliat-Dangus, J.-L., Desrues, J. & Foray, P. 1988. Triaxial testing of granular soil under elevated cell pressure. Proc.
ASTM Conf. on Advanced Triaxial Testing of Soil and Rocks, STP 977: 290–310.
Cook, N.G.W. 1992. Jaeger Memorial Dedication Lecture – Natural joints in rock: Mechanical, hydraulic and seis-
mic behaviour and properties under normal stress. Int. J. Rock Mech. Min. Sci. & Geomech., 29(3): 198–223.
Coop, M.R. & Airey, D.W. 2002. Carbonate sands. Proc. Int. Workshop on Characterisation and Engineering
Properties of Natural Soils, Singapore.
Cotecchia, F. & Chandler, R.J. 1997. The influence of structure on the pre-failure behaviour of a natural clay.
Géotechnique, 47(3): 623–544.
Cotecchia, F. & Chandler, R.J. 2000. A general framework for the mechanical behaviour of clays. Géotechnique,
50(4): 431–447.
Cotecchia, F. 2002. Mechanical behaviour of the stiff clays from the Montemesola Basin in relation to their geolog-
ical history and structure. Proc. Int. Workshop on Characterisation and Engineering Properties of Natural Soils,
Singapore.
Coutinho, R.Q. & Lacerda, W.A. 1987. Characterization-consolidation of Juturnaiba organic clays. Proc. Int. Symp.
on geotechnical Engineering of Soft Soils, Mexico City, Vol. 1: 17–24.
Crawford, C.B. 1964. Interpretation of the consolidation test. J. Soil Mech. and Found. Engrg. Div., ASCE, 90(5): 87–102.
Croney, D. & Coleman, J., 1961. Pore pressure and suction in soils. Proc. Conf. on Pore Pressure and Suction in
Soils, London, pp. 31–37.
Cuccovillo, T. & Coop, M.R. 1993. The influence of bond strength on the mechanics of carbonate soft rocks. Proc.
1st Int. Symp. on Geotechnical Engng. of Hard Soils-Soft Rocks, Athens, Vol. 1: 447–455.
Cuccovillo, T. & Coop, M.R. 1997. Yielding and pre-failure deformation of structured sands. Géotechnique, 47(3):
491–508.
Cui, Y.J. & Delage, P. 1996. Yielding and plastic behaviour of an unsaturated compacted silt. Géotechnique, 46(2):
291–311.
Danziger, F.A.B., Politano, C.F. & Danziger, B.R. 1998. CPT-SPT correlations for some Brazilian residual soils.
Proc. 1st Int. Conf. on Site Characterization, Atlanta, Vol. 2: 907–912.
Dasari, G.R. & Bolton, M.D. 1998. Comparison of field and laboratory stiffness of Gault clay. Symp. on Pre-Failure
Deformation Behaviour of Geomaterials, London, pp. 345–352.
De Campos, T.M.P. & Carrillo, C.W. 1995. Direct shear testing on an unsaturated soil from Rio de Janeiro. Proc.1st
Int. Conf. on Unsaturated Soils, Paris, 1: 31–38.
DeGroot, D.J. & Lutenegger, A.J. 2002. Geology and engineering properties of Connecticut Valley Varved Clay.
Proc. Int. Workshop on Characterisation and Engineering Properties of Natural Soils, Singapore.
Delage, P., Suraj de Silva, G.P.R. & De Laure, E. 1987. Un nouvel appareil triaxial pour les sols non saturés. Proc.9th
European Conf. on Soil Mech. and Found. Engrg., Dublin, 1: 26–28.
D’Elia, B. 1991. Deformation problems in the Italian structurally complex clay soils. Proc. 10th European Conf. on
Soil Mech. and Found. Engng., Florence, Vol. 4: 1159–1170.
D’Elia, B. 1994. Ricerca sperimentale sul comportamento reologico delle argille sovraconsolidate del Valdarno
Superiore (Ropporto Finale). Internal Report, Universita di Roma La Sapienza.
D’Elia, B., Picarelli, L., Leroueil, S. & Vaunat, J. 1998. Geotechnical characterization of slope movements in struc-
turally complex clay soils and stiff jointed clays. Italian Geotechnical Journal, Anno XXXII, No: 3: 5–32.
Demers, D. & Leroueil, S. 2002. Evaluation of preconsolidation pressure and the overconsolidation ratio from piezo-
cone tests of clay deposits in Quebec. Canadian Geotechnical J., 39(1): 174–192.
Den Haan, E.J. 1992. The formulation of virgin compression of soils. Géotechnique, 42(3): 465–483.
Desrues, J., Chambon, R., Mokni, M. & Mazerolle, F. 1996. Void ratio evolution inside shear bands in triaxial sand
specimens studied by computed tomography. Géotechnique, 46(3): 529–546.
De Waal, J.A. 1986. On the rate type compactation behavior of sandstone reservoir rock. Ph.D. Thesis, Delft
Technical University, Delft, The Netherlands.

239
09031-02[2].qxd 18/Oct/02 12:13 AM Page 240

Díaz-Rodríguez, J.A. 2002.Characterization and engineering properties of Mexico City lacustrine soils. Proc. Int.
Workshop on Characterisation and Engineering Properties of Natural Soils, Singapore.
Diaz-Rodriguez, J.A., Leroueil, S. & Aleman, J.D. 1992. Yielding of Mexico City clay and other natural clays.
J. Geotechnical Div., ASCE, 118(7): 981–995.
Diène, M. 1989. Mesure in situ de la perméabilité des argiles. Ph.D. Thesis, Université Laval, Québec, Canada.
Di Maio, C. 1996a. The influence of pore fluid composition on the residual shear strength of some natural clayey
soils. Proc. 7th Int. Symp. on Landslides, Trondheim. Vol. 2: 1189–1194.
Di Maio, C. 1996b. Exposure of bentonite to salt solution: osmotic and mechanical effects. Géotechnique,
46(4):695–707.
Di Maio, C. 2002. Personal communication.
Di Maio, C. & Onorati, R. 2000. Influence of pore liquid composition on the shear strength of an active clay. Proc.
8th Int. Symp. on Landslides, Cardiff, Vol. 1: 463–468.
Doanh, T., Ibraim, E., Dubujet, P., Mariotti, R. & Herle, I. 1999. Static liquefaction of very loose Hostun RF sand:
Experiments and modelling. Int. Workshop on the Physics and Mechanics of Soil Liquefaction, Baltimore,
pp. 17–28.
Donald, I.B. 1956. Shear strength measurements in unsaturated non-cohesive soils with negative pore pressures.
Proc. 2nd Australia-New Zealand Conf. Soil Mech. and Found. Engng, Christchurch, pp. 200–205.
D’Onofrio, A. 1996. Comportamento meccanico dell’arglla di Vallericca in condizioni lontane dalla rottura. Ph.D
Thesis, Universita di Napoli ‘Federico II’, Italy.
D’Onofrio, A., Santucci de Magistris, F. & Olivares, L. 1998. Influence of soil structure on the behaviour of two
natural stiff clays in the pre-rupture range. Proc. 2nd Int. Symp. on Hard Soils-Soft Rocks, Naples, Vol. 1:
497–505.
Duncan, J.M. & Chang, C. 1970. Nonlinear analysis of stress and strain in soils. J. Soil Mech. and Found, Engng,
ASCE, 96(5): 1629–1653.
Dyvik, R. & Madshus, C. 1985. Lab measurement of Gmax using bender elements. Proc. ASCE Annual Convention.
Advances in the Art of Testing Soils Under Cyclic Conditions, Detroit.
Edil, T.B. & den Haan, E.J. 1994. Settlement of peats and organic soils. Specialty Conf. on Vertical and Horizontal
Deformations of Foundations and Embankments, ASCE, Settlement ’94, College Station, 2: 1543–1572.
Feijo, R.L., & Martins I.S.M. 1993. Relacao entre compressao secundaria, OCR e K0. COPPEGO ’93. Simposio
Geotecnico Comemorativo dos 30 anos da COPPE-UFRJ, UFRJ, Rio de Janeiro, 27–40.
Fernandez, A.L. & Santamrina, J.C. 2001. Effect of cementation on the small-strain parameters of sands. Canadian
Geotechnical J., 38(1): 191–199.
Elliot, G.M. & Brown, E.T. 1985. Yield of a soft high porositiy rock. Géotechnique, 35(4): 413–423.
Eriksson, L.G. 1989. Temperature effects on consolidation properties of sulphide clays. Proc. 12th Int. Conf. on Soil
Mech. and Found. Engrg., Rio de Janeiro, 3: 2087–2090.
Escario, V. & Saez, J. 1986. The shear strength of partly saturated soils. Géotechnique, 36(3): 453–456.
Escario, V. & Juca, J.F.T. 1989. Strength and deformation of partly saturated soils. Proc. 12th Int. Conf. on Soil Mech.
and Found. Engrg., Rio de Janeiro, 1: 43–46.
Fahey, M. 1998. Deformation and in situ stress measurement. Proc. 1st Int. Conf. on Site Characterization, Atlanta,
Vol. 1: 49–68.
Félix, B., Magnan, J.P., Josseaume, H., Kenana, A., Piyal, M. & Shahanguian, S. 1985. Comportement triaxial de
l’argile molle de Cubzac. Proc. 11th Int. Conf. on Soil Mech. and Found. Engrg., San Francisco, 2: 451–454.
Fernandez, F. & Quigley, R.M. 1985. Hydraulic conductivity of natural clays permeated with simple liquid hydro-
carbons. Canadian Geotechnical J., 22(2): 205–214.
Fioravante, V., Jamiolkowski, M., Lo Presti, D.C.F., Manfredini, G. & Pedroni, S. 1998. Assessment of the coefficient
of the earth pressure at rest from shear wave velocity measurements. Géotechnique, 48(5): 657–666.
Fleureau, J.-M., Kheirbek-Saoud, S. Taibi, S. 1995. Experimental aspects and modelling the behaviour of soils with
a negative pressure. Proc.1st Int. Conf. on Unsaturated Soils, Paris, Vol. 1: 57–62.
Fredlund, D.G., Morgenstern, N.R. & Widger, R.A. 1978. The shear strength of unsaturated soils. Canadian
Geotechnical J., 15(3): 313–321.
Fredlund, D.G. & Rahardjo, H. 1993. Soil Mechanics for Unsaturated Soils. John Wiley and Sons, Inc., pp. 517
Fredlund, D.G. & Xing, A. 1994. Equations for the soil-water characteristic curve. Canadian Geotechnical J.,
31: 521–532.
Fredlund, D.G., Vanapalli, S., Xing, A. & Pufahl, D.E. 1995. Predicting the shear strength function for unsaturated
soils using the soil-water characteristic curve. Proc.1st Int. Conf. on Unsaturated Soils, Paris, Vol. 1: 63–69.
Fredlund, D.G., Xing, A., Fredlund, M.D. & Barbour, S.L. 1996. The relationship of the unsaturated soil shear
strength to the soil-water characteristic curve. Canadian Geotechnical J., 33(3): 440–448.
Fredlund, D.G., Fredlund, M.D. & Zakerzadeh, N. 2001. Predicting the permeability function for unsaturated soils.
Proc. Int. Symp. on Suction, Swelling, permeability and Structured Clays, IS-Shizuoka.
Futai, M.M., Almeida, M.S.S. & Lacerda, W.A. 2001. Geotechnical properties of Rio de Janeiro soft clays. Brazilian
Meeting on Soft Clay Properties, Rio de Janeiro.
Gan, J.K.M. & Fredlund, D.G. 1988. Multistage direct shear testing of unsaturated soils. Geotechnical Testing J.,
11(2): 313–321.
Geiser, F., Laloui, L. & Vulliet, L. 1999. Unsaturated soil modelling with special emphasis on undrained conditions.
Numerical models in Geomechanics, NUMOG VII, Graz, pp. 9–14.

240
09031-02[2].qxd 18/Oct/02 12:13 AM Page 241

Geiser, F., Laloui, L. & Vulliet, L. 2000. Modelling the behaviour of unsaturated silt. Proc. Int. Workshop on
Experimental Evidence and Theoretical Approaches in Unsaturated Soils, Trento, pp.: 155–175.
Gens, A. & Alonso, E.E. 1992. A framework for the behaviour of unsaturated expansive clays. Canadian
Geotechnical J. 29: 1013–1032.
Gens, A. & Nova, R. 1993. Conceptual bases for a constitutive model for bonded soils and weak rocks. Proc. 1st Int.
Symp. on the Geotechnics of Hard Soils – Soft Rocks, Athens, Vol. 1: 485–494.
Georgiannou, V.N., Burland J.B. & Hight D.W. 1990. The undrained behaviour of clayey sands in triaxial compres-
sion and extension. Géotechnique, 40(3): 431–449.
Georgiannou, V.N., Hight D.W. & Burland, J.B. 1991a. Undrained behaviour of natural and model clayey sands. Soils
& Foundations, 31(3): 17–29.
Georgiannou, V.N., Hight D.W. & Burland, J.B. 1991b. Behaviour of clayey sands under undrained cyclic triaxial
loading. Géotechnique, 41(3): 383–393.
Gilboy, G. 1928. The compressibility of sand-mica mixtures. Proc. ASCE Transactions, Vol. 54, 555–568.
Goto, S., Suzuki, Y., Nishio, S. & Ohoka, H. 1992. Mechanical properties of undisturbed Tone-River gravel obtained
by in-situ freezing method. Soils & Foundations, 32(3): 15–25.
Graham, J., Crooks, J.H.A. & Bell, A.L. 1983a. Time effects on the stress-strain behaviour of soft natural clays.
Géotechnique, 33(3): 327–340.
Graham, J., Noonan, M.L. & Lew, K.V. 1983b. Yield states and stress-strain relationships in a natural plastic clay.
Canadian Geotechnical J., 20(3): 502–516.
Graham, J., Tanaka, N., Grilly, T. & Alfaro, M. 2001. Modified Cam-Clay modelling of temperature effects in clays.
Canadian Geotechnical J., 38(3): 608–621.
Guerriero, G., Picarelli, L. & Urciuoli, G. 1993. Some considerations on the ultimate condition of hard clays in
triaxial tests. Proc. Int. Symp. on Geotechnical Engineering of Hard Soils – Soft Rocks, Athens, Vol. 3: 1909–1911.
Hamouche, K. 1995. Compréhension du comportement d’un massif argileux soumis à des sollicitations horizon-
tales. Ph.D. Thesis, Université Laval, Québec, Canada.
Hamouche, K., Leroueil, S., Roy, M. & Lutenegger, A.J. 1995a. In situ evaluation of Ko in eastern Canada clays.
Canadian Geotech. J., 32(4): 677–688.
Hamouche, K., Roy, M. & Leroueil, S. 1995b. A pressuremeter study of the sensitive Louiseville clay. Proc. 4th Int.
Symp. on Pressuremeter, Sherbrooke, pp. 361–366.
Handy, R.L., Remmes, R., Moldt, S., Lutenegger, A.J. & Trott, G. 1982. In Situ stress determination by Iowa stepped
blade. Journ. of the Geot. Eng. Div., ASCE, 108 (GT11): 1405–1422.
Hardin, B.O. 1978. The nature of stress-strain behavior for soils. Proc. ASCE Geot. Div. Specialty Conf. on
Earthquake Engng. and Soil Dynamics, Pasadena, Vol. 1: 3–90.
Hatanaka, M. & Uchida, A. 1995. Effects of test methods on the cyclic deformation characteristics of high-quality
undisturbed gravel samples. Proc. ASCE, Geotechnical Special Publication. No. 56, pp. 136–151.
Hatanaka, M. & Uchida, A. 1996. Empirical correlation between penetration resistance and  of sandy soils. Soils
& Foundations, 36(4): 1–9.
Hatanaka, M., Uchida, A., Taya, Y., Takehara, N., Hagisawa, T., Sakou, N. & Ogawa, S. 2001. Permeability charac-
teristics of high-quality undisturbed gravelly soils measured in laboratory tests. Soils & Foundations, 41(3):
45–55.
Hazen, A. 1911. Discussion on “Dams on sand formations” by A.C. Koenig. Transactions of the ASCE, 73: 199–203.
Hicher, P.Y. 1988. The viscoplastic behaviour of bentonite. Proc. Int. Conf. on Rheology and Soil Mechanics,
Coventry, 89–107.
Hight, D.W. 1983. Laboratory investigations of sea bed clays. Ph.D. Thesis, University of London.
Hight, D.W. 1998. Soil characterisation: the importance of structure and anisotropy. 38th Rankine Lecture, London, U.K.
Hight, D.W. 2001. Sampling effects in soft clay: an update on Ladd and Lambe (1963). Ladd Retirement Symposium,
Massachusetts Institute of Technology, Cambridge, USA (to be published).
Hight, D.W., Jardine, R.J. & Gens, A. 1987. The behaviour of soft clays. Chapter 2 of Embankments on Soft Clays,
Bulletin of the Public Works Research Center of Greece, Athens: 33–38.
Hight, D.W., Böese, R., Butcher, A.P., Clayton, C.R.I. & Smith, P.R. 1992. Disturbance of the Bothkennar clay prior
to laboratory testing. Géotechnique, 42(2): 199–217.
Hight, D.W. & Higgins, K.G. 1994. An approach to the prediction of ground movements in engineering practice:
Background and application. Proc. Int. Symp. on Pre-Failure Deformation Characteristics of geomaterials –
Measurement and Application. IS-Hokkaido, Sapporo, Vol. 2: 909–945.
Hight, D.W., Georgiannou V.N. & Ford C.J. 1994. Characterisation of clayey sands. Proc. 7th Int. Conf. on Behaviour
of Offshore Structures, Boston, USA, Vol. 1, 321–340.
Hight, D.W., Bennell, J.D., Chana, B., Davis, P.D., Jardine, R.J. & Porovic, E. 1997. Wave velocity and stiffness meas-
urements at Sizewell. Proc. Symp. on Pre-Failure Deformation Behaviour of Geomateirals, London, pp. 65–88.
Hight, D.W., Georgiannou, V.N., Martin, P.L. & Mundegar, A.K. 1998. Flow slides in micaceous sands. Proc. Int.
Symp. on Problematic Soils, IS-Tohoku’98, Sendai, Japan. Vol. 2:945–958.
Hight, D.W., McMillan, F., Powell, J.J.M., Jardine, R. J. & Allenou., C.P. 2002a. Some characteristics of London Clay.
Proc. Int. Workshop on Characterisation and Engineering Properties of Natural Soils, Singapore.
Hight, D.W., Paul, M.A., Barras, B. F., Powell, J.J.M., Nash, D.F.T., Smith, P.R., Jardine, R.J. & Edwards, D.H. 2002b.
The characterization of the Bothkennar Clay. Proc. Int. Workshop on Characterisation and Engineering
Properties of Natural Soils, Singapore.

241
09031-02[2].qxd 18/Oct/02 12:13 AM Page 242

Hight, D.W. & Leroueil, S. 2002. Characterisation of soils for engineering purposes. Proc. Int. Workshop on
Characterisation and Engineering Properties of Natural Soils, Singapore.
Hird, C.C. & Hassona, F. 1986. Discussion on “A state parameter for sands”. Géotechnique, 36(1): 124–127.
Hoek, E. 1965. Rock fracture under static stress conditions. Rep. MEG 303, CSIR, Pretoria, South Africa.
Hoeg, K., Dyvik, R. & Sandbaekken, G. 2000. Strength of undisturbed versus reconstituted silt and silty sand spec-
imens. J. Geotechnical and Geoenvironmental Engng., ASCE, 126(7): 606–617.
Holzer, T.L., Hoeg, K. & Arulanandan, K. 1973. Excess pore pressures during undrained clay creep. Canadian
Geotechnical J., 10(1): 12–24.
Holtz, R.D., Jamiolkowski, M.B. & Lancellotta, R. 1986. Lessons from odeometer test on high quality samples.
J. Geotech. Engng Div., ASCE 112(GT8): 768–776.
Houston, W.N. & Mitchell, J.K. 1969. Property interrelationships in sensitive clays. J. Soil Mech. and Found Engng.,
ASCE, 95(4): 1037–1062.
Houston, S.L., Houston, W.N. & Williams, N.D. 1985. Thermo-mechanical behavior of seafloor sediments.
J. Geotech. Engrg. Div., ASCE, 111(11): 1249–1263.
Huang, J.T. & Airey, D.W. 1993. Effects of cement and density on an artificially cemented sand. Proc. 1st Int. Symp.
on the Geotechnics of Hard Soils – Soft Rocks, Athens, Vol. 1: 553–560.
Hvorslev, M.J. 1937. Uber die Festigkeitseigenschaften Gestorter Bindiger Boden. Danmarks Naturvidenskabelige
Samfund. Ingeniorvidenskabelige Skrifter, A., No. 45.
Hueckel, T. & Pellegrini, R. 1989. Modelling of thermal failure of saturated clays. Numerical Models in
Geomechanics, Elsevier, New York: 81–90.
Hueckel, T. & Baldi, G. 1990. Thermoplasticity of saturated clays: Experimental constitutive study. J. Geotech.
Engrg. Div., ASCE, 116(12): 1778–1796.
Hungr, O. & Morgenstern, N.R. 1984. High velocity ring shear tests on sand. Géotechnique, 34(3): 415–421.
Hvorslev, M.J. 1937. Über die Festigkeit-seigenschaften gestörter bindiger Böden Ingeniorvidenskabelige Skrifter
A No. 45.
Imai, G. 1995. Analytical examinations of the foundations to formulate consolidation phenomena with inherent
time-dependence. Proc. Int. Symp. on Compression and Consolidation of Clayey Soils – IS-Hiroshima’s ’ 95,
Hiroshima, 2: 891–935.
Imai, G. & Tang, Y. 1992. A constitutive equation of one-dimensional consolidation derived from inter-connected
tests. Soils & Foundations, 32(2): 83–96.
Isenhower, W.M., & Stokoe, K.H. 1981. Strain-rate dependent shear modulus of San Francisco Bay Mud. Proc. Int.
Conf. on Recent Advances in Geotechnical Earthquake Engineering and Soil Dynamics. St. Louis, 2, 597–602.
Ishihara, K. 1993. Liquefaction and flow failure during earthquakes. Géotechnique, 43(3): 351–415.
Ishihara, K. 1996. Soil Behavior in Earthquake Geotechnics. Edit. Clarendon Press, Oxford.
Ishihara, K., Tatsuoka, F. & Yasuda, S. 1975. Undrained deformation and liquefaction of sand under cyclic stresses.
Soils & Foundations, 15: 29–44.
ISSMFE, 1989. Appendix A: “International reference test procedure for cone penetration test (CPT)”. Report of the
ISSMFE Technical Committee on Penetration Testing of Soils – TC 16, with reference to Test Procedures,
Swedish Geotechnical Institute, Linköping, Information, 7, 6–16.
Jaky, J. 1944. A nyugalmi nyomas tenyezoje (the coefficient of earth pressure at rest). J. Soc. of Hungarian Architects
and Engineers, Budapest, pp. 355–358.
Jamiolkowski, M., Ladd, C.C., Germaine, J.T. & Lancellotta, R. 1985. New developments in field and laboratory
testing of soils. Proc.11th Int. Conf. on Soil Mech. and Found. Engrg., San Francisco, Vol. 1, pp. 57–153.
Jamiolkowski, M., Leroueil, S. & Lo Presti, D.C.F. 1991. Theme Lecture: Design parameters, from theory to prac-
tice. Proc. Geo-Coast ’91, Yokohama. Vol. 2: 877–917.
Jamiolkowski, M., Lancellotta R. & Lo Presti D.C.F. 1994. Remarks on the stiffness at small strains of six Italian
clays. Proc. First Int. Conf. on Pre-failure Deformation Characteristics of Geomaterials, Sapporo, Japan, Vol. 2;
817–836.
Jamiolkowski, M., Lancellotta, R. & Lo Presti, D.C.F. 1995a. Remarks on the stiffness at small strains of six Italian
clays. Keynote Lecture 3, Proc. Int. Symp. on Pre-Failure Deformation Characteristics of Geomaterials,
Sapporo, Vol. 2: 817–836.
Jamiolkowski, M., Lo Presti, D.C.F. & Pallara, O. 1995b. Role of in-situ testing in geotechnical earthquake engi-
neering. Proc. 3rd Int. Conf. on Recent Advances in Geotech. Earthquake Engrg. and Soil Dynamic. State-of-the-
Art 7, Vol. 3: 1523–1546.
Jamiolkowski, M. & Lo Presti, D.C.F. 1998. Geotechnical characterization of gravelly deposits. Proc. 13th South-
East Asian Geotechnical Conf., Taipei.
Jamiolkowski, M. & Lo Presti, D.C.F. 2002. Geotechnical characterization of Holocene and Pleistocene Messina
sand and gravel deposits. Proc. Int. Workshop on Characterisation and Engineering Properties of Natural Soils,
Singapore.
Jang, D.-J. & Frost, J.D. 1998. Sand structure differences resulting from specimen preparation procedures. Proc.
ASCE Specialty Conf. on Geotechnical Earthquake Engineering and Soil Dynamics, Seattle, USA.
Jardine, R.J. 1985. Investigations of pile-soil behaviour with special reference to the foundations of offshore struc-
tures. Ph.D. Thesis, University of London, London, U.K.
Jardine, R.J. 1992. Some observations on the kinematic nature of soil stiffness. Soils & Foundations, 32(2):
111–124.

242
09031-02[2].qxd 18/Oct/02 12:13 AM Page 243

Jardine, R.J., Symes, M.J. & Burland, J.B. 1984. The measurement of soil stiffness in the triaxial apparatus.
Géotechnique, 34(3): 323–340.
Jardine, R.J., St-John, H.D., Hight, D.W. & Potts, D.M. 1991. Some practical applications of a non-linear ground
model. Proc. 10th European Conf. on Soil Mech. and Found. Engng., Florence, Vol. 1: 223–228.
Jardine, R.J., St. John, H.D., Hight, D.W. & Potts, D.M. 1991. Some practical applications of a non-linear ground
model. Proc.10th European Conf. on Soil Mech. and Found. Engrg., Florence, 1: 223–228.
Jardine, R.J., Zdravkovic, L. & Porovic, E. 1997. Anisotropic consolidation including principal stress axis rotation:
Experiments, results and practical implications. Proc. 14th Int. Conf. on Soil Mech. and Found. Engrg., Hamburg,
Germany, Vol. 4, 2165–2168, Panel contribution.
Jardine, R.J., Kuwano R., Zdravkovic L., & Thornton C. 2001. Some fundamental aspects of the pre-failure behav-
iour of granular soils. Proc. 2nd Int. Conf. on Pre-failure Deformation Characteristics of Geomaterials, Torino,
Italy, Vol. 2, 1077–1111
Jardine, R.J., Smith, P.R. & Nicholson, D.P. 2002. Properties of the soft Holocene Thames Estuary Clay from
Queenborough, Kent. Proc. Int. Workshop on Characterisation and Engineering Properties of Natural Soils,
Singapore.
Jennings, J.E.B & Burland, J.B. 1962. Limitations to the use of effective stresses in partly saturated soil.
Géotechnique, 12(2): 125–144.
Jovicic, V. & Coop, M. 1997. The influence of plastic strains during compression on the inherent anisotropy of Gmax.
Proc. Symp. on Pre-failure Deformation Behaviour of Geomaterials, London, pp. 395–397.
Kabbaj, M. 1985. Aspects rhéologiques des argiles naturelles en consolidation. Ph.D. Thesis, Université Laval, Québec.
Kabbaj, M., Tavenas, F. & Leroueil, S. 1988. In situ and laboratory stress-strain relations. Géotechnique, 38(1):
83–100.
Karube, D. 1988. New concept of effective stress in unsaturated soil and its proving test. In Advanced Triaxial Testing
of Soil and Rock, ASTM, STP 977: 539–552.
Karube, D. & Kato, S. 1989. Yield functions of unsaturated soil. Proc. 1th Int. Conf. on Soil Mech. and Found.
Engrg., Rio de Janeiro, 1: 615–618.
Kavvadas, M.J. 1998. General Report: Modelling the soil behaviour – Selection of soil parameters. Proc. 2nd
Int. Symp. on the Geotechnics of Hard Soils – Soft Rocks, Naples, Vol. 3: 1441–1481.
Kavvadas, M. & Amorosi, A. 1998. A plasticity approach for the mechanical behaviour of structured soils. Proc. 2nd
Int. Symp. on the Geotechnics of Hard Soils – Soft Rocks, Napoli, Vol. 2: 603–613.
Kavvadas, M., Anagnostopoulos, A. & Kalteziontis, N. 1993. A framework for the mechanical behaviour of
the cemented Corinth marl. Proc. Int. Symp. on Geotechnical Engineering of Hard Soils-Soft Rocks, Athens,
1: 577–583.
Kavvadas, M.J., Anagnostopoulos, A.G., Georgiannou, V.N. & Bardanis, M.E. 2002. Characterisation and engineering
properties of the Corinth marl. Proc. Int. Workshop on Characterisation and Engineering Properties of Natural
Soils, Singapore.
Kawai, K., Kato, S. & Karube, D. 2000. The model of water retention curve considering effects of void ratio. Proc.
Asian Conf. on Unsaturated Soils, Singapore, pp. 329–334.
Kenney, T.C. 1963. Permeability ratio of repeatedly layered soils. Géotechnique, 13: 325–333.
Kenney, T.C. 1967. The influence of mineral composition on the residual strength of natural soils. Proc. of the
Geotechnical Conference on the Shear Strength properties of Natural Soils and Rocks, Oslo, Vol. 1: 123–129.
Kenney, T.C., Lau, D. & Ofoegbu, G.I. 1984. Permeability of compacted granular materials. Canadian Geotechnical J.,
21(4): 726–729.
Khalili, N. & Khabbaz, M.H. 1998. A unique relationship for the determination of the shear strength of unsaturated
soils. Géotechnique, 48(5): 681–687.
Kirkgard, M.M. & Lade, P.V. 1991. Anisotropy of normally consolidated San Francisco Bay Mud. Geotechnical
Testing J., 14(3): 231–246.
Koerner, R.M. 1970. Effect of particle characteristics on soil strength. J. Soil Mech. and Found. Engng., ASCE,
96(SM4): 1221–1234.
Kohata, Y., Tatsuoka, F., Wang, L., Jiang, G.L., Hoque, E. & Kodaka, T. 1997. Modelling the non-linear deformation
properties of stiff materials. Géotechnique, 47(3): 563–580.
Kokusho, T., Yoshida, Y., & Esashi, Y. 1982. Dynamic properties of soft clay for a wide strain range. Soils &
Foundations, 22(4): 1–29.
Kondner, R.L. 1963. Hyperbolic stress-strain response: cohesive soils. J. Soil Mech. and Found. Engng, ASCE, 89(1):
115–143.
Konrad, J.-M. 1990. Minimum undrained strength versus steady-state strength of sands. J. Geotech. Engrg. Div.,
ASCE, 116(6): 948–963.
Konrad, J.-M. 1993. Undrained response of loosely compacted sands during monotonic and cyclic compression tests.
Géotechnique, 43(1): 69–89.
Konrad, J.M. 1997. In situ sand state from CPT: evaluation of a unified approach at two CANLEX sites. Canadian
Geotechnical J., 34(1): 120–130.
Konrad, J.-M., Bozozuk, M. & Law, K.T. 1985. Study of in-situ test methods in deltaic silt. Proc. 11th Int. Conf. on
Soil Mech. and found. Engng., San Francisco, Vol. 2: 879–886.
Konrad, J.-M. & Law, K.T. 1987. Preconsolidation pressure from piezocone tests in marine clays. Géotechnique,
37(2): 177–190.

243
09031-02[2].qxd 18/Oct/02 12:13 AM Page 244

Kuerbis, R., Negussey, D. & Vaid, Y.P. 1988. Effect of gradation and fines content on the undrained response of sand.
ASCE Specialty Conf. on Hydraulic Fill Structures, Colorado, USA, 330–345.
Kulhawy, F.H. & Mayne, P.W. 1990. Manual of estimating soil properties for foundation design. Geotechnical
Engineering Group, Cornell University, Ithaca.
Lacasse, S.M., Ladd, C.C. & Baligh, M.M. 1978. Evaluation of field vane, Dutch cone penetrometer and piezome-
ter testing devices, MIT Research Report.
Lacasse, S., Jamiolkowski, M., Lancellotta, R. & Lunne, T. 1981. In situ characteristics of two Norwegian clays.
Proc.10th Int. Conf. on Soil Mech. and Found. Engng., Stockholm, Vol. 2: 507–511.
Lacasse, S., Berre, T. & Lefebvre, G. 1985. Block sampling of sensitive clays. Proc. 11th Int. Conf. on Soil Mech.
and Found. Engrg, San Francisco, 2: 887–892.
Lacerda, W.A. 1976. Stress-relaxation and creep effects on soil deformation. Ph.D. Thesis, University of California,
Berkeley, U.S.A.
Lacerda, W.A. 1997. Stability of natural slopes along the tropical coast of Brazil. Proc, Symp. on Recent
Developments in Soil and Pavement Mechanics, Rio de Janeiro, pp. 17–39.
Lacerda, W.A. & Houston, W.N. 1973. Stress relaxation in soils. Proc. 8th Int. Conf. on Soil Mech. & Found. Engrg.,
1, Moscow, 221–227.
Lacerda, W.A. & Houston, W.N. 1978. Stress relaxation and creep effects on soil deformation. Ph.D. Thesis,
University of California, Berkeley, U.S.A.
Ladanyi, B. 1960. Étude des relations entre les contraintes et les déformations lors du cisaillement des sols pulvéru-
lents. Annales Trav. Publ. Belg., No. 3: 1–30.
Ladd, C.C. 1969. The prediction of in situ stress-strain behaviour of soft saturated clay during undrained shear. Proc.
Bolkesjö Symp., Norway, Norwegian Geotechnical Inst., pp. 14–20.
Ladd, C.C. 1991. Stability evaluation during staged construction. J. Geotech. Engng. Div., ASCE, 117(4): 540–615.
Ladd, C.C. & Varallyay, J. 1965. The influence of stress system on the behavior of saturated clays during undrained
shear. Research in Earth Physics Phase, Report No 1, Part II, Report R65-11, Massachusetts Institute of
Technology, p. 263.
Ladd, C.C. & Foott, R. 1974. New design procedure for stability of soft clays. J. of Geotech. Engng. Div., ASCE,
100(7): 763–786.
Ladd, C.C., Foott, R., Ishihara, K., Schlosser, F. & Poulos, H.G. 1977. Stress-deformation and strength characteris-
tics. State-of-the-Art Report, Proc. 9th Int. Conf. on Soil Mech. & Found. Engrg., Tokyo, Vol. 2, pp. 421–494.
Lade, P.V. 1993. Initiation of static instability in the submarine Nerlerk berm. Canadian Geotechnical J., 30(6):
895–904.
Lade, P.V. 1995. Instability of sand in the prefailure hardening regime, Proc.1st Int. Conf. on Pre-Failure
Deformation Characteristics of Geomaterials, Sapporo, Vol.2: 837–854.
Lade, P.V. 2002. Personal communication.
Lade, P.V. & Duncan, J.M. 1973. Cubical triaxial tests on cohesionless soil. J. of Soil Mech. and Found. Div., ASCE,
99(SM10): 793–812
Lade, P.V. & Duncan, J.M. 1975. Elastoplastic stress-strain theory for cohesionless soil. J. of Geotech. Div., ASCE,
101(10): 1037–1053.
Lade, P.V. & Kirkgard, M.M. 2000. Effects of stress rotation and changes in b values on cross-anisotropic behavior
of natural, Ko-consolidated soft clay. Soils & Foundations, 40(6): 93–105.
Lagoia, R. & Nova, R. 1995. An experimental and theoretical study of the behaviour of a calcarenite in triaxial com-
pression. Géotechnique, 45(4): 633–648.
Lam W.K. & Tatsuoka F. 1988. Effects of initial anisotropic fabric and 2 on strength and deformation characteris-
tics of sand. Soils & Foundations, 28(1): 89–106.
Lambe, T.W. 1967. Stress path method. J. Soil Mech., ASCE. 93(SM6): 309–331.
Lambe, T.W. 1973. Predictions in soil engineering. 13th Rankine Lecture. Géotechnique, 23(2):151–201.
Lancellotta, R. 1983. Reliability analysis in geotechnical engineering. Istituto Scienza delle Costruzioni, 625,
Politenico di Torino, Torino.
Lapierre, C., Leroueil, S. & Locat, J. 1990. Mercury intrusion and permeability of Louiseville clay. Canadian
Geotech. J., 27(6), 761–773.
La Rochelle, P., Sarrailh, J., Tavenas, F., Roy, M. & Leroueil, S. 1981. Causes of sampling disturbance and design of
a new sampler for sensitive soils. Canadian Geotechnical J., 18(1): 52–66.
Larsson, R. 1977. Basic behaviour of Scandinavian soft clays. Swedish Geotechnical Institute, Link öping, Report
No. 4.
Larsson, R. & Mulabdic, M. 1991. Piezocone tests in clay. Report No. 42, Swedish Geotechnical Institute,
Linköping, 240p.
Le Bihan, J.-P. & Leroueil, S. 2001. Simulation d’écoulement en milieu non saturé. Application au Barrage CD-00
de LG-2. Report CGT-01-11 prepared for Hydro-Québec, Université Laval, Québec, Canada.
Le Bihan, J.-P. 2002. Personal communication.
Leblond, P. 2002. Personal communication.
Lee, H.J., Ellen, S.D. & Kayen, R.E. 1988. Predicting transformation of shallow landslides into high-speed debris
flows. Proc. 5th Int. Symp. on Landslides, Lausanne, Vol. 1: 713–718.
Lee, K.L. 1965. Triaxial compressive strength of saturated sand under seismic loading conditions. Ph.D. Thesis,
University of California, Berkeley.

244
09031-02[2].qxd 18/Oct/02 12:13 AM Page 245

Lee, K.L. & Seed, H.B. 1967. Drained strength characteristics of sand. J. Soil Mech. and Found. Div., ASCE, 93(6):
117–141.
Lefebvre, G. 1970. Contribution à l’étude de la stabilité des pentes dans les argiles cimentées. Ph.D. Thesis,
Université Laval, Québec.
Lefebvre, G. 1981. Fourth Canadian Geotechnical Colloquium: Strength and slope stability in Canadian soft clay
deposits. Canadian Geotechnical J., 18, 420–442.
Lefebvre, G. & Leboeuf, D. 1987. Rate effects and cyclic loading of sensitive clays. J. Geotech. Engrg. Div., ASCE,
113(5): 476–489.
Lefebvre, G., Bozozuk, M., Philibert, A. & Hornych, P. 1991. Evaluating Ko in Champlain clays with hydraulic frac-
ture tests. Canadian Geotechnical J., 28(3): 365–377.
Lemos, L.J.L. 1986. The effect of rate of shear on residual shear strength of soil. Ph.D. Thesis, University of london,
London, U.K.
Leonards, G.A. & Altschaeffl, A.G. 1964. Compressibility of clay. J. of the Soil Mech. and Found. Engrg. Div.,
ASCE, 90(5): 133–155.
Leong, E.C. & Rahardjo, H. 1997. Review of soil-water characteristic equations. J. Geotech. and Geoenvironmental
Engineering, ASCE, 123(2): 1106–1117.
Leong, E. C., Rahardjo, H. & Tang, S. K. 2002. Characterisation and engineering properties of Singapore residual
soils. Proc. Int. Workshop on Characterisation and Engineering Properties of Natural Soils, Singapore.
Leong, W.K. 2001. Instability behaviour of a granular fill material. Ph.D. Thesis, Nanyang Technological University,
Singapore.
Leroueil, S. 1988. Tenth Canadian Geotechnical Colloquium: Recent developments in consolidation of natural clays.
Canadian Geotechnical J., 25(1): 85–107.
Leroueil, S. 1992. A framework for the mechanical behavior of structured soils, from soft clays to weak rocks. Proc.
US-Brazil NSF Geotechnical Workshop on Applicability of Classical Soil Mechanics Principles to Structured
Soils, Belo Horizonte: 107–128.
Leroueil, S. 1996. Compressibility of clays: fundamental and practical aspects. J. Geotech. Engrg. Div., ASCE,
122(7): 534–543. Also partially published in the Proc. of the ASCE Conf. on Vertical and Horizontal Deformations
of Foundations and Embankments, Settlement’s 94, College Station, 1: 57–76.
Leroueil, S. 1997. Critical state soil mechanics and the behaviour of real soils. Proc.Int. Symp. on Recent Developments
in Soil and Pavement Mechanics, Rio de Janeiro, pp. 41–80.
Leroueil, S. 1998. Elements of time-dependent mechanical behaviour of overconsolidated clays. Proc. 51st
Canadian Geotech. Conf., Edmonton, Vol. 2: 671–677.
Leroueil, S. 2000. Contribution to the Round Table: Peculiar aspects of structured soils. Proc. 2nd Symp. on hard
Soils – Soft Rocks, Naples, Vol. 3. pp. 1669–1678.
Leroueil, S. 2001. Natural slopes and cuts: movement anf failure mechanisms. Géotechnique, 51(3): 197–243.
Leroueil, S. & Tavenas, F. 1979. Strain rate behaviour of Saint-Jean-Vianney clay. Discussion. Canadian
Geotechnical J., 16(3): 616–620.
Leroueil, S., Tavenas, F., Brucy, F., La Rochelle, P. & Roy, M. 1979. Behaviour of destructured natural clays. J. of
Geotech. Engrg., ASCE, 105(6): 759–778.
Leroueil, S., Tavenas, F. & Le Bihan, J.P. 1983a. Propriétés caractéristiques des argiles de l’est du Canada. Canadian
Geotechnical J., 20(4): 681–705.
Leroueil, S., Tavenas, F., Samson, L. & Morin, P. 1983b. Preconsolidation pressure of Champlain clays – Part II:
Laboratory determination. Canadian Geotechnical J., 20(4): 803–816.
Leroueil, S., Kabbaj, M., Tavenas, F. & Bouchard, R. 1985a. Stress-strain-strain rate relation for the compressibility
of sensitive natural clays. Géotechnique, 35: 159–180.
Leroueil, S., Tavenas, F. & Locat, J. 1985b. Discussion on “Correlations between index tests and the properties of
remoulded clays” by Carrier & Beckman. Géotechnique, 35(2): 223–226.
Leroueil, S., Magnan, J.-P. & Tavenas, F. 1985c. Remblais sur Argiles Molles. Technique et Documentation,
Lavoisier. Also in English, “Embankments on Soft Clays” (1990), Ellis Horwood Publisher.
Leroueil, S., Kabbaj, M. & Tavenas, F. 1988. Study of the validity of a v  v  v* model in in situ conditions.
Soils & Foundations, 28(3): 3–25.
Leroueil, S. & Vaughan, P.R. 1990. The general and congruent effects of structure in natural soils and weak rocks.
Géotechnique, 40(3): 467–488.
Leroueil, S., Bouclin, G., Tavenas, F., Bergeron, L. & La Rochelle, P. 1990. Permeability anisotropy of natural clays
as a function of strain. Canadian Geotechnical J., 27(5): 568–579.
Leroueil, S., Lerat, P., Hight, D.W. & Powell, J.J.M. 1992. Hydraulic conductivity of a recent estuarine silty clay at
Bothkennar, Scotland. Géotechnique, 42(2): 275–288.
Leroueil, S., Demers, D., La Rochelle, P., Martel, G. & Virely, D. 1995. Practical applications of the piezocone in
Champlain sea clays. Proc. Int. Symp. on Cone Penetration Testing, CPT-95, Linköping, pp. 515–522.
Leroueil, S. & Marques, M.E.S. 1996. Importance of strain rate and temperature effects in geotechnical engineering.
Session on Measuring and Modeling Time Dependent Soil Behavior, ASCE Convention, Washington, Geot.
Special Publication 61: 1–60.
Leroueil, S., Perret, D. & Locat, J. 1996. Strain rate and structuring effects on the compressibility of a young clay.
Session on Measuring and Modeling Time Dependent Soil Behavior, ASCE Convention, Washington, Geot.
Special Publication 61: 137–150.

245
09031-02[2].qxd 18/Oct/02 12:13 AM Page 246

Leroueil, S., Guerriero, G., Picarelli, L. & Saihi, F. 1997. Large deformation shear strength of two types of structured
soils. Proc. Symp. on Deformation and Progressive Failure in Geomechanics, IS-Nagoya, Nagoya, pp. 217–222.
Leroueil, S. & Barbosa, P.S de A. 2000. Combined effect of fabric, bonding and partial saturation on yielding of
soils. Proc. Asian Conf. on Unsaturated Soils, Singapore, pp. 527–532.
Leroueil, S., Hamouche, K., Tavenas, F., Boudali, M., Locat, J., Virely, D., Roy, M., La Rochelle, P. & Leblond P.
2002. Geotechnical characterization and properties of a sensitive clay from Québec. Proc. Int. Workshop on
Characterisation and Engineering Properties of Natural Soils, Singapore.
Levadoux, J.-N. & Baligh, M.M. 1986. Consolidation after undrained piezocone penetration. J. of Geotechnical
Engng., ASCE, 112(7): 707–726.
Lingnau, B.E., Graham, J., & Tanaka, N. 1995. Isothermal modeling of sand-bentonite mixtures at elevated temper-
atures. Canadian Geotechnical J., 31(1): 78–88.
Lo K.Y. 1965. Stability of slopes in anisotropic soils. J. of Soil Mech. and Found. Div., ASCE, 91(SM4): 85–106.
Lo, K.Y. & Morin, J.P. 1972. Strength anisotropy and time effects of two sensitive clays. Canadian Geotechnical J.,
9(3): 261–277.
Locat, J. 1982. Contribution à l’étude de l’origine de la structuration dans les angiles sensibles dans l’est du
Canada. Ph.D. Thesis, Université de Sherbrooke, Québec, Canada.
Locat, J. & Lefebvre, G. 1982. The compressibility and sensitivity of an artificially sedimented clay soil: the Grande
Baleine marine clay, Québec. Proc. 2nd Canadian Conf. on Marine Geotechnique, Bedford, Session 2, Paper
No. 4.
Locat, J. & Lefebvre, G. 1985. The compressibility and sensitivity of an artificially sedimented clay soil: the Grande
Baleine marine clay, Québec. Marine Geotechnology, 6(1): 1–27.
Locat, J, Tanaka, H., Tan T.S & Dasari, G.R. 2002. Natural soils: Geotechnical behavior and geological knowledge.
Proc. Int. Workshop on Characterisation and Engineering Properties of Natural Soils, Singapore.
Long, M. 2002. Characterization and engineering properties of Athlone laminated clay. Proc. Int. Workshop on
Characterisation and Engineering Properties of Natural Soils, Singapore.
Lo Presti, D.C.F. 1989. Proprietà dinamiche dei terreni. Atti delle Conferenze di Geotecnica del Politecnico di Torino.
Lo Presti, D.C.F., Jamiolkowski, M., Pallara, O. & Cavallaro, A. 1996. Rate and creep effect on the stiffness of soils.
Session on Measuring and Modeling Time Dependent Soil Behavior, ASCE Convention, Washington, Geot.
Special Publication 61:166–180.
Lo Presti, D.C.F., Jamiolkowski, M., Pallara, O., Cavallaro, A. & Pedroni, S. 1997. Shear modulus and damping of
soils. Géotechnique, 47(3): 603–617.
Lo Presti, D.C.F., Pallara, O., Jamiolkowski, M. & Cavallaro, A. 1999. Anisotropy of small strain stiffness of undis-
turbed and reconstituted clays. Proc. 2nd Int. Symp. on Pre-Failure Deformation Characteristics of Geomaterials,
IS-Torino 99, Torino, Vol. 1: 11–18.
Lo Presti, D.C.F., Shibuya, S. & Rix, G.J. 2001. Innovation in soil testing. Proc. Symp. on Pre-failure Characteristics
of Geomaterials, Torino, Vol. 2: 1027–1076.
Lo Presti, D.C.F., Jamiolkowski, M. & Pepe, M. 2002. Geotechnical characterisation of the subsoil of Pisa Tower.
Proc. Int. Workshop on Characterisation and Engineering Properties of Natural Soils, Singapore.
Loudon, A.G. 1952. The computation of permeability from simple soil tests. Géotechnique, 3: 165–183.
Lowe, J.L. III, 1974. New concepts in consolidation and settlement analysis. J. Geotech. Engrg. Div., ASCE, 100(6):
574–612.
Lowe, J. III, Jones, E., Obrician, V. 1969. Controlled gradient consolidation test. J. Soil Mech. and Found. Div.,
ASCE, 95(SM1):77–97.
Lunne, T., Robertson, P.K. & Powell, J.J.M. 1997. Cone Penetration Testing in Geotechnical Practice. Blackie
Academic & Professional (An imprint of Chapman & Hall).
Lunne, T., Long, M. & Forsberg, C. F. 2002a. Characterization and engineering properties of Holmen, Drammen
sand. Proc. Int. Workshop on Characterisation and Engineering Properties of Natural Soils, Singapore.
Lunne, T., Long, M. & Forsberg, C. F. 2002b. Characterization and engineering properties of Onsøy clay. Proc. Int.
Workshop on Characterisation and Engineering Properties of Natural Soils, Singapore.
Lupini, J.F. 1980. The residual strength of soils. Ph.D. Thesis, University of London, U.K.
Lupini, J.F., Skinner, A.E. & Vaughan, P.R. 1981. The drained residual strength of cohesive soils. Géotechnique,
31(2): 181–213.
Maâtouk, A., Leroueil, S. & La Rochelle, P. 1995. Yielding and critical state of a collapsible unsaturated silty soil.
Géotechnique, 34(3): 465–477.
Maccarini, M. 1987. Laboratory studies of weakly bonded artificial soil. Ph.D. Thesis, University of London, U.K.
Machado, S. L. & Vilar, O. M. 2002. Geotechnical characteristics of an unsaturated soil deposit at São Carlos, Brazil.
Proc. Int. Workshop on Characterisation and Engineering Properties of Natural Soils, Singapore.
Manassero, M. 1994. Hydraulic conductivity assessment of slurry wall using piezocone test. J. Geotechnical
Engng., ASCE, 120(10): 1725–1746.
Marchand, G. 1982. Quelques considérations sur le comportement avant rupture des pentes argileuses naturelles.
M.Sc. Thesis, Université Laval, Québec, Canada.
Marchetti, S. 1980. In situ tests by flat dilatometer. J. Geot. Eng. Div. ASCE, Vol. 106 (GT3): 299–321.
Marques, M.E.S. 1996. Influência da velocidade de deformaçao O e da temperatura no adensamento de argilas nat-
urais. M.Sc. Thesis. Research performed at Université Laval, Ste-Foy, Canada in cooperation with COPPE –
Federal University of Rio de Janeiro, Brazil.

246
09031-02[2].qxd 18/Oct/02 12:13 AM Page 247

Marques, M.E.S., Leroueil, S. & Almeida, M. de S.S. (2002). Viscous behaviour of St-Roch-de-l’Achigan clay,
Québec. Canadian Geot. J., Submitted for publication.
Marshall, T.J. 1958. A relation between permeability and size distribution of pores. J. of Soil Sciences, 9(1): 1–9.
Martins, I.S.M. 1992. Fundamentos de um modelo de comportamento de solos argilosos. D.Sc. Thesis, COPPE-
Universidade Federal do Rio de Janeiro, Rio de Janeiro, Brazil.
Massarsch, K.R., Holtz, R.D., Holm, B.G. & Fredriksson, A. 1975. Measurement of horizontal in situ stresses. Proc.
ASCE Conf. On In Situ Measurement of Soil Properties, Raleigh, Vol. 1, pp. 266–286.
Matsuoka, H. & Nakai, T. 1974. Stress deformation and strength characteristics under three different principal
stresses. Proc. Japan Soc. Civil Engng., 232: 59–70.
Matthews, M.C., Clayton, C.R.I. & Own, Y. 2000. The use of field geophysical techniques to determine geotechni-
cal stiffness parameters. Geotechnical Engineering, 143(1): 31–42.
Matyas, E.L. & Radhakrishna, H.S. 1968. Volume change characteristics of partially saturated soils. Géotechnique,
18(4): 432–448.
Mayne, P.W. 1991. Determination of OCR in clays by piezocone tests using cavity expansion and critical state con-
cepts. Soils & Foundations, 31(2): 65–76.
Mayne, P.W. 2001. Stress-strain-strength-flow parameters from enhanced in-situ tests. Proc. Int. Conf. on In-Situ
Measurement of Soil Properties & Case Histories, Bali, Indonesia, pp. 27–48.
Mayne, P.W. & Kulhawy, F.H. 1982. Ko – OCR relationships in soil. J. of the Geotech. Engng. Div., ASCE, 108(GT6);
851–872.
Mayne, P.W. & Holtz, R.D. 1988. Profiling stress historyfrom piezocone soundings. Soils & Foundations, 28(1):
16–28.
Mayne, P.W. & Brown, D.A. 2002. Site characterization of Piedmont residuum of North America. Proc. Int.
Workshop on Characterisation and Engineering Properties of Natural Soils, Singapore.
McCinty, K., Karstunen, M. & Wheeler, S.J. 2001. Modelling the stress-strain behaviour of Bothkennar clay. Proc.
3rd Int. Conf. on Soft Soil Engineering, Hong Kong, 263–268.
Menkiti, C.O. 1995. Behaviour of clay and clayey sand, with particular reference to stress rotation. PhD Thesis,
Imperial College of Science, Technology and Medicine, University of London, UK.
Mesri, G. 1985. The uniqueness of the end-of-primary (EOP) void ratio – effective stress relationship. Proc. 1th
Int. Conf. On Soil Mech. And Found. Engrng., San Francisco, Vol.2, pp.587–590.
Mesri, G. 1987. The fourth law of soil mechanics: the law of compressibility. Proc. Int. Symp. on Geotechnical
Engineering of Soft Soils, Mexico City, 2: 179–187.
Mesri, G. & Olson, R.E. 1970. Shear strength of montmorillonite. Géotechnique, 20(3): 261–270.
Mesri, G. & Olson, R.E. 1971. Consolidation characteristics of montmorillonite. Géotechnique, 21 (4): 341–352.
Mesri, G. & Rokhsar, A. 1974. Theory of consolidaiton for clays. J. Geotech. Engng Div., ASCE, 100(GT8):
889–904.
Mesri, G., Rokhsar, A., & Bohor, B.F. 1975. Composition and compressibility of typical samples of Mexico City
clay. Géotechnique, 25(3): 527–554.
Mesri, G., & Godlewski, P.M. 1977. Time and stress compressibility interrelationships. J. Geotech. Enging Div.,
ASCE, 103(GT5): 417–430.
Mesri, G., Ullrich, C.R. & Choi, Y.K. 1978. The rate of swelling of overconsolidated clays subjected to unloading.
Géotechnique, 28: 281–307.
Mesri, G., & Hayat, T.M. 1993. The coefficient of earth pressure at rest. Canadian Geotechnical J., 30(4): 647–666.
Mesri, G., Lo, D.O.K. & Feng, T.W. 1994. Settlement of embankments on soft clays. Specialty Conf. On Vertical and
Horizontal Deformations of Foundations and Embankments, ASCE, Settlement ’94, College Station, Vol. 1: 8–56.
Mesri, G., Shahien, M. & Feng, T.W. 1995. Compressibility parameters during primary consolidation. Proc. Int.
Symp. on Compression and Consolidation of Clayey Soils – IS-Hiroshima’s 95, Hiroshima, 2: 1021–1037.
Mimura, M. 2002. Characteristics of some Japanese natural sands – Data from undisturbed frozen samples. Proc.
Int. Workshop on Characterisation and Engineering Properties of Natural Soils, Singapore.
Mitchell, J.K. 1964. Shearing resistance of soils as a rate process. J. Soil Mech. and Found. Engrg. Div., ASCE,
90(1): 29–61.
Mitchell, J.K. 1993. Fundamentals of soil behavior. 2nd Ed., John Wiley & Sons, Inc.
Mitchell, J.,K., Greenberg, J.A. & Witherspoon, P.A. 1973. Chemico-osmotic effects in fine-grained soils. J. Soil
Mech. Fdn Div., ASCE, 99(SM4), 307–322.
Mitchell, R.J. & Wong, P.K.K. 1973. The generalized failure of an Ottawa area Champlain Sea clay. Canadian Geot.
J. 10(4): 607–616.
Mitchell, J.K. & Solymar, Z.V. 1984. Time-dependent strength gain in freshly deposited or densified sand. J. of
Geotech. Engrg., ASCE, 110(11): 1559–1576.
Miura, S. & Yagi, K. 2002. Mechanical behavior and particle crushing of volcanic coarse-grained soils in Japan.
Proc. Int. Workshop on Characterisation and Engineering Properties of Natural Soils, Singapore.
Morgenstern, N.R. & Tchalenko, J.S. 1967. Microscopic structures in kaolin subjected to direct shear. Géotechnique,
17: 309–328.
Moritz, L. 1995. Geotechnical properties of clay at elevated temperatures. Swedish Geotechnical Institute,
Linköping, Report no. 47.
Mulilis, P.J., Seed, H.B., Chan, C., Mitchell, J.K. & Arulanandan, K. 1977. Effect of sample preparation on sand liq-
uefaction. J. Geotech. Engng., ASCE, 103(2): 91–109.

247
09031-02[2].qxd 18/Oct/02 12:13 AM Page 248

Murayama, S. & Shibata, T. 1961. Rheological properties of clays. Proc. 5th Int. Conf. on Soil Mech. and Found.
Engrg., Paris: 269–273.
Murayama, S., Michihiro, K. & Sakagami, T. 1984. Creep characteristics of sands. Soils & Foundations, 24(2): 1–15.
Nagaraj, T.S. & Srinivasa Murthy, B.R. 1986. A critical appraisal of compression index equations. Géotechnique,
36(1): 27–32.
Nash, D.F.T., Lings, M.L. & Pennington, D.S. 1999. The dependence of anisotropic Go shear moduli on void ratio
and stress state for reconstituted Gault clay. Proc. 2nd Int. Symp. on Pre-failure Deformation Characteristics of
Geomaterials, Torino, Italy, Vol.1, 229–238.
Neuzil, C.E. 1993. Low fluid pressure within the Pierre Shale: a Transient response to erosion. Water Resour. Res.,
29(7): 2007–2020.
Nishi, K., Okamoto, T. & Esashi, Y. 1983. Strength-deformation characteristics of mudstone under some kinds of
loading conditions and its unificative interpretation. Proc. of Japanese Society of Civil Engineering, No. 338:
149–158 (In Japanese).
Nishi, K., Ishiguro, T. & Kudo, K. 1989. Dynamic properties of weathered sedimentary soft rocks. Soils &
Foundations, 29(3): 67–82.
Nishio, S. & Tamaoki, K. 1988. Measurement of shear wave velocity in dilluvial gravel samples under triaxial con-
ditions. Soils & Foundations, 28(2): 35–48.
Öberg, A.L. & Sällfors, G. 1997. Determination of shear strength parameters of unsaturated silts and sands based on
the water retention curve. Geotechnical Testing J., 20(1):40–48.
Ochiai H. & Lade P.V. 1983. Three-dimensional behaviour of sand with anisotropic fabric. J. of Geotechnical Eng.,
ASCE, 109(GT10): 1313–1328.
Ohtsuki, H., Nishi, K., Okamoto, T. & Tanaka, S. 1981. Time dependent characteristics of strength and deformation
of a mudstone. Proc. Symp. Weak Rock, Tokyo, 1: 119–124.
Olivares, L., Picarelli, L. & Urciuoli, G. 1993. Yield and shear strength of intensely fissured clay shales. Proc. Int.
Symp. on The Geotechnics of Hard Soils-Soft Rocks, Athens, Vol. 3.
Oloo, S.Y. & Fredlund, D.G. 1996. A method for determination of b for statically compacted soils. Canadian
Geotechnical J., 33(2): 272–280.
Olson, R.E. 1985. State-of-the-art: consolidation testing. Proc. ASTM Symposium on Consolidation of Soils. Fort
Lauderdale. American Society for Testing and Materials, Special Technical Publication 892, 99. 7–68.
O’Neill, D. A. 1985. Undrained strength anisotropy of an overconsolidated thixotropic clay. M.Sc. Thesis, Dept. of
Civil Eng., Massachusetts Institute of Technology, Cambridge, USA, 359 pp.
Ovando-Shelley, E. 1986. Stress-strain behaviour of granular soils tested in the triaxial cell. Ph.D. Thesis, Imperial
College of Science, Technology and Medicine, University of London, UK.
Park, C-S. 1993. Deformation and strength characteristics of a variety of sands by plane strain compression tests.
Dr of Eng. Thesis, University of Tokyo, Japan (In Japanese).
Park, C-S & Tatsuoka F. 1994. Anisotropic strength and deformation in sands in plane strain compression. Proc. 13th
Int. Conf. on Soil Mech. and Found. Engrg, New Delhi, India, Vol. 1, 1–4.
Pellegrino, A. 1970. Mechanical behaviour of soft rock under high stresses. Proc. 2nd Int. Conf. on Rock Mechanics,
Beograd, 2: 173–180.
Pennington, D.S., Nash, D.F.T. & Lings, M.L. 1997. Anisotropy of Go shear stiffness in Gault clay. Proc. Symp. on
Pre-failure Deformation Behaviour of Geomaterials, London, pp. 5–12.
Perret, D. 1995. Diagénèse mécanique précoce des sédiments fins du Fjord Saguenay. Ph.D. Thesis, Université
Laval, Québec.
Petley, D.J. 1966. The shear strength of soils at large strains. Ph.D. Thesis, University of London, London, U.K.
Picarelli, L. 1991a. Resistenza e meccanismi di rottura nei terreni naturali. (Personal communication)
Picarelli, L. 1991b. Discussion on the paper “The general and congruent effects of structure in natural soils and weak
rocks” by Leroueil and Vaughan. Géotechnique, 41(2): 281–284.
Picarelli, L. 1993. Structure and properties of clay shales involved in earthflows. Proc. Int. Symp. on The
Geotechnics of Hard Soils – Soft Rocks, Athens, Vol. 3: 2009–2019.
Picarelli, L., Di Maio, C., Olivares, L. & Urciuoli, G. 2000. Properties and behaviour of tectonized clay shales in
Italy. Proc. 2nd Int. Symp. on the Geotechnics of Hard Soils – Soft Rocks, Naples, Vol. 3: 1211–1241.
Picarelli, L., Olivares, L., Di Maio, C., Silvestri, F., Di Nocera, S. & Urciuoli, G. 2002. Structure, properties and
mechanical behaviour of the highly plastic intensely fissured Bisaccia Clay Shale. Proc. Int. Workshop on
Characterisation and Engineering Properties of Natural Soils, Singapore.
Pintado, X. 1993. Estudi experimental de la relacio entre el modul de tall i la succio en un sol compactat. Tesina de
Especialidad ETS Ing. Caminos, Canales y Puertos de Barcelona, UPC.
Plum, R.L. & Esrig, M.I. 1969. Some temperature effects on soil compressibility and pore water pressure. Effects of
Temperature and Heat on Engineering Behavior of Soils, Highway Research Board Special Report 103: 231–242.
Poulos, S.J. 1981. The steady state of deformation. J. Geotech. Engrg. Div., ASCE, 107(5): 553–562.
Powell, J. J. M. & Butcher, A. P. 2002. Characterisation of a glacial clay till at Cowden, Humberside. Proc. Int.
Workshop on Characterisation and Engineering Properties of Natural Soils, Singapore.
Proctor, R.R. 1933. Four articles on “The design and construction of rolled-earth dams,” Eng. News-Record, 111,
pp. 245–248, 286–289, 348–351, 372–376.
Qian, X., Gray, D.H. & Woods, R.D. 1993. Voids and granulometry: effects on shear modulus of unsaturated sands.
J. of Geotech. Engng., ASCE, 119(2): 295–314.

248
09031-02[2].qxd 18/Oct/02 12:13 AM Page 249

Rampello, S. 1989. Effetti del rigonfamiento sul comportamento maccanico di argille fortemente sovraconsolidate.
Doctoral Thesis, University of Rome.
Rampello, S. 1991. Some remarks on the mechanical behaviour of stiff clays: The example of Todi clay. Proc.
Workshop on Experimental Characterization and Modelling of Soils and Soft Rocks, Naples, pp. 131–186.
Rampello, S. & Silvestri, F. 1993. The stress-strain behaviour of natural and reconstituted samples of two overcon-
solidated clays. Proc. 1st Int. Symp. on the Geotechnics of Hard Soils – Soft Rocks, Athens, Vol. 1: 769–778.
Rampello, S. & Viggiani, G.M.B. 2001. Pre-failure deformation characteristics of geomaterials. Proc. Symp. on
Pre-failure Characteristics of Geomaterials, Torino, Vol. 2: 1279–1289.
Rampello, S., Calabresi, G. & Callisto, L. 2002. Characterisation and engineering properties of a stiff clay deposit.
Proc. Int. Workshop on Characterisation and Engineering Properties of Natural Soils, Singapore.
Rampino, C., Mancuso, C. & Vinale P. 1998. Isotropic/anisotropic behaviour of a compacted silty sand under con-
trolled suction tests. Proc. 2nd Int. Symp. on the Geotechnics of Hard Soils – Soft Rocks, Napoli, Vol. 2: 781–787.
Rix, G.J. & Stokoe, K.H., II 1991. Correlation of initial tangent modulus and cone penetration resistance. Proc. 1st
Int. Symp. on Calibration Chamber Testing, Postdam.
Robertson, P.K. 1990. Soil classification using the cone penetration test. Canadian Geotechnical J., 27(1); 151–158.
Robertson, P.K. & Campanella, R.G. 1983. Interpretation of cone penetration tests: sands. Canadian Geotech. J.,
20(4): 719–733.
Robertson, P.K., Sasitharan, S., Cunning, J.C. & Sego, D.C. 1995. Shear wave velocity to evaluate flow liquefaction.
J. of Geotech. Engng., ASCE, 121(3): 262–273.
Robertson, P. K. & Ahmadi, M. M. 2002. Characterization of Syncrude Sand with special emphasis on potential for
flow liquefaction. Proc. Int. Workshop on Characterisation and Engineering Properties of Natural Soils,
Singapore.
Rojas, E. 2002. Modeling the soil-water characteristic curve during wetting and drying cycles. Proc. 3rd Int. Conf.
on Unsaturated Soils, Recife, Vol. 1: 215–219.
Romero, E., Barrera, M., Lloret, A. & Gens, A. 2002. Collapse under isotropic stress state of anisotropic and
isotropic compacted soils. Proc. 3rd Int. Conf. on Unsaturated Soils, Recife, Vol. 2: 635–640.
Romo, M.P. 1995. Clay behaviour, ground response and soil-structure interaction studies in Mexico City. Proc. 3rd
Int. Conf. on Recent Advances in Geotechnical Earthquake Engineering and Soil Dynamics, St-Louis, Vol. 2:
1039–1051.
Romo, M.P. & Ovando, E. 1994. Comportamiento dinàmico y estàtico de los suelos del sitio Proyecto Alameda.
Informe del Instituto de Ingenieria elaborado para Reichmann International, junio (ref. by Romo, 1995).
Roscoe, K.H., Schofield, A.N. & Wroth, C.P. 1958. On the yielding of soils. Géotechnique, 8(1): 22–53.
Roscoe, K.H. & Burland, J.B. 1968. On the generalized stress-strain behaviour of ‘wet’ clay. Proc. Symp. on
Plasticity, Cambridge: 535–610.
Rowe, P.W. 1962. The stress-dilatancy relation for static equilibrium of an assembly of particles in contact. Proc.
Royal Soc., Vol. 269A, 500–527.
Ruddy, I., Andersen, M.A., Pattillo, P.D., Bishlawi, M. & Foged, N. 1989. Rock compressibility, compaction, and
subsidence in a high-porosity chalk reservoir: a case study of Valhall Field. J. of Petroleum Tech., 41(7): 741–746.
Saada A.S. & Ou C.D. 1973. Stress-strain relations and failure of anisotropic clays. J. Soil Mech. and Found. Engng.,
ASCE, 99(SM12): 1091–1111.
Saihi F., Leroueil, S., La Rochelle, P. & French, I. 2002. Influence of microstructure on a Canadian stiff and sensi-
tive clay: The Saint-Jean-Vianney clay. Canadian Geotechnical J., In print.
Sällfors, G. 1975. Preconsolidation pressure of soft high plastic clays. Ph.D. Thesis, Chalmers University of
Technology, Gothenburg, Sweden.
Samson, L., Leroueil, S., Morin, P. & Le Bihan, J.P. 1981. Pressions de préconsolidation des argiles sensibles. DSS
Contract 1SX79-00026. Division of Building Research, National Research Council of Canada.
Sandroni, S.S. 1981. Solos residuals pesquisas realizades na PCC-RJ. Proc. Brazilian Symp. on Eng. of Tropical
Soils, Rio de Janeiro, 2: 30–65.
Sandven, R. 2002. Geotechnical properties of a natural silt deposit obtained from field and laboratory tests. Proc.
Int. Workshop on Characterisation and Engineering Properties of Natural Soils, Singapore.
Santos, O.F., Lacerda, W.A. & Ehrlich, M. 1996. Discussion of “Collapse of saturated soil due to reduction in con-
finement” by Anderson & Riemer (1995). J. of Geotech. Engng, ASCE, 122(6): 505–506.
Santos, O.F., Lacerda, W.A. & Ehrlich, M. 1997. Triaxial tests with cyclic pore pressure variation for simulation of
the seasonal variation of water level in slopes. Int. Symp. on Recent Developments in Soil and Pavement
Mechanics, Rio de Janeiro, pp. 279–282.
Santucci de Magistris, F., Koseki, J., Amaya, M., Hamaya, S., Sato, T. & Tatsuoka, F. 1999. A triaxial testing system
to evaluate stress-strain behaviour of soils for wide range of strain and strain rate. Geotechnical Testing J., 22(1):
44–60.
Sasitharan, S., Robertson, P.K., Sego, D.C. & Morgenstern, N.R. 1993. Collapse behavior of sand. Canadian
Geotechnical J., 30(4): 569–577.
Saxena, S.K., Hedberg, J. & Ladd, C.C. 1978. Geotechnical properties of Hachensack valley varved clays of New
Jersey. Geotechnical Testing J., 1(3): 148–161.
Schmertmann J.H. 1969. Swell sensitivity. Géotechnique, 19(4): 530–533.
Schmertmann, J.H. 1983. A simple question about consolidation. J. Geotech. Engrg. Div., ASCE, 109(1), 119–122.
Schmertmann, J.H. 1991. The mechanical aging of soils. J. of Geotech. Engrg., ASCE, 117(9): 1288–1330.

249
09031-02[2].qxd 18/Oct/02 12:13 AM Page 250

Schofield, A.N. & Wroth, C.P. 1968. Critical State Soil Mechanics. McGraw Hill, London.
Seah, T.H. 1990. Anisotropy of resedimented Boston Blue clay. Ph.D. Thesis, Dept. of Civil Eng., Massachusetts
Institute of Technology, Cambridge, USA, 1063 pp.
Sekiguchi, H. 1977. Rheological characteristics of clays. Proc. 9th Int. Conf. on Soil Mech. and Found. Engng.,
Tokyo, 1, 289–292.
Sharma, R.S. 1998. Mechanical behaviour of unsaturated highly espansive clays. Ph.D. Thesis, University of
Glasgow, U.K.
Sheahan, T.C. 1995. Interpretation of undrained creep tests in terms of effective stresses. Canadian Geotechnical J.,
32(2), 373–379.
Sheahan, T.C., Ladd, C.C. & Germaine, J.T. 1994. Time-dependent triaxial relaxation behavior of a resedimented
clay. Geotechnical Testing J., 17(4): 444–452.
Sheahan, T.C., Ladd, C.C. & Germaine, J.T. 1996. Rate dependent undrained behavior of saturated clay. J. Geotech.
Engng., ASCE, 122(2): 99–108.
Shen, C.K, Arulanandan, K. & Smith, W.S. 1973. Secondary consolidation and strength of a clay. J. Soil Mech. and
Found. Engng. Div., ASCE, 95(1): 95–110.
Sherard, J.L., Dunnigan, L.P. & Talbot, J.R. 1984. Basic properties of sand and gravel filters. J. Geotechnical Engng,
110(6): 684–700.
Sherif, M.A. & Burrous, C.M. 1969. Temperature effects on the unconfined shear strength of saturated, cohesive
soil. Effects of Temperature and Heat on Engineering Behavior of Soils, Highway Research Board Special Report
103: 267–272.
Shibuya, S. 1985. Undrained behaviour of granular materials under principal stress rotation. Ph.D. Thesis, Imperial
College of Science, Technology and Medicine, University of London, UK.
Shibuya, S. 2000. Assessing structure of aged natural sedimentary clays. Soils & Foundations, 40(3): 1–16.
Shibuya, S. 2001. Quasi-elastic stiffness in the behaviour of soft clay. Doctor of Engineering Thesis, University of
Tokyo.
Shibuya, S. & Hight, D.W. 1987. A bounding surface for granular materials. Soils & Foundations, 27(4): 123–136.
Shibuya S. & Hight D.W. 1991. Liquefaction Line – a new concept for initiation of liquefaction of soil. Proc. Int.
Conf. on Geotechnical Engineering for Coastal Development. Geo-Coast ’91, Yokohama, Japan, Discussion to
Session 5.
Shibuya, S. & Tanaka, H. 1996. Estimate of elastic shear modulus in Holocene soil deposits. Soils & Foundations,
36(4): 45–55.
Shibuya, S., Mitachi, T., Fukuda, F. & Hosomi, A. 1997. Modelling of strain-rate dependent deformation of clay at
small strains. Proc. 12th Int. Conf. on Soil Mech. and Geotech, Engng., Hamburg, Vol. 1: 409–412.
Shibuya, S. & Tamrakar, S. B. 2002. Engineering properties of Bangkok clay. Proc. Int. Workshop on Characterisation
and Engineering Properties of Natural Soils, Singapore.
Shimizu, 1999. Geotechnical features of volcanic-ash soils in Japan. Proc. Int. Symp. on Problematic Soils,
IS-Tohoku ’98, Sendai, pp. 907–927.
Shiwakoti, D.R., Tanaka, H., Tanaka, M. & Locat, J. 2002. Influence of diatom microfossils on engineering proper-
ties of soils. Soils & Foundations, 42(3): 1–17.
Sillers, W.S., Fredlund, D.G. & Zakerzadeh, N. 2001. Mathematical attributes of some soil-water characteristic curve
models. Geotechnical and Geological Engineering J., 19: 243–283.
Simons, N.E. 1975. General report ‘Normally consolidated and highly over-consolidated cohesive materials.’ Proc.
Conf. of the British Geotechnical Society on Settlements of Structures, Cambridge, pp. 500–530.
Singh, A.W. & Mitchell, J.K. 1968. General stress-strain-time function for soils. J. Soil Mech. and Found. Engng.
Div., ASCE, 94(1): 21–46.
Skempton, A.W. 1944. Notes on the compressibility of clays. Q. J. Geol. Soc. London, 100: 119–135.
Skempton, A.W. 1977. Slope stability of cuttings in Brown London clay. Proc. 9th Int. Conf. on Soil Mech.and
Found. Engng., Tokyo, Vol. 3: 261–270.
Skempton, A.W. 1985. Residual strength of clays in landslides, folded strata and the laboratory. Géotechnique, 35(1):
3–18.
Skempton, A.W. & Petley, D.J. 1967. The strength along structural discontinuities in stiff clays. Proc. Geotechnical
Conf., Oslo, Vol. 2: 55–69.
Sladen, J.A., D’Hollander, R.D. & Krahn, J. 1985. The liquefactiuon of sands, a collapse surface approach. Canadian
Geotechnical J., 22(4): 564–578.
Smith, R.E. & Wahls, H.E. 1969. Consolidation under constant rates of strain. J. of Soil Mech. and Found. Eng. Div.,
ASCE, 95 (SM2): 519–539.
Smith, P.R., Jardine, R.J. & Hight, D.W. 1992. On the yielding of Bothkennar clay. Géotechnique, 42(2):
257–274.
Soccodato, F.M. 2002. Geotechnical properties of Fucino clayey soil. Proc. Int. Workshop on Characterisation and
Engineering Properties of Natural Soils, Singapore.
Soga, K. & Mitchell, J.K. 1996. Rate-dependent deformation of structured natural clays. Session on Measuring and
Modeling Time Dependent Soil Behavior, ASCE Convention, Washington, Geotech. Special Publication 61:
243–257.
Sridharan, A. 1991. Engineering behaviour of fine grained soils – A fundamental approach. Indian Geotechnical J.,
21(1): 1–136.

250
09031-02[2].qxd 18/Oct/02 12:13 AM Page 251

Sridharan, A. 2001. Engineering behaviour of clays: Influence of mineralogy. Proc. Workshop on Chemo-Mechanical
Coupling in Clays: From Nano-Scale to Engineering Applications, Maratea, Italy, pp. 3–28.
Sridharan, A. & Sreepada Rao, A. 1981. Rectangular hyperbola fitting method for one dimensional consolidation.
Geotechnical Testing Journal, 4(4): 161–168.
Stark, T.D. & Eid, H.T. 1994. Drained residual strength of cohesive soils. J. of Geotech. Engng, ASCE, 120 (5):
856–871.
Stokoe, K.H. II & Wood, R. 1972. In situ shear wave velocity by cross-hole method. J. of Soil Mech. and Found. Div,
ASCE., 98(5): 443–460.
Stokoe, K.H., II & Nazarian, S. 1985. Use of Raleigh waves in liquefaction studies. Proc. Measurement and Use of
Shear Wave Velocity for Evaluating Dynamic Soil Properties. ASCE, pp. 1–17.
Stokoe, K.H. II, Lee, J.N.K. & Lee, S.H.H. 1991. Characterization of soil in calibration chambers with seismic
waves. Proc. 1st Int. Symp. on Calibration Chamber Testing, Postdam.
Stokoe, K.H. II, Hwang, S.K. & Lee, K.J.N. 1994. Effects of various parameters on the stiffness and damping of soils
at small to medium strains. Proc. 1st Int. Conf. on Pre-failure Deformation Characteristics of Geomaterials,
Sapporo, Vol. 2: 785–816.
Stokoe, K.H. II, Hwang, S.K., Lee, J.N.K & Andrus, R.D. 1995. Effects of various parameters on the stiffness and
damping of soils at small to medium strains. Proc. 1st Int. Conf. on Prefailure Deformation Characteristics of
Geomaterials, Sapporo, Vol. 2: 785–816.
Stokoe, K.H. II & Santamarina, J.C. 2000. Seissmic-wave-based testing in geotechnical engineering. Proc. Int. Conf.
on Geotechnical and Geological Engineering, Melbourne, Vol. 1: 1490–1536.
Suklje, L. 1957. The analysis of the consolidation process by the isotache method. Proc. 4th Int. Conf. on Soil Mech.
and Found. Engng., London, Vol. 1, pp. 200–206.
Symes M.J. 1983. Rotation of principal stress in sand. Ph.D. Thesis, Imperial College of Science, Technology and
Medicine, University of London, U.K.
Tan, T.S., Phoon, K.K., Lee, F.H., Tanaka, H., Locat, J. & Chong, P.T. 2002. A characterisation of Singapore Lower
Marine Clay. Proc. Int. Workshop on Characterisation and Engineering Properties of Natural Soils, Singapore.
Tanaka, H. 2000. Re-examination of established relations between index properties and soil parameters. Proc. Int.
Symp. on Coastal Geotech. Engng., Yokohama, in print.
Tanaka, H. 2002. A comparative study on geotechnical characteristics of marine soil deposits worlwide. Proc. Int. J.
of Offshore and Polar Engng. 12(2): in print.
Tanaka, H., Tanaka, M. & Iguchi, H. 1994. Shear modulus of soft clay measured by various kinds of tests. Proc.
Symp. on Pre-Failure Deformation of geomaterials, Sapporo, Vol. 1: 235–240.
Tanaka, H., Ritoh, F. & Omukai, N. 2002. Geotechnical properties of clay deposits of the Osaka Basin. Proc. Int.
Workshop on Characterisation and Engineering Properties of Natural Soils, Singapore.
Tatsuoka, F. 2000. Impacts on geotechnical engineering of several recent findings from laboratory stress-strain tests
on geomaterials. The 2000 Burmister Lecture, Columbia University, U.S.A.
Tatsuoka, F. & Shibuya, S. 1991. Deformation characteristics of soils and rocks from field and laboratory tests. Proc.
9th Asian Regional Conf. on Soil Mech. and Found. Engng., Bangkok, Vol. 2: 101–170.
Tatsuoka, F. & Shibuya, S. 1992. Deformation characteristics of soils and rocks from field and laboratory tests.
Report, Institute of Industrial Science – University of Tokyo, 37(1), Serial No. 235.
Tatsuoka, F. & Kohata, Y. 1995. Stiffness of hard soils and soft rocks in engineering applications. Proc. Int. Symp. on
Pre-Failure Deformation of Geomaterials, Sapporo, Vol. 2: 947–1063.
Tatsuoka, F., Lo Presti, D.C.F. & Kohata, Y. 1995. Deformation characteristics of soils and soft rocks under monot-
onic and cyclic loads and their relationships. Proc. 3rd Int. Conf. on Recent Advances in Geomechanical
Earthquake Engineering and Soil Dynamics, St-Louis, Vol. 2: 851–879.
Tatsuoka, F., Jardine, R.J., Lo Presti, D.C.F., Di Benedetto, H. & Kodaka, T. 1997. Characterising the pre-failure
deformation properties of geomaterials. Theme Lecture for Plenary Session No. 1. Proc. 14th Int. Conf. on Soil
Mech. and Found. Engng., Hamburg, Vol. 4: 2129–2164.
Tatsuoka, F., Santucci de Magistris, F., Hayano, K., Momoya, Y. & Koseki, J. 2000. Some new aspects of time effects
on the stress-strain behaviour of stiff geomaterials. Proc. 2nd Int. Symp. on the Geotechnics of Hard Soils – Soft
Rocks, Naples, Vol. 3: 1285–1371.
Tatsuoka, F., Uchimura, T., Hayano, K., Di Benedetto, H. Koseki, J. & Siddiquee, M.S.A. 2001. Time dependent
deformation characteristics of stiff geomaterials in engineering practice. Proc. 2nd Int. Conf. on Pre-Failure
Deformation Characteristics of Geomaterials, Torino, Vol. 2: 1161–1250.
Tatsuoka, F., Hayano, K. & Koseki, J. 2002. Strength and deformation characteristics of sedimentary soft rock in the
Tokyo metropolitan area. Proc. Int. Workshop on Characterisation and Engineering Properties of Natural Soils,
Singapore.
Tavenas, F. & Leroueil, S. 1977. Effects of stresses and time on yielding of clays. Proc. 9th Int. Conf. on Soil Mech.
and Found. Engng., Tokyo, Vol. 1, pp. 319–326.
Tavenas, F., Leroueil, S., La Rochelle, P. & Roy, M. 1978. Creep behaviour of an undisturbed lightly overconsoli-
dated clay. Canadian Geotechnical J., 15(3): 402–423.
Tavenas, F., Brucy, M., Magnan, J.P., La Rochelle, P. & Roy, M. 1979. Analyse critique de la théorie de consolidation
unidimensionnelle de Terzaghi. Revue Française de Géotechnique, 7: 29–43.
Tavenas, F., Jean, P., Leblond, P. & Leroueil, S. 1983a. The permeability of natural soft clays. Part II: Permeability
characteristics. Canadian Geotechnical J., 20(4): 645–660.

251
09031-02[2].qxd 18/Oct/02 12:13 AM Page 252

Tavenas, F., Leblond, P., Jean, P. & Leroueil, S. 1983b. The permeability of natural soft clays. Part I: Methods of
laboratory measurement. Canadian Geotechnical J., 20(4): 629–644.
Tavenas, F., Tremblay, M., Larouche, G. & Leroueil, S. 1986. In situ measurement of permeability in soft clays. Proc.
ASCE Specialty Conference In situ’86, Blacksburg, pp. 1034–1048.
Tavenas, F. & Leroueil, S. 1987. State-of-the-Art on “Laboratory and in situ stress-strain-time behavior of soft
clays”. Proc. Int. Symp. on Geotechnical Engng. of Soft Soils, Mexico City, Vol. 2: 1–46.
Taylor, R.K. & Cripps, J.C. 1987. Weathering effects: slopes in mudrocks and overconsolidated clays. Slope Stability,
Wiley, New York, Chapter 13, pp. 405–445.
Teh, C.I. & Houlsby, G.T. 1991. An analytical study of the cone penetration test in clay. Géotechnique, 41(1): 17–34.
Terzaghi, K. 1923. Die Berechnung der Durchlässigkeitsziffer des Tones aus dem Verlauf der Hydrodynamishchen
Spannungsercheinungen. Sitz, Akademie der Wissenschaften, Mathematish-naturwissenschafttliche, Klasse,
Vienna, Austria, Part Iia, 132, pp. 125–138.
Terzaghi, K. 1936. The shearing resistance of saturated soils and the angles between the planes of shear. Proc. 1st Int.
Conf. on Soil Mech., Cambridge, Vol. 1: 54–56.
Terzaghi, K. & Peck, R.B. 1948. Soil Mechanics in Engineering Practice. John Wiley & Sons, Inc., New York.
Terzaghi, K., Peck, R.B. & Mesri, G. 1996. Soil Mechanics in Engineering Practice, Third Edition John Wiley &
Sons, Inc., New York.
Tidfors, M. & Sällfors, G. 1989. Temperature effect on the preconsolidation pressure. Geotechnical Testing J., 12(1):
93–97.
Tiedemann, B. 1937. Über die Schubfestigkeit bindiger Böden. Die Bautechnik, 15: 400–403; 433–435.
Tien, Y.M., Lee, D.H. & Juang, C.H. 1990. Strain, pore pressure and fatigue characteristics of sandstone under vari-
ous load conditions. Int. J. of Rock Mech. and Mining Sciences, 27(4):283–289.
Tika, T.E., Vaughan, P.R. & Lemos, L.J. 1996. Fast shearing of pre-existing shear zones in soil. Géotechnique, 46(2):
197–233.
Tillard-Ngan, D., Desrues, J., Raynaud, S. & Mazerolle, F. 1993. Strain localization in Beaucaire marl. Proc. Symp.
on Geotechnical Engineering of Hard Soils – Soft Rocks, Athens, 2: 1679–1686.
Torstensson, B.A. 1984. A new system for groung water monitoring. Ground Water Monitoring Review, 4(4):
131–138.
Torstensson, B.A. & Petsonk, A.M. 1986. A device for in situ measurement of hydraulic conductivity. Proc. 4th
Int. Geotechnical seminar on Field Instrumentation and In Situ measurements, Singapore, pp. 157–162.
Urciuoli, G. 1992. Rigonfiamento di un’argilla di alta plasticita e modellazione dei fenomeni erosivi del colle di
Bisaccia. Ph.D. Thesis, Universita di Napoli ‘Frederico II’, Italy.
Uriel, S. & Serrano, A.A. 1973. Geotechnical properties of two collapsible volcanic soils of low bulk density at the
site of two dams in Canary Islands (Spain). Proc. 8th Int. Conf. on Soil Mech. and Found. Engrg., Moscow, 2(2):
257–264.
Uthayakumar, M. & Vaid, Y.P. 1998. Static liquefaction of sands under multiaxial loading. Canadian Geotech. J.,
35(2): 273–283
Vaid, Y.P. & Campanella, R.G. 1977. Time-dependent behaviour of undisturbed clay. J. Geotech. Engrg. Div., ASCE,
103(7): 693–709.
Vaid, Y.P., Robertson, P.K. & Campanella, R.G. 1979. Strain rate behaviour of Saint-Jean-Vianney clay. Canadian
Geotechnical J., 16(1): 34–42.
Vaid, Y.P. & Chern, J.C. 1985. Cyclic and monotonic undrained response of sands. Proc. Conf. Advances in the Art
of Testing Soils Under Cyclic Loading Conditions, Detroit, pp. 120–147.
Vaid, Y.P., Chung, E.K.F. & Kuerbis, R.H.V. 1990. Stress path and steady state. Canadian Geotechnical J., 27(1):
1–7.
Vaid, Y.P., Uthayakumar, M., Sivathayalan, S., Robertson, P.K. & Hofmann, B. 1995. Laboratory testing of Syncrude
sand. Proc. 48th Canadian Geotech. Conf., Vancouver, Vol. 1: 223–232.
Vaid, Y.P. & Eliadorani, A. 1998. Instability and liquefaction of granular soils under undrained and partially drained
states. Canadian Geotechnical J., 35(6): 1053–1062.
Vaid, Y.P. & Sivathayalan, S. 2000. Fundamental factors affecting liquefaction susceptibility of sands. Canadian
Geotechnical J., 37(3): 592–606.
Vanapalli, S.K. 1994. Simple test procedures and their interpretation in evaluating the shear strength of an unsatu-
rated soil. Ph.D Thesis, University of Saskatchewan, Canada.
Vanapalli, S.K., Fredlund, D.G., Pufahl, D.E. & Clifton, A.W. 1996. Model for the prediction of shear strength with
respect to soil suction. Canadian Geotechnical J., 33(3): 379–392.
Vanapalli, S.K., Fredlund, D.G. & Pufahl, D.E. 1999. The influence of soil structure and stress history on the soil-
water characteristics of a compacted till. Géotechnique, 49(2): 143–159.
Vaughan, P.R. 1997. Engineering behaviour of weak rock: Some answers and some questions. Proc. 1st Int. Conf. on
Hard Soils and Soft Rocks, Athens, Vol. 3: 1741–1765.
Vaughan, P.R. & Hamza, M.M. 1977. Clay embankments and foundations: Monitoring stability by measuring defor-
mations. Specialty Session 8: Deformation of earth-rockfill dams. Proc. 9th Int. Conf. on Soil Mech. and Found.
Engng, Tokyo, pp. 37–48.
Vesic, A.S. & Clough, G.W. 1968. Behaviour of granular materials under high stresses. J. Soil Mech. and Found.
Div., ASCE, 94(3): 661–688.

252
09031-02[2].qxd 18/Oct/02 12:13 AM Page 253

Viana da Fonseca, A. 2002. Characterising and deriving engineering properties of a saprolitic soil from granite, in
Porto. Proc. Int. Workshop on Characterisation and Engineering Properties of Natural Soils, Singapore.
Viggiani, G., Rampello, S. & Georgiannou, V.N. 1993. Experimental analysis of localization phenomena in triaxial
tests on stiff clays. Proc. Symp. on Geotech. Engng of Hard Soils – Soft Rocks, Athens, 1: 849–856.
Viggiani, G. & Atkinson, J.H. 1995a. Stiffness of fine grained soils at very small strains. Géotechnique, 45(2):
249–265.
Viggiani, G. & Atkinson, J.H. 1995b. The interpretation of bender element tests. Géotechnique, 45(1): 149–155.
Vinale, F., d’Onofrio, A., Mancuso, C. & Santucci de Magistris, F. 2001. The pre-failure behaviour of soils as con-
struction materials. Proc. 2nd Int. Symp. on Pre-Failure Deformation Characteristics of Geomaterials, Torino,
Vol. 2: 955–1007.
Vucetic, M. & Dobry, R. 1991. Effect of soil plasticity on cyclic response. J. of Geotech. Engng. Div., ASCE, 117(1):
89–107.
Vulliet, L. 1986. Modélisation des pentes naturelles en mouvement. Thesis No. 635, École Polytechnique Fédérale
de Lausanne, Switzerland.
Ward, W.H., Marsland, A. & Samuels, S.G. 1965. Studies of the properties of undisturbed London Clay at the
Ashford Common shaft: in-situ and immediate strength tests. Géotechnique, 15(4): 321–344.
Watabe, Y., Leroueil, S. & Le Bihan, J.P. 2000. Influence of compaction conditions on pore-size distribution and
saturated hydraulic conductivity of a glacial till. Canadian Geotechnical J., 37(6): 1184–1194.
Watabe, Y., Tanaka, M., Tanaka, H. & Tsuchida, T. 2002. Ko-consolidation in a triaxial cell and evaluation of in-situ
Ko for marine clays with various characteristics. Soils & Foundations, In press.
Wesley, L.D. 1977. Shear strength properties of halloysite and allophane clays in Java, Indonesia. Géotechnique,
27(2): 125–136.
Wesley, L. D. 2002. Geotechnical characterisation and behaviour of allophone clays. Proc. Int. Workshop on
Characterisation and Engineering Properties of Natural Soils, Singapore.
Weston, T.R. 1996. Effects of grain size and particle distribution on the stiffness and damping of granular soils at
small strains. M.Sc. Thesis, The University of Texas at Austin (ref. by Jamiolkowski and Lo Presti, 1998).
Wheeler, S.J. (2002). Keynote Lecture presentation. Proc. 3rd Int. Conf. on Unsaturated Soils, Recife.
Wheeler, S.J & Sivakumar, V. 1995. An elasto-plastic critical state framework for unsaturated soil. Géotechnique,
45(1): 35–53.
Wheeler, S.J. & Karube, D. 1996. Constitutive modelling. Proc. 1st Int. Conf. on Unsaturated Soils, Paris, 3:
1323–1356.
Wong, P.K.K. & Mitchell, R.J. 1975. Yielding and plastic flow of sensitive cemented clay. Géotechnique, 25(4):
763–782.
Wood, D.M. 1990. Soil Behaviour and Critical State Soil Mechanics. Cambridge University Press, Cambridge.
Wroth, C.P. & Loudon, P.A. 1967. The correlation of strains within a family of triaxial tests on overconsolidated
samples of kaolin. Proc. of the Geotechnical Conf. on the Shear Strength Properties of Natural Soils and Rocks,
Oslo, pp. 163–189.
Wroth, C.P. & Wood, D.M. 1978. The correlation of index properties of soils. Canadian Geotechnical J., 15(2):
137–145.
Wroth, C.P. 1984. The interpretation of in situ soil test. 24th Rankine Lecture, Géotechnique, 34(4): 449–489.
Wroth, C.P. 1988. Penetration testing – a more rigorous approach to interpretation. Proc. Int. Symp. on Penetration
testing, ISOPT-1, Orlando, Vol. 1: 303–311.
Wroth, C.P. & Hughes, J.M.O. 1973. An instrument for the in situ measurement of the properties of soft clays. Proc.
8th Int. Conf. on Soil Mech. and Found. Engng, Moscow, Vol. 1.2, pp. 487–494.
Wu, S., Gray, D.H. & Richart, F.E. 1984. Capillary effects on dynamic modulus of sands and silts. J. of Geotech.
Engng., ASCE, 110 (9): 1188–1203.
Yoshida, T. & Tatsuoka, F. 1997. Deformation property of shear band in sand subjected to plane strain compression
and its relation to particle characteristics. Proc. 14th Int. Conf. on Soil Mech. and Found. Eng., Hamburg, Vol. 1:
237–240.
Yoshikuni, H., Hirao, T., Nishiumi, H. & Ikegami, S. 1993. The behavior of swelling and recompression due to
unloading. Proc. 48th Annual Conf. of the Japan Society of Civil Engineers, 48(3), 1010–1011. (In Japanese, ref.
by Imai, 1995)
Yoshikuni, H., Nishiumi, H., Ikegami, S. & Seto, K. 1994. The creep and effective stress-relaxation behavior on one-
dimensional consolidation. (in Japanese). Proc. 29th Japan National Conf. on Soil Mech. and Found. Engrg., 29:
269–270.
Yoshikuni, H., Kusakabe, O., Okada, M., & Tajima, S. 1995. Mechanism of one-dimensional consolidation.
Proc. Int. Symp. on Compression and Consolidation of Clayey Soils – IS-Hiroshima’s 95, Hiroshima, 1(4):
97–504.
Yoshimine, M. & Ishihara, K. 1998. Flow potential of sand during liquefaction. Soils & Foundations, 38(3): 189–198.
Yoshimine, M., Robertson, P.K. & (Fear) Wride, C.E. 1999. Undrained shear strength of clean sands to trigger flow
liquefaction. Canadian Geotechnical J., 36(5): 891–906.
Yoshinaka, R. & Osada, M. 1995. The comparison between dynamic and static strength of soft sedimentary rocks. In
Rock Foundation. Edited by R. Yoshinaka and K. Kikuchi. Balkema, Rotterdam, pp. 109–114.
Yoshinaka, R. & Osada, M. 1998. Personal communication.

253
09031-02[2].qxd 18/Oct/02 12:13 AM Page 254

Zakaria, I., Wheeler, S.J & Anderson, W.F. 1995. Yielding of unsaturated compacted kaolin. Proc. 1st Int. Conf. on
Unsaturated Soils, Paris, 1: 223–228.
Zavoral, D.Z. & Campanella, R.G. 1994. Frequency effects on damping/modulus of cohesive soil. Proc. Conf. on
Dynamic Geotechnical Testing, ASTM, STP 1213, 191–201.
Zdravkovic L. 1996. The stress-strain-strength anisotropy of a granular medium under general stress conditions.
PhD Thesis, Imperial College of Technology, Science and Medicine, University of London, UK
Zdravkovic, L. & Jardine, R.J. 1997. Some anisotropic stiffness characteristics of a silt under general stress condi-
tions. Proc. Symp. on Pre-failure Deformation Behaviour of Geomaterials, London, pp. 21–51.
Zdravkovic L. & Jardine R.J. 2000. Undrained anisotropy of a K0 consolidated silt. Canadian Geotech. J., 37(1):
178–200.

254

S-ar putea să vă placă și