You are on page 1of 10

COMPRESSIBILITY OF CLAYS: FUNDAMENTAL AND

PRACTICAL ASPECTS·

By S. Leroueil1

ABSTRACT: The compressibility of natural clays is influenced by numerous factors: strain rate, temperature,
sampling disturbance, stress path, and some restructuring factors. The first part of the paper reviews the effects
of these factors, in particular of strain rate and temperature. The influence of drainage conditions on the effective
stress-strain curves followed in various sUbelements of a consolidating clay layer is also discussed. In a second
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DE PARANA on 08/22/13. Copyright ASCE. For personal use only; all rights reserved.

part, in-situ conditions are considered. In the overconsolidated range, and at the preconsolidation pressure, the
behavior is influenced by most of the aforementioned factors, and can be compared with laboratory test results
only on the basis of a serniempirical approach. In the normally consolidated range, major factors are strain rate
and temperature, and their effect can be evaluated. In some cases, however, structuring phenomena can exist
and decrease the viscous effects. Finally, practical conclusions concerning the evaluation of long-term settlements
are given.

INTRODUCTION ship. This rheologic model, originally proposed by Suklje


(1957), was recently confirmed by Imai and Tang (1992) for
Compressibility of clays has always been a topic for lively the Yokohama Bay mud. There last authors showed in partic-
discussions among soil engineers. Those involving Bjerrum ular that the model remains valid when a clay specimen is
(1967, 1972) and Leonards (Leonards and Altschaeffl 1964; reloaded during primary consolidation. Leroueil et al. (1985)
Leonards 1972) are particularly famous. In fact, both were
also showed that the (a~, £v, tv) relationship can be simply
right, but they were not talking about or referring to the same described by two curves, one giving the variation of the pre-
aspect of clay behavior. Leonards (1972) was mostly referring consolidation pressure a; with strain-rate [a; = j(t v )] and the
to his work on very recent materials in which "structuring other presenting the normalized stress-strain curve [a ~/a ;(tv )
effects" are important, whereas Bjerrum (1967, 1972) was re- = g(£v)]. Figs. l(a) and (b) show a~/a; = j(t v) curves for
ferring to clays which were formed thousands of years ago Champlain Sea clays and Finnish clays, respectively. They are
and in which creep becomes a major factor when they are very similar, even if the scatter is more important for the or-
loaded. The materials of main practical interest fall in this ganic [organic matter (OM) > 3%] Finnish clays, and seem to
second category. The paper will be divided into two main reflect the behavior of most natural clays.
parts, the first one describing the fundamental aspects of clay These strain-rate effects have important implications, which
behavior and the second giving the views of the writer on how have been discussed by Leroueil et a1. (1985) and Leroueil
to estimate settlements of clay deposits. (1988). One of them is that constant rate of strain (CRS)
oedometer tests give stress-strain curves that are generally dif-
FUNDAMENTAL ASPECTS OF CLAY BEHAVIOR ferent from those obtained in conventional 24 h oedometer
Clay behavior in one-dimensional or nearly one-dimen- (a) 1 . 3 , r - - - - -......-------r-----r-----:'~
sional conditions is mainly influenced by strain rate (£v), tem- c"..",,* S.. c"".
perature (T), disturbance, the stress path followed, and restruc-
turing phenomena. The influence of these factors is briefly
~'" 1.1l-----+-----+---:,ttf,o;r-.:.;.-::.--=-I
described hereafter. ~
>C
~
Strain-Rate Effects
i 0.91-------7--r---,---:>lI!II!R~!Ib:.,..."---I--O------l
The influence of time and strain rate on the compressibility ~
of clays has been evidenced for decades (Taylor 1942; Craw-
ford 1965; Bjerrum 1967). Studying in detail the rheologic

behavior of natural clays tested in different types of oedometer
test, Leroueil et a1. (1985) showed that during one-dimensional
compression, the behavior is controlled by a unique effective
stress-strain-strain rate (a~, £v, tv) [or (a~, e, e)] relation- (b) 1.3 •matter
Deplh(m Ref.

"This paper was originally published in "Vertical and Horizontal De- PaImIo 5.7
PIcku-lluclpolabl 8.2 •• 0
formation of Foundations and Embankments." Under a special program, •
-
TorpparIn_ 2.1-2.7 D
the paper was nominated for potential republication, was reviewed in the 'co v_ 4.3-4.8 0
" 1.1 Ota_ 2.0-2.4 D
same manner as all other journal papers, and was accepted for republi-
cation. This paper differs from the original paper to incorporate sugges-
~ 2.0 - 2.8
8.0-8.3 ••
tions of the reviewers and to recognize more recent developments, most ~
>C
T_ 3.0-3.4 •
of which were presented at the 1995 International Symposium on Com- i 0.9
pression and Consolidation of Clayey Soils in Hiroshima, Japan. ~
'Prof. of Civ. Engrg., Universite Laval, Sainte-Foy, Quebec, G IK 7P4,
Canada.
Note. Discussion open until December 1, 1996. To extend the closing
date one month, a written request must be filed with the ASCE Manager
of Journals. The manuscript for this paper was submitted for review and
possible publication on June 9, 1995. This paper is part of the }ouTluzl FIG. 1. Strain Rate Effect on Preconsolldatlon Pressure:
of Geotechnical Engineering, Vol. 122, No.7, July 1996. ©ASCE, ISSN (a) Canadian Experience [after Lerouell et al. (1983, 1985));
0733-9410/96/0007-0534-0543/$4.00 + $.50 per page. Paper No. (b) Finnish Experience [after Paakkunalnen (1990) and Holkkala
11135. (1991)]

534/ JOURNAL OF GEOTECHNICAL ENGINEERING / JULY 1996

J. Geotech. Engrg. 1996.122:534-543.


tests. Contrary to what is indicated in ASTM Standard D 4186 Temperature Effects
("Standard" 1993), the strain rates used in CRS tests are gen-
In relation to new concerns such as nuclear-waste isolation
erally not within the range encountered in conventional 24 h
and ~e use of soil deposits for heat energy storage, several
tests [24-h multiple stage loading (MSL24)]. They are in the
expenmental studies on the effects of temperatures on natural
order of, or less than, 10- 7 S-1 at the end of the loading periods
cl.ay behavior have recently been performed (Erickson 1989;
in the conventional MS~ tests (Leroueil 1988), while they
Tldfors and Siillfors 1989; Hueckell and Baldi 1990; Boudali
are generally between 1 X 10-6 and 4 X 10-6 S-1 in the CRS
et al. 1994). In terms of compressibility, the observed behavior
tests. A typical comparison of MS~ and CRS test results is can be illustrated by Fig. 3 and described on the basis of a
shown in Fig. 2. Other comparisons obtained on a variety of
study performed at Laval University (Boudali et al. 1994) as
natural clays are summarized in Table 1, where it can be seen follows: (1) The vertical strain generated by a change in tem-
that the preconsolidation pressure deduced from CRS tests is
perature of 30°C, under a constant effective stress in the
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DE PARANA on 08/22/13. Copyright ASCE. For personal use only; all rights reserved.

typically 25% higher than that deduced from conventional 24


overconsolidated range, remains small, in the order of 0.5%.
h tests. The same applies to the entire stress-strain curve. As
(2). In the normally consolidated range and at the preconsoli-
our practical experience is essentially based on conventional datIOn pressure, a temperature change of 30°C has a significant
24 h oedometer test results, it implies that CRS test results effect on compressibility of clays. At a given strain rate, the
have to be corrected for strain rate effects before being used smaller the temperature, the higher is the effective stress at a
in practice. given strain or void ratio. (3) As shown in Fig. 3, temperature
effects combine with the effects of strain rate to influence the
4.0 l"""T"T~_r---T---r-'l"""""""',Jr"T'T,~k--"'T--'T-- viscous nature of natural clays. For this reason, the stress-
strai~ curves obtained on the Berthierville clay at (T = 5°C
.... ~ ~// (£=3.3x10-6 s-1 ) and Ev = 1.6 X 10- 7 S-I) and at (T = 35°C and Ev = 1.6 X
3.6 H-H+t---+~\""+'-:~~I-++l+--.I--l---J-J 10- 5 S-I) coincide. (4) The preconsolidation pressure can be
described. as a function of strain rate and temperature, 0' ~ = f
(t", T) [FIg. 4(a)]. (5) It appears that all the stress-strain curves
MSL 24 ~ ~
3.2 t-t-11-ttt---+-~:!...j--3~\~I+--~--+-t
obtained at various strain rates and temperatures come into a
unique one when they are normalized with respect to the pre-
consolidation pressure corresponding to the strain rate and the
Q)
o , Mean vertical effective stress, kPa
, 2.8 t+++t+-_f---+--+-1~~\)---W~
:g o 50 100 150 200 250 300
~ O~....:-r---..,r----.-----.--~----.
T EV
2.4 H+ttt---t-+-+-H-++ffi~tH~~-t

~,
(0 C) (s -1 )
5
~ 35 l.OxlO-5

<!. ---+- 5 l.OxlO- 5


2.0 H+t++---t-+-+-+-+++++--~~~-t 10

"'a~ 'f
-;
.~
35
5
1.6xlO-7
1.6xlO-7

1.6 ~-::-:~_~~~~!-!-.&+I-_ _L...J--I.-1 15


5 6 7 6~ 0 1 2 3 4 5 6 78 ~02 2 3 4 5 -<
Effective stress cry (kPa)
20
FIG. 2. Comparison between Stres..straln Curves Obtained
In Conventional 24 hand CRS Oedometer Tests [from Hanzawa
(1989)] 25

TABLE 1. Comparison between Preconsolldatlon Pressures 40


Measured In Conventional 24 h and In CRS (1 to 4 x 10-1 S-l)
Oedometer Tests
35
Location Reference ~'"

(1 ) (3) gj 30
~
Qu6bec, dozen of clays
Sweden, several sites
1.28 Leroueil et al. (1983)
Larsson and Slillfors
(1985)
.sil 25

~
Finland, one site 1.16 Kolisoja et al. (1989)
Osaka, Japanb 1.3-1.5 Hanzawa et al. (1990) 20
Fucino, Italy Burghignoli et al. (1991) '"
Ariake and Kuwana, Japan
1.2
1.3-1.4 Hanzawa (1991) ~ 15
Yokohama, Japan 1.25 Okumura and Suzuki
(1991) i 10
Finland, three sites 1.3 Hoikkala (1991) '"'"
Japan, several clays 1.18 Mizukarni and Motoyash- ~
~
iki (1992) 5
Bothkennar, United Kingdom 1.33 Nash et al. (1992)
"Larsson and Siillfors also found a ratio larger than 1.0, but correct the 0
CRS test results with a graphical method.
bClay from the Pleistocene period. FIG. 3. 'lYplcal CRS Oedometer Test Results Obtained at Dif-
ferent Strain Rates and Temperatures [after Boudall et al. (1994)]

JOURNAL OF GEOTECHNICAL ENGINEERING 1JULY 1996/535

J. Geotech. Engrg. 1996.122:534-543.


temperature used in each test [Fig. 4(b)]. The model proposed Sampling Disturbance
by Leroueil et at. (1985) for rate effects can thus be extended
for including temperature effects. (6) On the basis of data ob- As shown for a variety of clays, sampling disturbance has
tained by several researchers on seven different clays, an idea a major consequence on compressibility parameters (Schmert-
of the variation of the preconsolidation pressure (or the vertical mann 1955; La Rochelle et at. 1981; Lacasse et at. 1985; Ler-
effective stress at a given void ratio) with temperature can be oueil and Kabbaj 1987; Hight et al. 1992; Magnan et at. 1994).
obtained, as shown in Fig. 5. It is significant since typically a; It decreases the preconsolidation pressure and, more generally,
increases the estimated settlements under any stress. Typical
decreases by 35% when the temperature increases from 5°C
to 40°C. oedometer tests performed on Vasby (Sweden) clay and shown
in Fig. 6 illustrate well the effect of sampling disturbance. The
Vertical strain rate, £1 (s'1) specimen taken with the large diameter (200 mm) Laval sam-
7 5 pler shows a sharp knee at the passage of a;
whereas the
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DE PARANA on 08/22/13. Copyright ASCE. For personal use only; all rights reserved.

10-9 10'" 16 10'" 16 10-4


100 specimen taken in 1967 with the Swedish sampler presents a
'iil90 more rounded curve. In average, the oedometer tests per-
~80
... formed on specimens taken with the Laval sampler gave pre-
b 70
consolidation pressure about 30% higher (Leroueil and Kabbaj
l! 1987).
i
Q.
60
Stress Paths
c
.250

I~40 CRS
Creep
5· 20· 35·
... • •
0
Clay behavior is known to be essentially stress-path depen-
dent. In one-dimensional or nearly one-dimensional condi-
MSL p II tions, the stress path remains close to the K one ("" 1 - simp')
Do
MSL 24 m line when the soil is normally consolidated. On the other hand,
30 it can significantly vary when the soil is overconsolidated, de-
pending in particular on the initial stress conditions. As shown
Normalized effective stress 0"1/O"p (£1, T) later [(2) and Figs. 11 and 12] this aspect may have important
0 0.5 1.0 1.5 2.0 2.5 3.0
0 consequences when estimating the in situ vertical yield stress
of overconsolidated clays.

51---+--~lr--+---I---+---1 Restructuring Phenomena


~ Leonards and Altschaeffl (1964), and Perret (1995) have
~ 101-----l---+- shown that in the early stages of the diagenesis of clay sedi-
c
'il! ments, there can be some strengthening of the soil structure
'Iii
with time that modifies the effects of secondary compression
~ 15l---+--4---+-=
and gives preconsolidation pressure higher than those associ-
:!: ated to the void ratio of the clay only.
201-----l---+--4---+----::: Perret (1995) performed one-dimensional compression tests
on the resedimented Jonquiere clay (Ip = 32%) in large-di-
ameter (200 mm) cells equipped with bender elements to fol-
25L-_---I._ _....L..._ _1......_-L_ _..1-_....J
(From Kabbaj. 1985 and Boudalle! aI., 1994) 3.0
FIG. 4. One-Dimensional Compression of Berthierville Clay ~ r-a.. I"b[\
[from Boudall et al. (1994) and Kabbaj (1985)]: (a) Preconsollda-
tion Pressure As Function of Strain Rate and Temperature; and 2.8
~ ,
(b) Normalized Effective Stress-Strain Curve

1.40 , . - - - . . , - - -.......- - - . . . . - - - - . . - - - - ,
2.6
~ 1\

1.20
• III 2.4
1\ 1\

u
~ i:2 ~I\
0

l<1 i • o
~
> 2.2
1.00 0
~

-Il
... '*"
I!I •
..... 8. 10)
-r: 181
<> •
~. 2.0

0.80 • ••
0 • 1.8 -
C Specimen taken with
the Laval sampler \
0.60
0 20 40 60
0

80
1"\

100 1.6
o Specimen taken in 1967
I1II
\
Temperature, 0 C 1 10 100
Vertical effective stress <Tv (kPa)
FIG. 5. Variation of Normalized Preconsolldation Pressure, or
Vertical Effective Stress at a Given Void Ratio, with Temperature FIG. 6. Typical Compression Curves for Vasby Clay [after Ler-
for Different Clays [after Boudali et al. (1994)] ouell and KabbaJ (1987)]

536/ JOURNAL OF GEOTECHNICAL ENGINEERING / JULY 1996

J. Geotech. Engrg. 1996.122:534-543.


low the evolution of the maximum shear modulus Go. As Go and subjected to a constant loading during the same time, but
is directly related to the preconsolidation pressure, its mea- with a higher organic matter content of 4.6% does not show
surement makes possible to follow the evolution of O'~, and any significant structuration. Organic matter thus seems to be
thus of the .structure, during consolidation. a factor influencing the structuration of clayey soils. Leroueil
Typical results are shown in Fig. 7 for a test in which the et al. (1985) and Kabbaj (1985) also observed that, in some
soil was allowed to consolidate during 120 d under 10 kPa, cases, CRS oedometer tests performed at very low strain rates,
before being loaded again. Just on the basis of the change in lower than 2 X 10- 8 S-I, give compression curves that cross
void ratio during secondary consolidation, the preconsolidation compression curves obtained at larger strain rates, indicating
pressure would have increased from 10 to 11.5 kPa [Fig. 7(a)]. that some structure develops.
However, due to structuration, the preconsolidation pressure It comes out from these observations that structuring effects
has increased to 18.5 kPa. The variation in Go under the sus- can develop during the consolidation of clayey soils, but that
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DE PARANA on 08/22/13. Copyright ASCE. For personal use only; all rights reserved.

tained load of 10 kPa is shown with black dots in Fig 7(b). If these effects seem to differ from soil to soil. It can also be
Go would only be associated to the void ratio, it would follow noted that the existence of nearly normally consolidated clay
the aa line. In fact, Go first goes on the left side of aa, in- deposits indicates that structuration can be extremely small in
dicating a destructuration just after loading, and then, pro- some cases.
gressively goes on the right side of aa as structure develops.
These results show that compression and structuration can be INFLUENCE OF DRAINAGE CONDITIONS ON
intimately linked during both primary and secondary consoli- STRESS-STRAIN BEHAVIOR
dations.
Casagrande (1932), Locat and Lefebvre (1985), and Burland Hypothesis A versus Hypothesis B
(1990) indicated that natural clays often develop some struc- In relation to the influence of the drainage length on the
ture during their diagenesis. Perret et al. (1995) also evidenced stress-strain curve followed during primary consolidation,
that structuring phenomena could also exist for natural clays there are two extreme possibilities, hypotheses A and B (Ja-
loaded in the normally consolidated range. As shown in Fig. miolkowski et al. 1985). "Hypothesis A assumes that creep
8(a), the specimen of the Saguenay Fjord with a low organic occurs only after the end of primary consolidation," and con-
matter content (1.1 %) developed a significant strengthening sequently, the same stress-strain curve [which would also be
during the 82 d under a constant load, with a preconsolidation the end-of-primary consolidation (EOP) curve] is followed
pressure 22% higher than that corresponding to its void ratio whatever the drainage conditions. The EOP approach for cal-
only. On the other hand, the specimen from the same origin culating in situ settlements proposed by Mesri and Choi
(1985a) and Mesri et al. (1994) is based on this hypothesis.
a,\
2.8
\ "Hypothesis B assumes that some sort of 'structural viscosity'
is responsible for creep, that this phenomenon occurs during
\
\ pore pressure dissipation .. " and consequently, the stress-
strain curve followed depends on drainage conditions. Detailed
~
2.• observations of clay behavior can indicate which hypothesis
\ is the most representative.
1\
l\ Laboratory Observations
120d ~ ~,
,, ,
2.0
, \
,, :\ Berre and Iversen (1972), Mesri and Feng (1986), and Imai
and Tang (1992) have put subspecimens of clay in series to
examine local stress-strain behavior within a consolidating
\~
(8)
5 8 7 8 i
10
"~2 33
(b)
4 5 e 75 ~03 2
~~
3 4 5 8
clay layer. The results obtained by Leroueil et al. (1986) and
by Imai and Tang (1992) clearly show that the stress-strain
curves followed depend on the position of the sub-specimen
0'. (kP.) Go (kP.) relatively to the drainage boundary (Fig. 9). At the beginning
(from "'rtet, 1895) of the loading period, the strain rate near the drainage bound-
ary is higher than that near the impervious boundary. In agree-
FIG. 7. Re88dlmented Jonqul.re Clay [from Perret (1995)]: <a)
Compression Curve; and (b) Variation of Maximum Shear Mod· ment with the stress-strain-strain rate model previously men-
ulus Go with Void Ratio tioned, clay element 1 near the drainage boundary mobilizes
higher effective stresses than clay element 4 near the imper-
Vertical eII8c:tIve pnI8IUIW 0'. (kPa) vious boundary. However, when the clay specimen is ap-
20 30 40 60 80 80 100 80 80 100 200 300 400 500 proaching the end of primary consolidation, there is unifor-
3.0
1.8 "- mization of the strain rate in the clay layer, and consequently,
I I II
1\
1\ 11°... ··1.1% 2.7 0 ..... ,.5%/" the stress-strain curves of the different subspecimens converge.
The same happens during the successive steps of loading [Fig.

,
1.7
'a 9(b)], which gives stress-strain curves that are different from
• \,
~
one subspecimen to the other, and also different from the con-
11.8
:g ~~ 01
ventional curve deduced from the end points only.
>1.5

~b
Field Observations
'\
1.4
1.5
~ In 1990, Hydro-Quebec (Montreal, Canada) built a test em-
1\ bankment on site Olga in the Province of Quebec, about 10
(8) (11)
1.3 km northeast of Matagami, Quebec, and 600 km northwest of
(from Pertet et aI., 1895) Montreal, to study the efficiency of vertical wick drains in the
FIG. 8. Influence of secondary Consolidation on Structura- sensitive clays from this area (Lavallee et al. 1990). This em-
tlon of Saguenay Fjord Sediments after 82 Days of Sustained bankment is 6 m high and has total width and length of 106
Load [from Perret et al. (1995)] m and 146 m, respectively. It comprises four different sections:
JOURNAL OF GEOTECHNICAL ENGINEERING / JULY 1996/537

J. Geotech. Engrg. 1996.122:534-543.


EffectlYe .tr... CT~, kPa
80 90 100 110 120 130 140
0

0,5 ~~~
c..c,
"\J.
.,~
l'....

"" ~\\
.~

.)

W
\

'~l,.\.~~
\
\
~l~~ 10'
LDgt (I Ind)
10
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DE PARANA on 08/22/13. Copyright ASCE. For personal use only; all rights reserved.

~~
<Iv (kPa) <Iv (kPa)
25 50 7 1'" 1 0 25 I 75 1 0 12
fi·_v.~,- 00

CD 2-·-
1.1'0
i"Z~
'i"j
7 I \\
2,0 I- go - 3--- ">'il 2
L":",,.......... I

,
1\\
\
\
\
(0) 4 ~

2,5
I
P·'06
(7.7m)
\
I

\
\r P·l'S
(7.5m)

3.0 I

~
I

~ ~~
I
'.
1== P·l07

, ('0.4m) \
~2
~

1'2.6 -
Dralnage
\
)\ 12

~-1~
~
> 2.4

2.2
-
-
(b)
tid
Impervious

20
I
40 6080100 200
7 I

~
~

500
14
(6)
SfdlGrtA,N_

FIG. 10. Olga-c Test Embankment [after St-Arnaud et al.


(1992)]: (a) Variation of Pore Pressures with Time In Section A,
without Vertical Dralnsj (b) Approximate Stress-5traln Curves In
Section A, without Vertical Drains; (c) Variation of Pore Pres-
sures with Time In Section B, with Drain Spacing of 1.5 mj (d)
(4)
SfdIGrt B, _
(_91_".1.,1982)
4.5m
(1.5 • _ . )

Vertical effective stress (kPal Approximate Stress-5traln Curves In Section B, with Drain
Spacing of 1.5 m
FIG. 9. Observed Effective Stress-Strain (or Void Ratio) Rela-
tions In Various Sublayers of Sample Subjected to Incremental
Loading: (a) Berthlervllle Clay [from Lerouell et al. (1986) after lO(b) and lO(d). In section B (and also in sections C and D),
Mesrl and Feng (1986)]; (b) Yokohama Bay Mud [from Imal and where the drainage length is relatively small, the pore pres-
Tang (1992)] sures start decreasing just after the end of construction and,
consequently, the vertical effective stress is continuously in-
creasing during the consolidation of the clay deposit. On the
section A has no drains; section B has vertical drains with a other hand, in section A, where the drainage length is longer,
spacing of 1.5 m; in sections C and D, the spacing is of the clay mass was not able to expel the excess pore pressures
1.0 m. generated by creep and the pore pressures continue to increase
The geotechnical profile consists of a sensitive varved clay during 15-50 d after construction, before starting to decrease.
deposit, about 14 m thick, over a moraine layer of sandy and Similar behavior has been observed, in particular by Crooks
silty till. The upper 2 m of the clay deposit is an oxidized, et al. (1984) and Kabbaj et al. (1988). Consequently, the ver-
brown stiff crust. Then up to a depth of about 10 m, there is tical effective stress first decreases before it starts to increase.
the soft, highly plastic grey clay with water content essentially The stress-strain curves followed in the two sections are thus
constant between 80% and 100%, liquidity index between 1.15 different and this is due to creep effects.
and 1.5, and undrained shear strength between 20 kPa and 30
kPa. At larger depths, the plasticity index decreases and the Conclusion
undrained shear strength progressively increases.
The instrumentation of the clay foundation is described by These observations made both in laboratory an in situ con-
Lavallee et al. (1990) and the observations are presented by sistently show that the stress-strain curve followed by a soil
St-Arnaud et al. (1992). The distribution of settlement with element during primary consolidation depends on drainage
depth, as observed at different times, is rather linear in all conditions, and on strain rate and temperature. Hypothesis A
sections; 300 d after construction, the average vertical strain thus is incorrect.
was 6% under section A, 13% under section B, and 16.5%
under sections C and D. The lateral displacements measured PRACTICAL ASPECTS OF CLAY
under the berms of the embankments are relatively uniform COMPRESSIBILITY-ESTIMATION OF LONG-TERM
over the height of the clay deposit, being equal to about 6% SETTLEMENTS
of the settlement in section C and about 9% in section A. Soil
conditions are thus close to one dimensional. Figs. lO(a) and Long-term settlements of clay deposits are generally esti-
lO(c) respectively show the evolution with time of pore pres- mated on the basis of oedometer tests. However, as previously
sures observed in section A, where there are no drains, and in indicated, natural clay behavior is complex and there are im-
section B, in the middle of the grids formed by the drains. portant differences between soil conditions during a laboratory
Corresponding effective stress-strain curves are shown in Figs. oedometer test, and under an embankment in situ: (1) initial

538/ JOURNAL OF GEOTECHNICAL ENGINEERING / JULY 1996

J. Geotech. Engrg. 1996.122:534-543.


stress conditions and stress paths in general are different (Ler- eration during construction of embankments (point P; on Fig.
oueil 1988); (2) the strain rate is typically between 5 X 10- 8 12) have been compared with preconsolidation pressure values
and 5 X 10-6 S-I in oedometer tests and between 5 X 10- 12 obtained in conventional 24-h oedometer tests. The results are
and 10- 8 S-I under embankments; (3) the temperature is typ- shown in Fig. 11 for the sites where OCRs larger than 2.5
ically between 7°C and lOoC in the ground and is generally existed. Even if the scatter is large, the (]' ~onj(]' ~ ratio clearly
between 18°C and 22°C in the laboratory, and (4) sampling increases with the OCR, to reach a value of 2.0 at an OCR of
disturbance makes the soil in laboratory different from what about 5. Leroueil et al. (1978) explained this behavior by the
it is in situ. The determination of representative parameters for shape of the limit state curve of natural clays and the peculiar
the calculation of long-term settlements has thus to be made stress path followed in heavily overconsolidated clays during
on the basis of a semi-empirical approach. construction of an embankment. This is illustrated in Fig. 12
where a typical excess pore pressure increase l1u is shown as
a function of the total vertical stress increase 11(]' u [Fig. 12(a)]
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DE PARANA on 08/22/13. Copyright ASCE. For personal use only; all rights reserved.

Preconsolldation Pressure
and where the corresponding stress path is shown in Fig.
Morin et al. (1983) compiled preconsolidation pressure val- 12(b). During the early stages of construction, the excess pore
ues mobilized in situ under several embankments. Their results pressure generated in heavily overconsolidated clays is low (fJ
show that for overconsolidation ratios (OCRs) between 1.2 and = 11u1I1(]'u typically equal to 0.22) and the stress path such as
2.5, there is a good agreement with the preconsolidation pres- IP; is reaching the limit state curve at P;, at a vertical effective
sure obtained in conventional 24 h oedometer tests «(]'~onY)' At stress (]' ~ value lower than the preconsolidation pressure of
lower OCRs, laboratory tests generally slightly underestimate the clay (]';. The yield stress considered in Fig. 11 corresponds
in situ values, whereas at high OCRs, they overestimate in situ to this point P;. If the loading of the embankment is continued
values. Morin et al. (1983) suggested a small correction to the fJ increases but remains lower than 1.0 for overconsolidated
measured laboratory values clays, so that the vertical effective stress (]' ~ continues to in-
crease up to (]'; and the corresponding stress path is denoted
(1) by P; P~A' in Fig. 12(b).
So, if the loading of an embankment is stopped while the
with (XI = 1.1 for OCR < 1.2, (XI = 1.0 for 1.2 < OCR < 2.5,
stress state is between P; and P~, the settlement corresponding
and on the basis of few data, (XI = 0.9 for 2.5 < OCR < 4.5.
Other data obtained since that time confirm the general ten- to the normally consolidated range starts at a vertical effective
stress smaller than the measured in laboratory preconsolidation
dency for OCRs < 2.5 and allow a reevaluation of the situation
pressure of the clay, the minimum value of this vertical stress
for larger OCRs. Available vertical yield stress (]' ~ values ev-
being given by Fig. 11 and the approximation equation
idenced by a rapid increase in the rate of pore-pressure gen-
2.5 , aJ,conv
(2)
I
(]'.." = 0.64 + 0.260CR
~~
I
I I
I I
I I
,
I I
I
If the loading is stopped when fJ is close to 1.0 and the
2.0 stress path is between P~ and A', then the settlement corre-
.- -.,
J
,_.!
..,
'L
~ ___ J sponding to the normally consolidated range starts at a vertical
effective stress close to the preconsolidation pressure of the
clay. In such a case, this preconsolidation pressure can be de-

~0
V
"'" 0'vy = 0'P cony. termined by using (1).
~ r64+0.261R
I I Long-Term Settlements In Normally Consolidated

-,
,_.
~
Rupert-7 } I Range

•• At stresses well in excess of the preconsolidation pressure,


Arles Leroueil et al.. 1978
Interstate 95
0.5
Saint-Esprit BoucHn.199O
the clay behavior is less affected by sampling disturbance than
~ Olga-C 5t-Arnaud et aI., 1992 at small strains, and the stress paths are similar in laboratory
(} Olga-B 5EBJ,1983 and in situ, Le., close to the K one line. The main factors that
a I could induce differences between laboratory and in situ be-
2 345 6 7 haviors are thus strain-rate, temperature, and structuring phe-
OCR = O'pconv./O'YO
nomena. Detailed studies performed by Kabbaj et al. (1988)
FIG. 11. Variation of Ratio a :.conia ~y with Overconsolidation on four different sites from Canada and Sweden show that,
Ratio of Clay (Only Sites with OCRs Larger Than 2.5 Are Consid- compared to end-of-primary (EOP) compression curves ob-
ered) tained in laboratory on high-quality samples obtained by Laval
sampler (La Rochelle et al. 1981), the in situ stress-strain
curves are well below (Fig. 13). At a given effective stress,
the in situ strain (or settlement) is larger than that expected on
the basis of EOP laboratory curves. Other studies reviewed by
Leroueil (1988) also show that in all cases (about 10) but one,
the vertical effective stress in situ is smaller than that predicted
on the basis of conventional 24 h oedometer tests at the same
strain, thus confirming the results of Kabbaj et al. (1988). All
(b)
these detailed studies and experiences gained in Sweden (Lars-
son 1986) and in France (Magnan 1992) indicate that rate ef-
4CJy (Cl'1 + Cl'3)12 fects during in situ primary consolidation are predominant.
(nr l8.....u 01 .... 1878) Considering only the effects of strain rate, Leroueil et al.
FIG. 12. Pore Pressure Generated and Stress Path Followed (1988) suggested that a strain 11£, be added to the strain esti-
In Heavily Overconsolldated Clay under Center of Embankment mated on the basis of the conventional 24 h oedometer test
[after Lerouellet al. (1978)] (which, in itself already incorporates some secondary com-
JOURNAL OF GEOTECHNICAL ENGINEERING f JULY 1996 f 539

J. Geotech. Engrg. 1996.122:534-543.


CT~ I kPo CT~ I kPo
n0r-_...;20;,=-__4;.r:0;....".~60~_ _8;;;O~......:I~OO~--:.:12:;O0i-_-:2T0'-!I":':4:!!O;-.....:6::y:O:....~8~0i'-_~IOO/lf-_1.l;120.

:==:=:=~\~-ft---+-----..:-.-+--:~_. ~---I \._O"'_\F,


.. \~-\ ";F I
.: 6t---t---t---\ \
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DE PARANA on 08/22/13. Copyright ASCE. For personal use only; all rights reserved.

8 J---I--~I-----!- \ - \~+---; .---1--"'----11----+---+-\'\-+__---I


8erthierville • ~ Gloucester
o -EOP (4,4m) I-+--~ 1\_ -EOP (3 , 6m)
_ •• In situ 13.9-4,8m) " \ - - In situ (2,4-4,9m) \

02r'~
--==:r:
1;:i15~==~~~ --=~:::::==~~=::=:::=~
Th ~"- I I
r-~ Vosby
21---+---+--., \ I: ' -EOP (5,9m)
-.-In situ (4,3-7,3m)

pression) to evaluate the strain at the end of in situ primary of strain rate and temperature differences between the labo-
consolidation. At. increases with increasing CJ(1 + eo) and ratory and in situ. To avoid confusion, this new strain to be
decreasing strain rate at the end of in situ primary consolida- added to the strain estimated on the basis of the conventional
tion £EQP as follows: 24 h oedometer test will be described as At:.

...t.
A
= m(l C+c eo) [log10- 7
- Iog(tEop)]
' (3) At: = Cc
~1+~
[ H 2
(a.
log 6 x 10-9 - log -16kU;)]
~
(4)

where £EOP = (O.16k U~/"YwH2); C c = compression index under At: is shown in Fig, 14 as a function of CJ(1 + eo) and
the final effective stress; eo = initial void ratio; k = hydraulic £EQP'
conductivity of the clay; m = a strain rate factor deduced from Applying (3) to the test embankments of Berthierville,
Fig. 1 and approximately equal to 32; u~ = initial excess pore Saint-Alban-D, and Vlisby, Leroueil et al. (1988) found At,
pressure; and H = maximum drainage length. values equal to 2.8%, 4% and 8%, respectively, and considered
Considering that the effects of temperature are, up to a cer- these values realistic. Taking into account the effect of tem-
tain degree, compensating for the effects of strain rate, the perature, (4) gives for the same embankments AE~ values re-
additional strain At, value given by (3) is too high. From the spectively equal to 1.4%,2.5%, and 5.6%. On the basis of the
data on the effects of temperature compiled by Boudali et al, field data, these values can still be considered realistic but
(1994), a difference in temperature of about 12°C would have certainly are lower limits for the strain to add to that given by
an effect similar to a change in strain rate of 1 to 1.5 logarithm the conventional 24 h oedometer test.
cycles. Eq. (3) can be modified to take into account the effects These results, which include the effects of temperature and
540 I JOURNAL OF GEOTECHNICAL ENGINEERING I JULY 1996

J. Geotech. Engrg. 1996.122:534-543.


12.5 Co
, +.0·
than 0.25] or for embankments of special importance (test
1.0 embankments, embankments for which the magnitude of
10.0 1-_....3~_+- +- +- -I
the settlements may have important consequences), vis-
0.8 cous effects could be significant. In such cases, as previ-
ously indicated, the EOP and the viscous approaches
could be considered. On the basis of the results, the en-
gineer has to judge which method seems to be the most
appropriate for the considered project.
• For usual projects in clays of low compressibility [Cc /(1
+ eo) values less than 0.25], tit. is smaller than 2.5% and
is often compensated by some disturbance of the clay
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DE PARANA on 08/22/13. Copyright ASCE. For personal use only; all rights reserved.

samples and some reduction of the final effective stress


uj'O
due to settlement and a raise of the water table in the
e • 0.'8k~oO embankment. In these cases, it is thought that conven-
EOP 'Yw H tional 24 h oedometer test results, which already include
some secondary compression, could be used without any
FIG. 14. Estimation of End-of-Prlmary In situ Strain 4£: In Ex-
ee.. of That Determined from Conventional 24 h Laboratory
correction.
Oedometer Test, when Strain Rate and Temperature Are Consid- • When vertical drains are used to accelerate the consoli-
ered dation process or when the clay deposit is thin or strati-
fied, the strain rate at the end of primary consolidation is
decrease the correction due to rate effects to apply to conven- larger than 3 X 10- 10 S-1 in most cases. Consequently,
tional test results, would be supported by the French experi- a£~ remains relatively small and, for the same reasons as
ence (Magnan 1992), which indicates that there is some vis- those indicated in the preceding item, it is thought that
cous compression during primary consolidation, but that the the conventional 24 h oedometer test results could be used
amount is less than that implied by rate effects only. without any correction for calculating the settlements.
It is worth noting however that the structuring phenomena
previously described, when they exist, become more important CONCLUSION
at low strain rates and could modify the consolidation process
of thick clay layers. In such a case, they would decrease the The present analysis shows that there are several factors
importance of the viscous component of deformation a£~. This affecting the compressibility of natural clays: strain rate, tem-
could explain why Mesri and Choi (1985b) in laboratory, and perature, sampling disturbance, stress path, and structuring
Mesri et al. (1994) for several embankments found that the phenomena. At small strains, while the soil is in the overcon-
end-of-primary e-Iog cr ~ compression curve was independent solidated range or passing its preconsolidation pressure, the
of the thickness of .the clay layer. It thus appears that the vis- effects of the four first factors are important and the soil be-
cous approach including the strain rate and the temperature havior is complex. It follows that the preconsolidation pressure
effects, and characterized by (4) and Fig 14, would give an has to be determined on the basis of a semiempirical approach.
upper limit of settlements, whereas the EOP approach sug- In the normally consolidated range, the possible factors are
gested by Mesri and Choi (1985a) and Mesri et al. (1994) strain rate, temperature, and structuring phenomena. The com-
could be a lower limit of settlements. In cases of soils not bined effect of strain rate and temperature has been quantified,
prone to develop structure, the settlement should be close to and a viscous strain a£~ is proposed to be added to the strain
that given by the viscous approach. In cases of soils devel- estimated on the basis of the conventional 24 h oedometer test
oping structure with time or at low strain rates, the settlement to evaluate the strain at the end of in situ primary consolida-
should be smaller than that given by the viscous approach, and tion. In low compressibility clays, and also in cases where
possibly closer to that given by the EOP approach. Unfortu- vertical drains are used so that the strain rate at the end of
nately, there is no validated method for evaluating the ability primary consolidation is relatively high, a£~ is small and can
of a soil to develop structure in in situ conditions. be neglected in many cases. This corresponds to the experi-
Numerous numerical models, in which strain rate or time ence gained in the United States and in other countries where
effects are taken into account, have been recently developed the clays are of low compressibility and the 24 h oedometer
for simulating the consolidation of clay deposits (Magnan test gives reasonable settlement predictions. On the other hand,
et al. 1979; Oka et al. 1986; Larsson 1987; Szavits-Nossan in highly compressible clays, a£~ can be significant and
1988; Yin and Graham 1989; Borja 1992; Liang and Ma 1992; should be considered. This corresponds to the experience
Rajot 1992). None of them considers the influence of temper- gained in Eastern Canada and Scandinavia, where the clays
ature, sampling disturbance and structuring phenomena, and are sensitive and highly compressible. There could be cases
the soil engineer must be aware of these weaknesses when however where structuring phenomena exist and can reduce
using them. the viscous component a£~.
On the basis of all these remarks and observations, Leroueil
(1995) suggested that, in a general manner, both the viscous ACKNOWLEDGMENTS
and the EOP approaches be considered for estimating in situ
settlements. When the difference is small, any method would The writer would like to thank M. Lojander, A. L. Paakkunainen, and
be acceptable. When the difference is significant, the engineer S. Hoikkala for providing data on the effects of strain rate on Finnish
has to figure out the sampling disturbance and its conse- clays; Hydro-Qu~bec and in particular Y. Hammamji and B. Touileb for
allowing publication of some data related to Olga-C Test embankment;
quences on the reference compression curve, the ability of the and D. Perret for the permissions to present results from his PhD thesis.
soil to develop structure, and the practical consequences of
settlements larger than expected. Finally, using his or her judg-
ment and local experience, the engineer will decide how to APPENDIX. REFERENCES
calculate the settlements of the considered structure. More Berre, T., and Iversen, K. (1972). "Oedometer tests with different spec-
practically, the following approach is recommended: imen heights on a clay exhibiting large secondary compression." Glo-
technique, London, U.K., 22(1), 27-52.
• In highly compressible clays [Cc!(l + eo) values larger Bjerrum, L. (1967). "Engineering geology of Norwegian normally con-

JOURNAL OF GEOTECHNICAL ENGINEERING / JULY 1996/541

J. Geotech. Engrg. 1996.122:534-543.


solidated marine clays as related to the settlement of buildings." Geo- Leroueil, S. (1988). "Recent developments in consolidation of natural
technique, London, U.K., 17(2), 83-119. clays." Can. Geotech. J., 25(1), 85 -107.
Bjerrum, L. (1972). "Embankments on soft ground." Proc. Spec. Con! Leroueil, S. (1995). "Could it be that clays have no unique way of be-
on Perf. of Earth and Earth-Supported Struct., ASCE, New York, N.Y., having during consolidation?" Int. Symp. on Compression and Con-
2, I-54. soliOOtion of Clayey Soils, Keynote Lecture, Hiroshima, Vol. 2.
Borja, R. I. (1992). "Generalized creep and stress relaxation model for Leroueil, S., and Kabbaj, M. (1987). "Discussion of 'Settlements analysis
clays." J. Geotech. Engrg., ASCE, 118(11), 1765-1786. of embankments on soft clays,' by G. Mesri and Y. K. Choi." J. Geo-
Bouclin, G. (1990). "Anisotropie de perm6abilit6 des argiles de Saint- tech. Engrg., ASCE, 113(9), 1067-1070.
Esprit," MS thesis, Laval Univ., Qu6bec, Canada. Leroueil, S., Tavenas, F., Mieussens, C., and Peignaud, M. (1978). "Con-
Boudali, M., Leroueil, S., and Srinivasa Murthy, B. R. (1994). "Viscous struction pore pressures in clay foundations under embankments, Part
behaviour of natural clays." Proc.. 13th ICSMFE, 1,411-416. II: generalized behaviour." Can. Geotech. J., 15(1), 66-82.
Burghigholi, A., Cavalera, L., Chieppa, v., Jamiolkowski, M., Mancuso, Leroueil, S., Tavenas, F., Samson, L., and Morin, P. (1983). "Precon-
C., Marchetti, S., Pane, V., Paoliani, P., Silvestri, F., Vinale, F., and solidation pressure of Champlain clays. Part II: Laboratory determi-
nation." Can. Geotech. J., 20(4), 803-816.
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DE PARANA on 08/22/13. Copyright ASCE. For personal use only; all rights reserved.

Vittori, E. (1991). "Geotechnical characterization of Fucino clay."


Proc. 10th Eur. Con! on Soil Mech. and Found. Engrg., 1,27-40. Leroueil, S., Kabbaj, M., Tavenas, F., and Bouchard, R. (1985). "Stress-
Burland, J. B. (1990). "On the compressibility and shear strength of strain-strain rate relation for the compressibility of sensitive natural
natural clays." Geotechnique, London, U.K., 40(3),329-378. clays." Geotechnique, London, U.K., 35(2),159-180.
Casagrande, A. (1932). "The structure of clays and its importance in Leroueil, S., Kabbaj, M., Tavenas, F., and Bouchard, R. (1986). "Closure
foundation engineering." J. Boston Soc. Civ. Engrg., 19(4), 168-209. to 'Stress-strain-strain rate relation for the compressibility of sensitive
Crawford, C. B. (1965). "The resistance of soil structure to consolida- natural clays.''' Geotechnique, London, U.K., 36(2), 288-290.
tion." Can. Geotech. J., 2(2), 90-97. Leroueil, S., Kabbaj, M., and Tavenas, F. (1988). "Study of the validity
of a a; - E" - ~v model in in-situ conditions." Soils and Found.,
Crooks, J. H. A., Becker, D. E., Jefferies, M. G., and McKenzie, K.
Tokyo, Japan, 28(3), 13-25.
(1984). "Yield behaviour and consolidation-I: pore pressure re-
Liang, R. Y., and Ma, F. (1992). "A unified elasto-viscoplasticity model
sponse." Proc. ASCE Symp. on Sedimentation Consoliootion Models: for clays, Part I: Theory." Compo and Geotechnics, 13(2), 71-87.
Predictions and Valiootion, ASCE, New York, N.Y., 356-381. Locat, J., and Lefebvre, G. (1985). "The compressibility and sensitivity
Eriksson, L. G. (1989). "Temperature effects on consolidation properties of an artificially sedimented clay soil: The Grande Baleine marine clay,
of sulphide clays." Proc. 12th ICSMFE, Rio de Janeiro, 3, 2087 -2090. Qu6bec, Canada." Marine Geotechnol., 6(1), 1-27.
Hanzawa: H. (1989). "Evaluation of design parameters for soft clays as Magnan, J. P. (1992). "Le rille du f1uage dans les calculs de consolidation
related to geological stress history." Soils and Found., Tokyo, Japan, et de tassem~nt des sols compressibles." Bulletin de Liaison des La-
29(2), 99 -111. boratoires des Ponts et Chaussees, 180, 19-24.
Hanzawa, H. (1991). "A new approach to determine the shear strength Magnan, J. P., Baghery, S., Brucy, M., and Tavenas, F. (1979). "Etude
of soft clay." Proc., Int. Con! on Geotech. Engrg. for Coast. Dev., num6rique de la consolidation unidimensionnelle en tenant compte des
Geo-Coast '91, Yokohama, Japan, 1,23-28. variations de la perm6abilit6 et de la compressibilit6 du sol, du f1uage
Hanzawa, H., Fudaya, T., and Suzuki, K. (1990). "Evaluation of engi- et de la non-saturation." Bulletin de Liaison des Laboratoires des Ponts
neering properties for an Ariake clay." Soils and Found., Tokyo, Japan, et Chaussees, 103, 83-94.
30(4), 11-24. Magnan, J. P., Khemissa, M., and Josseaume, H. (1994). "Influence du
Hight, D. w., Boese, R., Butcher, A. P., Clayton, C. R. I., and Smith, P. pr61~vement sur Ie comportement des argiles." Proc. 13th ICSMFE,
R. (1992). "Disturbance of the Bothkennar clay prior to laboratory New Delhi, 1,317-320.
testing." Geotechnique, London, U.K., 42(2), 199-217. Mesri, G., and Choi, Y. K. (1985a). "Settlement analysis of embankments
Hoikkala, S. (1991). "Continuous and incremental loading oedometer of soft clays." J. Geotech. Engrg., ASCE, 111(4),441-464.
tests," MS thesis, Helsinki Univ. of Technol., Espoo, Finland (in Finn- Mesri, G., and Choi, Y. K. (1985b). "The uniqueness of the end-of-
ish). primary (EOP) void ratio-effective stress relationship." Proc. 11th IC-
Hueckell, T., and Baldi, G. (1990). "Thermoplasticity of saturated clays: SMFE, San Francisco, 2, 587 -590.
experimental constitutive study." J. Geotech. Engrg., ASCE, 116(12), Mesri, G., and Feng, T. W. (1986). "Discussion of 'Stress-strain-strain
1778-1795. rate relation for the compressibility of sensitive natural clays,' by S.
Imai, G., and Tang, Y. X. (1992). "A constitutive equation of one-di- Leroueil, M. Kabbaj, F. Tavenas, and R. Bouchard." Geotechnique,
mensional consolidation derived from inter-connected tests." Soils and London, U.K., 36(2), 283-287.
Found., Tokyo, Japan, 32(2), 83-96. Mesri, G., Lo, D. O. D., and Feng, T. W. (1994). "Settlement of em-
Jamiolkowski, M., Ladd, C. C., Germaine, J. T., and Lancellotta, R. bankments on soft clays." Proc. Con! on Vertical and Horizontal De-
(1985). "New developments in field and laboratory testing of soils." formations of Found. and Embankments: Settlement '94, Geotech.
Proc. 11 th ICSMFE, San Francisco, I, 57 -153. Spec. Publ. No. 40, ASCE, New York, N.Y., I, 8-56.
Kabbaj, M. (1985). "Aspects rh6ologiques des argiles naturelles en con- Mizukarni, J. I., and Motoyashiki, M. (1992). "Consolidation yield stress
solidation," PhD thesis, Laval Univ., Qu6bec, Canada. by constant rate of strain test." Proc. 28th Annual Con! of the Japa-
Kabbaj, M., Tavenas, F., and Leroueil, S. (1988). "In situ and laboratory nese Soc. of Soil Mech. and Found. Engrg., 419-420 (in Japanese).
stress-strain relations." Geotechnique, London, U.K., 38(1), 83-100. Morin, P., Leroueil, S., and Samson, L. (1983). "Preconsolidation pres-
Kolisoja, P., Sahi, K., and Hartikainen, J. (1989). "An automatic triaxial- sure of Champlain clays. Part I: In-situ determination." Can Geotech.
oedometer device." Proc. 12th ICSMFE, Rio de Janeiro, 1,61-64. J., 20(4), 782-802.
Lacasse, S., Ben'e, T., and Lefebvre, G. (1985). "Block sampling of sen- Nash, D. F. T., Sills, G. C., and Davison, L. R. (1992). "One-dimensional
sitive clays." Proc. 11th ICSMFE, San Francisco, 2, 887-892. consolidation testing of soft clay from Bothkennar." Geotechnique,
La Rochelle, P., Sarrailh, J., Tavenas, F., Roy, M., and Leroueil, S. (1981). London, U.K., 42(2),241-256.
Oka, F., Adachi, T., and Okano, Y. (1986). "Two-dimensional consoli-
•'Causes of sampling disturbance and design of new sampler for sen-
dation analysis using an elasto-viscoplastic constitutive equation." Int.
sitive soils." Can. Geotech. J., 11(1), 142-164.
J. Numer. and Anal. Methods in Geomech., 10(1), 1-16.
Larsson, R. (1986). "Consolidation of soft soils." Swedish Geotech. Inst.
Okumura, T., and Suzuki, K. (1991). "Analysis of consolidation settle-
Rep. No. 29, Linkoping, Sweden. ment considering the change in compressibility." Proc. Int. Con! on
Larsson, R. (1987). "Long term behaviour of two test fills in Sweden." Geotech. Engrg.for Coast. Dev., Geo-Coast-91, Yokohama, 1,57-62.
Proc. Int. Symp. on Geotech. Engrg. of Soft Soils, Mexico City, I, Paakkunainen, A. L. (1990). "Continuous loading oedometer tests in set-
239-247. tlement calculation," MS thesis, Tempere Univ. of Techno!., Tempere,
Larsson, R., and SaIlfors, G. (1985). "Automatic continuous consolida- Finland (in Finnish).
tion testing in Sweden." Consoliootion of soils, ASTM STP 892, Perret, D. (1995). "D6veloppement de la r6sistance dans un s6diment fin
ASTM, Philadelphia, Pa., 299-328. en formation," PhD thesis, Laval Univ., Qu6bec, Canada.
Lava1l6e, J. G., St-Arnaud, G., Morel, R., and Hammamji, Y. (1990). Perret, D., Locat, J., and Leroueil, S. (1995). "Strength development with
"Remblai d'essai pour v6rifier la consolidation de l'argile avec des burial during early diagenesis in fine grained sediments from Saguenay
drains synth6tiques." Proc. 43rd Can. Geotech. Con!, Qu6bec, Can- Fjord, Qu6bec, Canada." Can. Geotech. J., 32(2), 247-262.
ada, 525-531. Rajot, J. P. (1992). "A theory for the time dependent yielding and creep
Leonards, G. A. (1972). "Discussion of 'Shallow foundations.''' Proc., of clay," PhD thesis, Virginia Polytech. Inst. and State Univ., Blacks-
ASCE Spec. Con! on Perf. of Earth and Earth-Supported Struct., burg, Va.
ASCE, New York, N.Y., 3, 169-173. Schmertmann, J. H. (1955). "The undisturbed consolidation behavior of
Leonards, G. A., and Altschaeffl, A. G. (1964). "Compressibility of clay." Trans ASCE" 120(1), 12-16.
clay." J. Soil Mech. Found. Div., ASCE, 90(5),133-155. Soci6t6 d'Energie de la Baie James. (1983). "Experts Committee on Soft
542/ JOURNAL OF GEOTECHNICAL ENGINEERING / JULY 1996

J. Geotech. Engrg. 1996.122:534-543.


Clays from the NBR Complex," Final Rep., Montreal, Quebec, Can- isotache method." Proc., 4th ICSMFE, London, U.K., 1,200-206.
ada. Szavits-Nossan, V. (1988). "Intrinsic time behavior of cohesive soils dur-
"Standard test method for one-dimensional consolidation properties of ing consolidation," PhD thesis, Univ. of Colorado, Boulder, Colo.
soils using controlled-strain loading." (1993). D 4186-89: Annual Book Taylor, D. W. (1942). "Research on consolidation of clays." Series 82,
of ASTM Standards, ASTM, Philadelphia, Pa., 04.08, 637-641. Massachusetts Inst. of Techno!., Cambridge, Mass.
St-Arnaud, G., Morel, R., and Lavallee, J. G. (1992). "Comportement de Tidfors, M., and Siillfors, G. (1989). "Temperature effect on precon-
la fondation argileuse traitee avec des drains synthetiques sous Ie rem- solidation pressure." Geotech. Test J., 12(1),93-97.
blai d'essai Olga-C." Internal Rep., Hydro-Quebec, Service geologie Yin, J., and Graham, J. (1989). "Viscous-elastic-plastic modelling of one-
et structures, Montreal, Canada. dimensional time-dependent behaviour of clays." Can. Geotech. J.,
Suklje, L. (1957). "The analysis of the consolidation process by the 26(2), 199-209.
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DE PARANA on 08/22/13. Copyright ASCE. For personal use only; all rights reserved.

JOURNAL OF GEOTECHNICAL ENGINEERING / JULY 1996/543

J. Geotech. Engrg. 1996.122:534-543.

You might also like