You are on page 1of 23

Tenth Canadian Geotechnical Colloquium:' Recent developments in consolidation of

natural clays
SERGELEROUEIL
Universite' Laval, Qukbec, Que., Canada GIK 7P4
Received Februa~y13, 1987
Accepted October 15, 1987
A global analysis of the consolidation of natural clays is realized considering the consolidation process to be a combination
of the effects of compressibility and of permeability. The compressibility or stress-stzin curve followed is strongly
influenced, both in the laboratory and in situ, by the strain rate. The self-boring permeameter appears to be an excellent tool for
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

permeability measurement; however, in homogeneous clays direct measurement in the laboratory also gives representive
results. The coefficients of consolidation determined graphically strongly underestimate the in situ coefficient. The consolida-
tion must thus be analysed by considering compressibility and permeability parameters measured separately.
Key words: consolidation, clay, compressibility, permeability.

Une analyse globale de la consolidation des argiles naturelles est effectuCe en considkrant que le processus de consolidation
est constituC d'une combinaison des effets de compressibilitC et de permtabilitC. La compressibilitt ou la courbe effort-
dCformation suivie est fortement influencte par la vitesse de dCformation, tant en laboratoire que sur le terrain. Le permkamktre
auto-foreur semble &re u11 excellent outil pour mesurer la permkabilitC; cependant, la mesure directe en laboratoire donne
Cgalement des rCsultats reprksentatifs dans les argiles homogknes. Les mCthodes graphiques sous-estiment fortement le
coefficient de consolidation in situ. La consolidation doit donc &treanalyste en partant des paramktres de compressibilitCet de
permCabilitC mesures en place.
Mots cle's : consolidation, argile, compressibilitC, permCabilitC.
[Traduit par la revue]
Can. Geotech. J. 25, 85-107 (1988)
For personal use only.

Foreward the soft soils of the Euphrate valley; one of these temples, the
The topic of the Colloquium is the consolidation of clays. White temple of Eridou, has now its foundations 12 m below
The subject is limited to the consolidation of natural clays the street. Since that time, a great number of other structures
without consideration of dredged or other materials consoli- have settled and suffered damage. In the last century, with the
dating under their own weight. Moreover, very little attention industrialization of the planet and the growth of road and rail-
is given to numerical models and, in particular, models related way networks, it has become necessary to occupy lands and
to two- and three-dimensional consolidation. For these topics, marshes that were before avoided, thus increasing the number
the reader is referred to general reports presented recently by of structures built on soft soils.
Schiffman et al. (1984), Murray (1978), and Balasubramanian Another consequence of industrialization is the rapid growth
and Brenner (1981). of cities and the increasing need for water. This water is often
Because of the large extent of soft clay deposits in populated taken from the subsoil, resulting in settlements of large areas
areas of Canada and the importance of associated problems, of land, such as in Mexico City, Tokyo, Nagoya, Bangkok,
research on consolidation has been very active in the country Stockholm, Shanghai, and Venice. Some of these cities are
for at least 30 years. The researchers involved are numerous only slightly above sea level and could sink unless the consoli-
but I think it is worth mentioning the contribution made by dation process is stopped. Already, lowlands of great value for
Mr. Carl B. Crawford. He was one of the first to evidence the agriculture have become submerged; Yamamoto (1977) esti-
importance of strain rate effects on clay behaviour and the fact mated that in Japan, owing to this phenomenon, 1200 km2 has
that oedometer test results depend on the testing technique come below the sea level.
used; in numerous papers, he compared laboratory test results Thus, consolidation is an important problem and has always
with in situ behaviour to try to establish which of the labora- been with us. However, our comprehension of the processes
tory tests is the most representative. The work presented here involved has improved very slowly. Even if the Sumerians
follows the path laid out by Crawford and the author would may have understood some aspects of consolidation and pore
like to dedicate this Colloquium to him. pressure dissipation (KCrisel 1985), the most important bench-
mark of our understanding of consolidation was set by Ter-
Introduction zaghi in 1923. In the last 60 years considerable work and
Consolidation has been a major concern to builders for a very progress have been achieved to improve our understanding of
long time. KCrisel(1985) reported that 6000 years ago, Sumer- clay behaviour. It must be recognized, however, that in 1986
ians built ziggourats-embankments to support temple,Q-on soil engineers are still using the theory of Terzaghi (1923) and
wondering. whether there is a combination of ~ r i m a r vand sec-
'Editor's footnote: the Canadian Geotechnical Colloquium is pre- ondary compression during primary consolidation.
sented annually by a young engineer to the annual meeting of the The experimental work performed in the laboratory
Canadian Geotechnical Society. The subject of the lecture, as chosen and in situ needs to be anal~sedwith the aim of bringing out
by the Associate Committee on Geotechnical Research of the National the rheological behaviour of natural clays. The data from
Research Council of Canada, is one of national interest and im~or- embankment foundations with vertical drains is, however, dis-
tance. regarded in this paper, the analysis being complicated by pore
Printed in Canada / Imprim6 au Canada
86 VOL. 25, 1988
CAN. GEOTECH. .I.

pressures that vary with the distance to the drain and by a con-
solidation process that depends on the type of drain used and
its installation (Magnan 1983).
Before examining the data related to soil consolidation, it is
of interest to recall that the basic and most general equation for
one-dimensional consolidation can be written (Berry and
Poskitt 1972) as

[I]
aev - 1 + eo a
at yw az
(+
I
k
e
au)
a~
where ev = vertical strain, z = depth related to initial thick-
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

ness, t = time, eo = initial void ratio, e = present void ratio,


k = permeability, yw = unit weight of water, and u' = excess J
pore-water pressure. .-c
The only assumptions required to establish [I] are that E
?ii
Darcv's law be valid and the soil be saturated. It is worth
noting that, from [I], at a given time, the strain rate iv= aglat
FIG. 1. Stress-strain curves and the consolidation process.
depends essentially on the pore pressure distribution and the
and very little<through (1 f eo) and (1 + e)) on
the compressibility of the soil. For example, if two different recognized that time could be an important factor, a
soils, one strongly overconsolidated (soil A) and one soft and decrease of void ratio under constant effective stress was
sensitive (soil B) (Fig. I), have, at a given time t, the same observed, The approaches developed by Koppejan (1948) and
excess pore pressure distribution, the same permeability, and Bjermm (1967), with rheological equations of the type
the same effective stress (a; at point I, Fig. l), they will
undergo about the same strain increase Atv = i,At during the [41 R(o;,e,t) = 0 or R1(u;,ev,t) = 0
time increment At. are directly derived from such observations. A major difficulty
For personal use only.

Owing to different shapes of the stress -strain curves, soil A with models of this type is to define an origin of time; this is
will move from point I to point VA with a corresponding particularly true when the applied load varies with time.
increase of the effective stress and a decrease of the excess Bjermm's model implies that in multiple-stage loading oed-
pore pressure; on the contrary, going from point I to point VB, ometer tests, the longer the application of the load, the smaller
the soil B undergoes a decrease in cr; and thus an increase in u' is the preconsolidation pressure. This was observed in partic-
(as shown by Crooks et al. (1984), this latter behaviour is not ular by Crawford (1964) and by Tavenas and Leroueil(1977).
exceptional). At time t + At, the new effective stress 0; at In the last two decades, numerous constant rate of strain (CRS)
points VAor VBwill give the new excess pore pressure and thus oedometer tests have been performed on natural clays at
the new strain rate. The consolidation process is thus the com- various strain rates (on eastern Canada clays: Crawford 1965;
bination of two phenomena: the permeability, which controls Vaid et al. 1979; Leroueil et al. 1983a; Silvestri et al. 1985;
the rate of settlement at any time, and the compressibility, on Swedish clay: Sallfors 1975; on Belfast clay: Graham et al.
which controls the evolution of the consolidation process. 1983). All the tests show that the strain rate influences the
In the simple case where the applied total stress, the perme- compression curve of clays: at a given strain, the higher the
ability k , and the deformation modulus M = du;/dev are con- strain rate, the higher the effective stress. Constant rate of
stant, the equation controlling the consolidation takes the loading tests (Jarrett 1967; Burghignoli 1979), stage loading
well-known form tests (Larsson 1981), and controlled gradient tests (Shields
1974; Leroueil et al. 1983a) confirm the important effect of
strain rate.
To take these effects into account, two families of rheolog-
with ical models have been developed:
[3] c kM
=-='-Sf

Yw (Taylor and Merchant 1940; Gibson and Lo 1961; Wu et al.


The coefficient of consolidation cv is a function of the com- 1966; Poskitt and Birdsall 1971; Sekiguchi and Toriihara
pressibility parameter M and the permeability parameter k. 1976), and
To understand the consolidation process of natural clays, it
seems logical to analyse separately the compressibility and the
permeability; this will be made in the first two sections of the (~uklje1957, 1969; Barden 1965; Poorooshasb and Siva-
Colloquium. In the last two sections, the overall consolidation patham 1969; Poorooshasb et al. 1981; Oka 1981; Kabbaj
will be examined and practical conclusions will be suggested. et al. 1985).
Leroueil et al. (1985b) camed out different types of oedom-
Compressibility eter tests (CRS, controlled gradient, stage loading, long-term
Compressibility in laboratory creep) on several Champlain sea clays to determine which one
In Terzaghi's theory of consolidation, the effective stress - of equations [5] and [6] is the most representative of soil
strain response of the soil is unique, linear, and time indepen- behaviour. In the CRS tests, the effective stress was always
dent. However, as early as 1936 (Buisman), it has been increasing ($4 > 0) while in long-term creep tests, the rate of
LEROUEIL

Effective stress T;, kPo


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13
For personal use only.

FIG. 2. Typical stress - strain - strain rate relation for natural clays

change in effective stress was 0 (6; = O), and it was thus pos- effective stress and which varies with strain rate, must be con-
sible to evaluate the influence of 6; on the rheological behav- sidered as a rheological parameter. This yield stress should
iour of clays. From the results, it appears that this factor does have a particular appellation but engineers usually refer to it as
not have a significant influence, so that, at least for laboratory "preconsolidation pressure" and this term will also be used in
conditions, an effective stress - strain - strain rate model this paper.
([6]) seems to be most representative of the rheological behav- During secondary compression, when u' is almost equal to
iour of natural clays. This means that whatever the test, at a zero, the classical secondary compression approach with the
given strain, there is a unique relation between the effective use of C,, and the strain rate approach are equivalent, i, and
stress and the strain rate. This behaviour can be described by C,, being related by
curves of equal strain rate in a c,-o: diagram as shown in
Fig. 2; SuMje (1957) used the e-a: diagram and called the
e = cStcurves isotaches.
Similar strain rate effects have been observed on numerous If, as indicated by Mesri and Godlewski (1977), CaelCc =
soft clays. Figure 3 shows the normalized preconsolidation 0.04 in conventional oedometer tests, then at 24 h, the strain
pressure - logarithm of strain rate relationship for eastern rate is
Canada clays; Sallfors (1975), Larsson (198 l), Graham et al.
(1983), and Silvestri et al. (1985) have found very similar
results. As shown by Mesri and Choi (1979), the fact that the
variation of a; per logarithm cycle of time or strain rate is For eastern Canada clays for which Cc/(l + eo) is typically
about the same for a variety of clays is consistent with observa- between 0.3 and 0.7, the conventional oedometer curve corre-
tions made by Mesri and Godlewski (1977) and Mesri and sponds to a strain rate of about s-I, as observed by
Choi (1984) that the ratio C,,/CC (Cue = AelA log t and Cc = Leroueil et al. (1983b). For less sensitive clays having smaller
AelA log a:) is approximately constant and equal to 0.04 for Cc/(l + eo) ratios, the strain rate at 24 h would be slightly
inorganic clays. smaller.
The preconsolidation pressure, as determined in oedometer According to Mesri and Feng (1986), the strain rate corre-
tests, is thus strain rate dependent and this seems to be valid sponding to the end-of-primary (EOP) (i.e., when the excess
not only for sensitive clays but for all clays. In fact, as indi- pore pressure becomes approximately equal to zero) compres-
cated by L. Suklje (personal communication, 1984), the geo- sion curve would be
logical preconsolidation pressure is unique and constant; on
the contrary, the yield stress separating small and large strains,
which does not necessarily represent the geological maximum
CAN. GEOTECH. 1. VOL. 25, 1988

1.3
.,
; +Go' '

1.2

-,-
1.1
0
U)
'0

YG 1.0 0
Y

-
0
.,O 0.9 u

\
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

.,,
.a
/

0.8
0-
0.7 ---.------c
0.6
10-a 10- lo-6 lo-5
2",s-'
Normalized preconsolidation pressure - strain rate relationship for Champlain sea clays (from Leroueil et al.

cr;/ui0
0.8 1.0 1.5 2 . 0 2.5
0
For personal use only.

Increment 4 Incr. 5
2 k '

4
$"
>
\U 6
.-
C

e
G
8

10 -

I I 1
12 20
(Berre and lversen 1972) ( Samson et 01. 1981 )
(a) Drammen c/uy (b1 Jo/ieffe c/ay
FIG.4. Multiple-stage loading (MSL) tests.

In contrast to the strain rate at 24 h (assuming that the speci- eter test, it is possible to characterize two different phases: pri-
men is in the secondary compression range), the strain rate at mary and secondary. Various definitions have been given for
the end of primary is highly variable, depending on the charac- these two phases (Bjermm 1967; Leonards 1977; Crawford
teristics of the clay specimen, particularly its permeability. 1985). The author agrees with Crawford, who says that
"during the so-called primary phase, the permeability of the
Practical implications of strain rate effects
soil is the predominant factor controlling the rate of volume
As shown be Leroueil et al. (1985b), the consolidation change and during the secondary phase, it is the resistance of
behaviour of clay can be entirely described by a relatively the structure that determines the rate." Of course, during the
simple effective stress - strain - strain rate relationship secondary compression phase, settlement occurs and thus
(Fig. 2). This has important implications on the response of water flows out of the specimen. Consequently, from Darcy's
the soil during testing. law, there must -be a gradient and some excess pore pressures,
- - -
Stage loading tests but these are very small under usual laboratory conditions.
In stage loading tests, such as the conventional 24-h oedom- Furthermore, as indicated by Leroueil et al. (1985b), the soil
LERI

Effective stress ci , kPa Effective stress 6,


kPo
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

FIG.6. Observed stress-strain relations in various sublayers of a


sample of clay from Berthierville (from Leroueil et al. 1986, after
Mesri and Feng 1986).
For personal use only.

FIG.5. Schematic state path followed during a multiple-stage load-


ing oedometer test.

has only one rheological behaviour independently of the fact


thzt the clay could be in the primary or secondary phase. In
fact, most of the models (for example, Poorooshasb et al.
1981; Oka 1981) do not separate primary and secondary com-
pressions.
During secondary consolidation, i.e., at an essentially con-
stant effective stress, the strain rate decreases with time
(Fig. 2) as the void ratio (or strain) decreases with time in the
C,, concept.
The behaviour during primary consolidation is more difficult
to analyse, as the effective stress is continuously changing. An
idea of the effective stress-strain curve followed can be
obtained by measuring the pore pressure at the base of the spe-
cimen and deducting an average effective stress at various
times and strains. This has been canied out by Berre and Qii from conventional tests, kPa
Iversen (1972), Rizkallah et al. (1977), and Samson et al.
(1981). Figure 4 shows typical results. The stress -strain FIG.7. Correlation between the preconsolidation pressures obtained
curves followed are very different from the conventional com- from CRS and conventional 24-h oedometer tests (from Leroueil
pression curve obtained by simply joining the points corre- et al. 19836).
sponding to the end of the loading periods. In fact, at each
step, at the beginning of the loading period, the pore pressure sures and deformations at various depths in the specimen.
and the strain rate are high, and from the 04 - E , - t , model Berre and Iversen (1972) and Mesri and Feng (1986) have
shown in Fig. 2, the clay goes to high stresses (point D in solved the problem by putting subspecimens in series. Figure 6
Fig. 5); with time, the excess pore pressure and the strain rate presents results obtained by Mesri and Feng (1986) in an iso-
decrease, and the stress-strain state moves down along lines tropic consolidation test where four 12.7 cm high triaxial spe-
of lower strain rates. As previously mentioned, after 24 h the cimens were used; the step under consideration was entirely in
strain rate is approximately equal to lop7 s-I (point F in the normally consolidated range. The stress-strain curves fol-
Fig. 5). The complete stress-strain curve followed is a suc- lowed by the four subspecimens are very different and this can
cession of steps and not the smooth conventional compression be explained as follows: at the beginning of the loading period,
curve (Fig. 4). the strain rate is higher near the drainage boundary than near
To know what happens within the specimen during these the impervious boundary and consequently, in agreement with
successive steps, it becomes necessary to measure pore pres- the a: - E , - t , model, the clay near the drainage boundary (ele-
90 CAN. GEOTECH. J . VOL. 25, 1988

HVPOTHESIS A: E O P # Settlement gauge


Piezometer

A N D 0 ( T H I N SAMPLE)

.. .
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

8 ( T H I C K SAMPLE I100 31 THIN)

ON S A M R E T H I C K N E S S

(b) S T R A I N V S . T I M E C o n OCR=1 SAMPLES H A V I N G EQUAL


INITIAL CONDITIONS A N 0 AfTV

FIG.9. Determination of in situ stress-strain curves from settle-


ment and pore pressure observations.

In general terms, this means that each testing technique leads


to different results, depending on the strain rate history fol-
For personal use only.

lowed in that test. As all our experience has been established


on the basis of the conventional test, it is absolutely necessaw
to calibrate the results of the new tests (CRS, CGT, or others)
log TIME- with those of the conventional test before using them for
design.
FIG.8. Illustration of hypotheses A and B in terms of (a) strain vs.
stress and (b) strain vs. time (from Jamiolkowski et al. 1985). Compressibility in situ
Hypothesis A and hypothesis B
ment 1 in the figure) mobilizes higher effective stresses than The most important practical question for the soil engineer is
the clay element 4 near the impervious boundary. to know what is the relevant compression curve for in situ con-
ditions. As indicated by Ladd et al. (1977) and more recently
Continuous oedometer tests by Jamiolkowski et al. (1985), there are two extreme possibil-
In a constant rate of strain test the stress-strain curve ities (Fig. 8). "Hypothesis A assumes that creep occurs only
depends on the strain rate used. As the strain rates generally after the end of primary consolidation . . . " and consequently
used in practice are chosen between and 5 x s-I, that the stress-strain curve followed in situ is the same as the
the stress -strain curve obtained in a CRS test is different from one obtained in the laboratory at the end of primary (EOP
the one obtained in a conventional 24-h oedometer test, which curve). "Hypothesis B assumes that some sort of 'structural
was seen to correspond to strain rates of about lo-' s-I ([8]). viscosity' is responsible for creep, that this phenomenon
Data compiled by Leroueil et al. (19836) for eastern Canada occurs during pore pressure dissipation, and therefore that the
clays show that the preconsolidation pressure obtained in CRS strain at the end of primary consolidation increases" with
tests (6, = 4 x lop6 s-') is on average 28% higher than that sample thickness.
obtained in conventional tests (Fig. 7). Holtz et al. (1986) Numerous investigators have tried to demonstrate with lab-
have found similar results on the very stiff Montalto di Castro oratory tests which one of the hypothesis, A or B, is the most
clay. It thus follows that to obtain a "conventional curve" representative of clay behaviour. A review of the literature
from a CRS test camed out at a strain rate of 4 x s-I, the made by Mesri and Choi (19856) indicated that the answer is
effective stresses have to be divided by 1.28. To take into not clear-some results supporting hypothesis A, others sup-
account these strain rate effects, Larsson and Sallfors (1985) porting hypothesis B, and others (Aboshi 1973) giving results
suggest a graphical correction for CRS tests that also gives cor- in between. In the same paper, Mesri and Choi present end-of-
rected curves similar to 24-h curves. primary compression curves obtained in isotropic consolida-
In controlled gradient tests (CGT), the experience on eastern tion tests on specimens with heights varying between 2.5 and
Canada clays shows that the strain rate decreases when the 50 cm, and conclude that the EOP curve is unique. While their
strain increases in the normally consolidated range (Samson results seem conclusive, Leroueil et al. (1985~)commented
et al. 1981). It follows that the CGT oedometer curve in the that the C,, values are much smaller in isotropic conditions
normally consolidated range is steeper (C, higher) than the than in one-dimensional conditions and that the void ratio dif-
curves obtained in CRS tests (Kabbaj et al. 1985) or in conven- ference between hypotheses A and B at the end of primary
tional tests (Shields 1974). could be difficult to quantify. In the author's opinion, the rela-
LEROUEIL
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13
For personal use only.

FIG. 10. Comparison of in siru stress-strain curves and laboratory end-of-primary compression curves (after Kabbaj et al. 1986).

tive validity of hypotheses A and B has not yet been definitely curves obtained as indicated above and in Fig. 9, on four dif-
established experimentally in the laboratory. ferent sites with the stress-strain curves obtained at the end-
An approach frequently used to evaluate the representativ- of-primary consolidation in laboratory multiple-stage loading
ness of laboratory compression curves consists in comparing oedometer tests (Fig. 10). For this comparison and in general
computed and measured surface settlements of embankment in this study, the strain E, =AHIHo has been considered rather
foundations. However, as indicated by Leroueil and Tavenas than the void ratio because it erases the small initial void ratio
(1981) and Magnan et al. (1983), in such a problem there are differences unavoidably existing in natural soil deposits. On
too many parameters varying with depth and time to be able to the Berthierville site, the diameter of the embankment is nine
reliably evaluate one particular parameter. times the thickness of the clay layer; in Saint-Alban (embank-
A much better approach has been proposed by Pelletier et al. ment D), the maximum lateral displacement is about 2% of the
(1979). As schematized in Fig. 9, it consists in using deep surface settlement; on the Vasby site, Chang (1981) compared
settlement and pore pressure records to determine the stress - measured settlements with settlements computed on the basis
strain curves followed in situ. Homogeneous sublayers, well of water content changes, and found good agreement. It can
defined between two deep settlement gauges, are considered; thus be considered that the compression was essentially one
the settlements measured at the two settlement gauges give the dimensional in these three cases. A these three sites, samples
strain and strain rate, and the measured pore pressures give the were taken with the 200 mm diameter Laval sampler (La
effective stress, allowing a complete description of the in situ Rochelle et al. 1981) and are thus considered to be of an exel-
stress -strain curve. lent quality. From Fig. 10, it can first be observed that in Saint-
Alban and Berthierville, the effective stress decreases slightly,
Is the laboratory end-ofprimary curve representative of as the stress - strain curve has just passed the preconsolidation
in situ compressibility? pressure. This condition corresponds to an increase of the pore
Kabbaj et al. (1988) have compared in situ stress-strain pressure under constant applied load. Such a behaviour could
92 CAN. GEOTECII. J. VOL. 25, 1988
2", 5-1

Expected excess
pore pressure
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

pore pressure

FIG. 11. Compared stress states both in situ and in the laboratory.

be viewed as unexpected. However, Crooks et al. (1984) have


reported similar observations under 11 embankments out of 30
cases studied. Most of the investigators explain this behaviour
by breakdown of the clay structure or by pore pressures
induced by creep higher than the rate of decrease allowed by
Darcy's law. ~ h e s phenomena
e are generally considered to be CRS M S L ~ ~n sit"

associated with very sensitive materials. However, among the 0 CREEP @% MSLZ4 e. = 10%
For personal use only.

11 cases reported by Crooks et al. (1984), some are located in FIG.12. Effective stress - strain rate relations observed in the lab-
Great Britain, Northern Ireland, Thailand, and Trinidad, oratory and in situ at a vertical strain of 10% (after Leroueil et al.
places where the clays are not very sensitive. 19886).
Also evidenced in Fig. 10 is the major difference between
the EOP laboratory and in situ curves; at a given effective
stress, the in situ strain is much higher than the strain expected when the effective stress exceeds the preconsolidation pressure
from laboratory tests. Moreover, as indicated by the arrows, or due to creep phenomena; in such cases, pore pressures will
the strain on the EOP laboratory curve under the final stress is continue to decrease and settlements to increase, so that the
in all cases smaller than the existing strain under the embank- good agreement between observed and computed settlements
ments, where the primary consolidation is still not completed. reported by the authors could no longer be valid for long
This is particularly true for the Vasby case where the present periods of observation. Another possible explanation for high
effective stress increase is only 50% of the final stress increase. pore pressures would be a rise of the water table in the
There is certainly some uncertainty in the determination of the embankment due to the expelled water from the clay founda-
effective stresses in situ, but never to explain the typical differ- tion and (or) to rainfall; such a water table rise has been
ence of 24 kPa between the laboratory and in situ curves observed several times in embankments of sand or till.
observed in Fig. 10. The laboratory end-of-primary curves are Whatever the explanation for these high pore pressures, it
therefore not representative of the in situ behaviour and hypo- remains that at the settlement or strain considered, it implies
thesis A is not valid. that the effective stress is smaller than expected. As shown in
To the author's knowledge, there is no other comparison of Fig. 11 the predicted pore pressure is given by the distance FL
in situ and laboratory stress-strain curves in the literature. between the final effective stress and the considered stress-
However, there are some cases of embankments in which strain curve (in these cases, laboratory end-of-primary or 24-h
measured and computed behaviour in terms of settlements and curves). If the pore pressure (FS in Fig. 11) is, as generally
pore pressures have been compared and analysed: Avonmouth observed, higher than the calculated one, it is because the soil
(Lewis et al. 1976); Aiko (Shoji and Matsumoto 1976); is following in situ a stress-strain curve such as S'S" below
Penang (Adachi and Todo 1979); Saint-Alban-D (Leroueil and the laboratory one. This is also confirmed by Larsson (1986):
Tavenas 1981); MIT-195 (SouliC and Silvestri 1984); Cubzac- after studying the cases of Vasby, Ska-Edeby, and Drammen,
les-Ponts, test fill B (Magnan et al. 1983; SouliC and Silvestri he concluded that "creep deformations occur within the pro-
1984); Vasby (Mesri and Choi 198%); New York Port (Aron- cess of pore pressure dissipation and afterwards, and the com-
witz 1986). pressibility is thus time dependent"; then, refemng to "several
In all these cases, it was possible to find a relatively good field observations of settlements made during recent years in
agreement between observed and computed settlements (most connection with road constructions in Sweden," he writes,
of the time by modifying the value of the permeability coeffi- "no final settlements have been measured but the general
cients), but the measured excess pore pressures were higher observation is that the settlements that have occurred during
than the computed ones in all cases except Aiko test embank- the times for observation have been faster and larger than what
ment. Investigators generally consider that these high excess was predicted from calculations where creep effects were dis-
pore pressures are due to a breakdown of the soil structure regarded." Ail these data confirm that creep effects, with dis-
LEROU EIL
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

FIG. 14. Effective stress paths, laboratory and in situ.

M = da:/dc,) in the overconsolidated range also show differ-


ences between laboratory and in situ behaviours: the moduli
observed in situ at Berthierville, Saint-Alban, and Gloucester
(Fig. 10) are higher than the ones measured in the laboratory.
I 0 CREEP

IS8 MSL,

MSL,,
in situ
Similar behaviours have been observed by other researchers:
Crawford and Bum (1962) measured settlements of a mat foun-
FIG. 13. Preconsolidation pressure - strain rate relations observed dation in Toronto, equal to only 1/10 of the estimated value
in the laboratory and in situ (after Leroueil et al. 19886). from recompression in oedometer tests; at Kars Bridge, Folkes
and Crooks (1985) observed an almost rigid behaviour of the
placement of the stress-strain curve followed, exist in situ clay in the overconsolidated range and thus a modulus much
during primary consolidation, and that the laboratory EOP higher than the one obtained in the laboratory by Eden and
stress-strain curve is not representative of the in situ clay Poorooshasb (1968).
Several explanations can be given to explain such discrep-
For personal use only.

behaviour.
ancies between laboratory and in situ behaviours at small
Is the laboratory 4 -c, - i, model relevant in situ? strains: the first one could be that the a{-c,-i, concept is not
The a{ - c, - i, model described in a previous section is, in general; the second could be that the clay specimens, as they
fact, a particular case of type-B theories and it is possible to are in the oedometer cell, are disturbed; the third one could be
evaluate its validity by comparing the effective stress - strain associated with stress path effects. These possible explanations
rate relation at a given strain deduced from laboratory oedom- will now be examined.
eter tests with the effective stress - strain rate in situ state at Stress paths followed during oedometer tests have been
the same strain. This has been done at the sites of Berthierville, investigated by Sallfors (1975), Lefebvre and Philibert (1979),
Saint-Alban, and Vasby where numerous, some very long Silvestri (1981), and Mesri and Castro (1987); they show small
term, oedometer tests have been carried out (Kabbaj 1985; lateral effective stress increases in the overconsolidated range
Leroueil et al. 19886). and a;/a; ratios approximately equal to (1 - sin 4 ' ) in the
Figure 12 shows a:-& relations obtained at a large strain of normally consolidated range (Fig. 14). The stress paths fol-
10%. It can be seen that the agreement between laboratory and lowed in situ under embankments have been studied by
in situ behaviours is excellent, which indicates that the Leroueil et al. (1978), Folkes and Crooks (1985), and Leroueil
a{ - E, - i, model established in laboratory conditions is also and Tavenas (1986); they are variable depending on the con-
valid in situ. This clearly proves that there is an accumulation solidation properties of the clay in the overconsolidated range,
of creep deformztions during primary consolidation. However, the thickness of the compressible foundation, the rate of con-
before recommending the use of numerical program based on a struction, and the geometry of the embankment. However, in
a{-E,-i, model to compute settlements, it is necessary to all cases, the in situ stress p~.thsare different from the ones
verify the practical validity of such a model at other strains, existing in oedometer tests (Fig. 14) and could explain a differ-
and in particular at the strain corresponding to the preconsoli- ence in behaviour.
dation pressure. As indicated by the high in situ moduli in the overconsoli-
Comparing the preconsolidation pressure states in the labor- dated range, another possible factor is disturbance. In the cases
atory and in situ for the site of Gloucester, Leroueil et al. of Berthierville and Saint-Alban, the samples were taken with
(1983~)indicated that the in situ a;-i, state "appeared to lie the Lava1 sampler of 200 mm diameter. This sampler has been
within the extrapolated boundaries defined in laboratory shown to give samples of quality equivalent to blocks (La
tests"; however, there was one and one-half log-cycle separat- Rochelle et al. 1981); however, some disturbance is always
ing laboratory and field data, so that no definite answer could possible, although not otherwise evidenced. Disturbance could
be given at that time. Figure 13 shows the a;-i, relations for also come from the trimming and the mounting of the speci-
the Saint-Alban and Berthierville clays. In contrast to the men in the oedometer ring. Burland and Symes (1982) have
previous figure, it clearly appears in this case that the in situ shown that the overall deformation measured in the triaxial cell
preconsolidation pressure is higher than the laboratory precon- overestimates the deformation in the central part of the speci-
solidation pressure measured at the same strain rate. It thus men; this could also be true for oedometer tests.
seems that the a{ - E , - 6, relationship deduced from oedometer So, when examining the representativeness of the a{ - E, - i,
tests is not representative of the in situ behaviour at a;. model, it appears that at large strains, when the stress paths are
Observations of the deformations (or deformation moduli about the same ix situ and in the laboratory and the soils in situ
CAN. GEOTECH. J. VOL. 25, 1988
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

1.2cOCRc2.5
2.5< OCR A
For personal use only.

ub ,kPa
FIG. 15. Correlation between in situ and laboratory conventional 24-h preconsolidation pressures.

and in the laboratory are destructured in the same manner, the


a{ - E, - iv model applies. On the other hand, at small strains,
where the stress paths are different in the laboratory and in situ
and where the clay behaviour in the laboratory is the most
affected by possible disturbances, the a{ - E, - iv model estab-
lished in laboratory is not representative of the in situ behav-
iour. In practical terms, that means that numerical
consolidation programs based on a a{ - 6, -4 model cannot be
used at the present time to compute settlements and settlement
rates.
Search for a representative stress -strain curve
It has been previously seen that the end-of-primary consoli-
dation laboratory a: -E, curve and the a$- E, - 6, model estab-
lished from laboratory tests are not representative of the overall
in situ behaviour. Thus, it becomes necessary to define a repre-
sentative stress -strain curve in a semiemperical manner.
Sallfors (1975), Leroueil et al. (1978), Morin et al. (1983),
and Crooks et al. (1984) have used different techniques to esti-
mate the preconsolidation pressure mobilized under embank-
ments: determination of stress-strain curves as shown in
Fig. 9; analysis of of pore pressures during and after construc-
tion; comparison of water content profiles-before construction FIG.16. Comparison between in situ and conventional 24-h labora-
and during the process of consolidation; observation of settle- tory stress-strain curves for eastern Canada clays.
ments of embankments of different heights. Their results are
summarized in Fig. 15 where in situ preconsolidation pressures
are compared with preconsolidation pressures measured in OCR's, it overestimates that pressure. Morin et al. (1983) pro-
conventional 24-h oedometer tests. For overconsolidation posed a small correction to the laboratory values:
ratios (OCR) between 1.2 and 2.5, both values are approxi-
mately equal. At lower OCR's, the conventional test underesti- [lo] 0; in situ = a10;cconv
mates the in situ preconsolidation pressure, whereas at high whereal = 1.1 for OCR < 1.2; al = 1.0 for 1.2 < OCR <
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

FIG. 17. Estimation of the end-of-primary in situ strain in excess of that determined from conventional 24-h laboratory test (from Leroueil
et al. 1988b).

2.5; and a l = 0.9 for 2.5 < OCR < 4.5. i, = m A log a:. Experimental data give m = 32.
For personal use only.

Larsson and Sdlfors (1985) have compared a; values The proposed equation for AE, is
obtained in CRS tests and corrected for strain rate effkcts with
in situ preconsolidation pressures, and have also found good
agreement. It thus appears that in situ a; is approximately
equal to the conventional 24-h preconsolidation pressure, this
value being obtained directly from standard stage loading tests +
A g is given in Fig. 17 as a function of Cc/(l eo) and of the
or from CRS tests with a correction for strain rate effects. This approximate strain rate at the end of primary iEOP.Equation
correction could be done with a multiplying factor, as sug- [ll.] was applied by Leroueil et al. (19886) to the cases of
gested by Leroueil et al. (1983b), or with a graphical method, Berthierville, Saint-Alban-D, and Vasby and A q values found
as suggested by Larsson and Sallfors (1985). are respectively 2.8, 4, and 8 % , implying additional settle-
From all these data (Figs. 12 and 15), it becomes possible to ments respectively in the order of 9, 20, and 110 cm. These
compare the in situ compressibility curve with the compression values are realistic when it is considered that at Berthierville
curve obtained in conventional (24-h) oedometer tests or corre- and Saint-Alban, the strains computed on the basis of the con-
sponding to a strain rate of s-l. The conclusions can ventional oedometer test, respectively 10 and 13.5%, have
be summarized as follows (Fig. 16): in the overconsolidated been exceeded when the pore pressure in the middle of the clay
range, the modulus M is higher in situ by a factor that will vary layer was still 25 and 35% of the excess pore pressure at the
with the sampling technique used, the trimming technique, and end of construction. At Vasby, total settlements have not been
probably with the soil; as discussed earlier, the preconsolida- calculated; however, in the clay layer between 4.3 and 7.3 m
tion pressure is about the same; in the normally consolidated (Fig. lo), the strain estimated from conventional oedometer
range, the in situ strain is larger than the laboratory strain. tests (about 15%) has been exceeded when the excess pore
Leroueil et al. (1988b) have estimated the strain AE, (Fig. 16) pressure was more than 50% of the applied load. It is worth
that can be expected at the end-of-primary consolidation in situ, noting, however, that in clays of very low sensitivity having
in excess of the strain estimated on the basis of the conven- +
Cc/(l eo) values of about 0.25, A q would be much smaller
tional (24-h) oedometer test. Their assumptions can be sum- and hardly in excess of 3 % .
marized as follows: The calculated strains and A q (Figs. 16 and 17), which
-the end of primary is reached when the excess pore pressure depend on the reference conventional oedometer curve, could,
in the middle of the clay layer is 8% of the initial excess pore however, vary with the sampling technique used. As indicated
pressure uo; this value is deducted from the Terzaghi theory in by Bozozuk (1971), La Rochelle and Lefebvre (1971), Lacasse
which it corresponds to a degree of consolidation equal to 95 %; et al. (1985), and Holtz et al. (1986) (Fig. 18), the stress-
-during the last phases of primary consolidation, the pore strain curve can be strongly affected by sampling disturbance
pressure isochrone is parabolic, which implies, together with and the comparison of in situ versus laboratory behaviours is
the previous assumption, that the strain rate at the end of pri- modified accordingly. An example of this problem is well
mary is iEoP = (0.16ku,!,)l(yWfP) (H is the drainage length); illustrated in the case of Vasby. Mesri and Choi (1985a), con-
-at large strains and for strain rates smaller than s-l, the sidering EOP oedometer test results obtained on samples taken
strain rate dependence of the effective stress at a given strain in 1967, came to a relatively good agreement between com-
can be described by a linear log a{ - log i, relation: A log puted and measured settlements. On the other hand, Leroueil
CAN. GEOTECH. J. VOL. 25, I988

log (TI
,kPa
lo2 lo3
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

-- (Holtz et ol. 1986)

Gloucester cloy Montolto di Cosfro cloy


FIG. 18. Sampling disturbance effect on the compression curves of natural clays.

and Kabbaj (1987), using test results obtained on samples


taken with the Laval 200 mm diameter sampler, found in situ
strains much higher than the ones obtained in the laboratory
under the same effective stress (Fig. 10). The reason for such a
discrepancy between the Mesri and Choi (1985~)and Leroueil
and Kabbaj (1987) results is clearly shown in Fig. 19, where
For personal use only.

typical oedometer curves are shown. The specimens taken with


the Laval sampler have a well-defined knee at the preconsoli-
dation pressure; on the contrary, the other curve is much more
rounded, indicating a high degree of disturbance and present-
ing larger strain under a given effective stress. It is thought that
in the study by Mesri and Choi (1985~)strain rate effects have
been balanced by disturbance effects.

Permeability
Validity of Darcy's law
While Darcy's law is very simple and presented in all text
books, its validity is still debated. Some authors (Hansbo
1960; Law and Lee 1981) suggest nonlinear gradient -velocity
relationships, whereas others (Miller and Low 1963) suggest
the existence of a threshold gradient below which no flow
occurs. When reviewing and analysing published data, Olsen
(1985) concluded that there is a linear relation between the
flow and the hydraulic gradient. However, owing to chemical
osmosis and electrical osmosis gradients that could be superim- 10'
Vertical effective stress 0-;,kPa
posed on the hydraulic gradient, the relation between the flow
and the hydraulic gradient can present an intercept. FIG. 19. Typical compression curves for the Vasby clay (after
Leroueil and Kabbaj 1987).
It is thought that in most natural intact clays, chemical or
electrical gradients are small and that Darcy's law could be
considered valid. Indeed, Tavenas et al. (1983b) confirm this formed with a Mariotte bottle (Olson and Daniel 1981) or with
validity in the laboratory for gradients smaller than 0.07, and mercury pots.
Leroueil et al. (1983~)observed at Gloucester some settle- Permeability tests along the vertical and horizontal directions
ments-thus flow of water-under average gradients of 0.1. on intact specimens at their natural void ratio have shown that
the anisotropy is very small in massive marine clays (rk =
Laboratory permeability tests kholkvo = 1.1) (Larsson 1981; Tavenas et al. 1983~).For
Constant-head and falling-head permeability tests can be strongly overconsolidated clays, stratified materials, or varved
performed in the laboratory in triaxial or oedometer cells; for clays, rk is larger. Chan and Kenney (1973) found values
eastern Canada clays, both tests lead to the same k values, but between 3 and 5 for a varved clay from New Liskeard; Table 1
for soils having higher swelling index C, the change in effec- gives possible ranges of rk values (from Jamiolkowski et al.
tive stress during the falling-head test could introduce volume 1985).
changes, which could alter the test results. In general, it is thus The laboratory tests are the only ones that can give the varia-
preferable to use constant-head tests, which can easily be per- tion of the permeability with compression. They show a
TABLE1. Range of possible field values of the ratio k,lk, for soft clays
(from Jamiolkowski et al. 1985)

Nature of clay kdk~


No macrofabric, or only slightly developed macrofabric,
essentially homogeneous deposits 1- 1.5
From fairly well to well-developed macrofabric, e.g.,
sedimentary clays with discontinuous lenses and layers
of more permeable material 2- 4
Varved clays and other deposits containing embedded and
more or less continuous permeable layers 3-15
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

decrease of the permeability according to relationships that can


be approximated, for all practical purposes, by a linear e -
log k curve for strains smaller than 20 % . The slope Ck = AelA
log k of these curves is approximately equal to 0.5eo (Tavenas
et al. 1 9 8 3 ~ ) . I I I I I I
Finally, the coefficients of permeability obtained in the lab- 2 4 6 8 10 12 14
oratory have to be corrected for temperature effects before P/d
being used in designs. The reduction factor to be applied, FIG. 20. Shape factor for cylindrical probe (from Tavenas et al.
when temperatures in the laboratory and in situ are respectively 19860).
20 and 7 " C , is typically equal to 1.5.
In situ permeability tests Constant-head tests can be interpreted with the following
Testing techniques equation:
Permeability measurements are sometimes made in cavities
formed in a borehole between two impervious bentonite plugs.
For personal use only.

This method is, however, uncertain, since plugs are often not
reliable, the geometry of the cavity not well known, and the
soil around the cavity is often affected by the boring process in which q , is the measured stabilized flow.
and possible stress release. Most often, in situ permeability Willunson (1968) and Mieussens and Ducasse (1977) sug-
tests are performed in piezometer probes pushed into position. gested that for isotropic materials and for porous elements
However, this technique presents the following shortcomings having a length-to-diameter ratio equal or greater than 2 , the
(Tavenas et al. 1 9 8 3 ~ ) . flow could be considered as a pure radial flow with an error of
-during the installation, any layering of the natural soil is dis- less than 3 % . However, Randolph and Booker (1982) indi-
torted and more or less erased; cated that this statement is not valid, as confirmed numerically
-the clay around the probe is remoulded by the installation and experimentally by Tavenas et al. ( 1 9 8 6 ~ ) In . fact, as
and then reconsolidates with time to a void ratio smaller than shown by these authors, who have performed constant- and
the natural one; falling-head tests in piezometers and self-boring permeameters
-the porous element of most of the piezometers is subjected to with length-to-diameter ratios varying between 1 and 14.4,
clogging. both [12] and [13] lead to the same in situ k value, indepen-
All these phenomena lead to a reduction of the measured coef- dently of the probe geometry, when the "recommended rela-
ficient of permeability. tion" shown in Fig. 20 is used to choose the shape factor F.
To overcome these problems, Baguelin et al. (1974) and Comparison of test results
Mieussens and Ducasse (1977) have proposed a self-boring To quantify the relative effects of remoulding and clogging
permeameter to practically eliminate the disturbance of the soil associated with the use of piezometer probes, Tavenas et al.
and its reconsolidation. However, this apparatus was equipped (1986b) have compared test results obtained in a pushed-in-
with a fine-grained porous element subject to clogging. Roc- place piezometer, in a nonclogging self-boring permeameter,
test Ltd. has since developed another self-boring permeameter and in the same permeameter equipped with a closed conical tip
in which the porous element consists in a series of -9 mm and pushed into position. In this last technique, there was some
holes that eliminates clogging. During the placement of remoulding but no clogging. The results (Fig. 21) show that
the probe, the holes are closed with an inflated membrane the remoulding affects the measured permeability by factors
(Tavenas et al. 1983c, 19866). varying between 1 . I at Saint-Polycarpe, where the clay deposit
As in the laboratory, two types of permeability tests can be is very homogeneous, and 9.1 at Saint-Alban, where the
performed in situ: variable-head tests with a graduated burette deposit is more stratified. Overall, clogging and remoulding
and constant-head tests with a Mariotte bottle. Variable-head were found to reduce the coefficient of permeability by a factor
tests are commonly interpreted by the Hvorslev (1951) of 2 - 17, which proves that the tests in piezometers underesti-
approach: mate the in situ permeability.
Figures 22 and 23, give the permeability profiles for the
Louiseville and Saint-Alban sites, which are respectively
exceptionally homogeneous and moderately stratified; in situ
in which A is the section of the tube, H I and H2 the hydraulic permeabilities have been determined with [12]and [13]and the
heads at times tl and t2, and F a shape factor, given in Fig. 20. shape factor proposed by Tavenas et al. ( 1 9 8 6 ~(Fig.
) 20). The
CAN. GEOTECH. J. VOL. 25, 1988

I U
2 3 4 5 6 7 8 9 ' " 2 3 4 5 6
F-------- Saint- Polycorpe

I I
Louiseville
Plezorneter
I I
Berthierville

II - - . I I I I I I
0
I I
Saint-Alban

I Effectof ?loggin{'< r e m o l d i n g 4
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

FIG.21. Results of permeability tests with different probes illustrating the effects of remoulding and clogging (after Tavenas et al. 1986b).

K.m/s
For personal use only.

FIG. 22. Permeability profile at Louiseville (after Tavenas et al. 1986b, modified with method proposed by Tavenas et al. 1986a).

following remarks can be made: between 0.6 and 1.7 with an average value of 1.2. This ratio is
-the permeability coefficients obtained in constant-head tests approximately equal to the ratio of the horizontal and vertical
and variable-head tests have the same value on both sites and at permeabilities measured in the laboratory (Tavenas et al.
all depths, thus confirming the validity of the Tavenas et al. 1 9 8 3 ~ )The
. permeability measured in the self-boring perme-
( 1 9 8 6 ~approach;
) ameter being influenced by the horizontal permeability, it can
-the permeability obtained with pushed-in piezometers is be concluded that the permeability measured in the laboratory
smaller than the one measured with the self-boring permeam- is representative of the in situ permeability of homogeneous
eter; as previously mentioned, the effect of clogging and clays;
remoulding is much more important in stratified deposits than -on the Saint-Alban site (Fig. 23), as on the two other strati-
in homogeneous deposits; fied sites studied by Tavenas et al. (1986b),the laboratory per-
-on the Louiseville site, as on the three other homogeneous meability underestimates the in situ permeability.
sites studied by Tavenas et al. (1986b), the permeability coef- Hence, for general soil conditions, the self-boring permeam-
ficient measured in the self-boring permeameter is very similar eter seems to be the best test to evaluate the in situ permeabil-
to the one measured directly in the laboratory on intact clay at ity. However, it must be mentioned that the permeability
its natural void ratio. The ratio k,,,,lklab was found to vary measured with this apparatus is, in fact, a function of the hori-
LEROUEIL
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

FIG.23. Permeability profile at Saint-Alban (after Tavenas et al. I 9866, modified with method proposed by Tavenas et al. 1986~).
For personal use only.

zontal and vertical in situ permeabilities. At the present time, it (1988~);Fig. 25 shows the results obtained at Berthierville, in
is not possible to separate the two values but, by using self- the central part of the clay layer where the soil is very homo-
boring permeameters of various geometries and developing geneous. The diameter of the embankment being nine times
appropriate theories of interpretation, both horizontal and ver- the thickness of the clay layer, the conditions are excellent to
tical permeabilities could certainly be defined. Research on validate the approach and to verify the representativeness of
this subject is presently in progress at UniversitC Laval. the permeability coefficients measured with various tech-
niques. Also shown in Fig. 25 is the range of the directly meas-
Direct determination of in situ permeability ured permeability coefficient - strain relations obtained from
In the section on compressibility, it has been seen that a well- laboratory tests on four different specimens and also the in situ
instrumented test embankment can be a most effective means permeability measured in pushed-in piezometers and with the
to determine one of the basic soil parameters, i.e., the stress- self-boring permeameter, at a vertical strain of zero. The fol-
strain curve (Figs. 9 and 10); such a test can also be used to lowing comments can be made:
determine the real value of the permeability at various points in -the results deduced from [15] are slightly dispersed; how-
the foundation, and at various times and strains. ever, on the average, the permeability decreases with the strain
During the consolidation of a clay layer, at a given time ti, (or void ratio) along a line parallel to the k - E , curve obtained
the excess pore pressure (u') distribution could be as given in in the laboratory;
Fig. 24a; at two times ti-, and ti+,, the settlement distribution -the "real" in situ permeability is slightly higher than the
could be as shown in Fig. 24b. From Darcy's law, there is no permeability measured in the laboratory; the average ratio is
water flow through depth ZN, as au'laz = 0 at that level. On about 1.6;
the other hand, at a depth such as ZA,there is a vertical flow -the permeability measured in pushed-in-place piezometers
rate, which is underestimates the "real" in situ permeability;
-the permeability measured with the self-boring permeameter
is very close to the "real" in situ permeability and can cer-
tainly be considered as representative;
If the consolidation is one dimensional and if the soil is satu- -all the results would have been different in more heterogen-
rated, the flow rate at point A is equal to the settlement rate eous deposits, the permeability being, as described by Rowe
between point N and point A. With the use of Fig. 24b, the (1972), strongly influenced by the stratification and the fabric
flow rate is calculated, and the permeability coefficient can be of the soils. It is thought, however, that the self-boring perme-
computed by using the following equation: ameter is an excellent tool to measure permeability in all
clayey deposits.
Consolidation
During the consolidation process of a saturated clay, com-
pressibility and permeability are intimately linked. In the very
simple case of Terzaghi's theory in which a linear stress-
This approach has been successfully tried by Leroueil et al. strain relation and a constant permeability are assumed, the
CAN. GEOTECH. J. VOL. 25, 1988

Settlement S Permeability k ,m/s


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

FIG. 24. Schematic excess pore pressure and settlement distribu-


tions for the determination of in situ permeability.

consolidation is controlled by the coefficient of consolidation


c, ([3]); the settlement - logarithm of time curve obtained
after an instantaneous loading has its typical S-shape and the
degree of consolidation expressed in terms of settlement is
equal to the degree of consolidation expressed in terms of pore
pressure.
When the compressibility curve is bilinear in an e - log a:
diagram, the coefficient of consolidation, as defined by [3],
varies with time and also with the position of the soil element
within the specimen. The coefficient c, is high in the overcon-
solidated range and much lower in the normally consolidated
For personal use only.

(from Leroueil et 01. 19860 )


range. As shown by Mesri and Rokhsar (1974) and by Tavenas
et al. (1979) (Fig. 26), the degree of consolidation expressed FIG. 25. Comparison between permeabilities deduced from pore
in terms of pore pressure presents, in these conditions, a step pressure and settlement observations and permeabilities measured in
corresponding to the passage of the preconsolidation pressure. the laboratory and in situ for Berthierville clay (Leroueil et al.
The degree of settlement shows a typical S-shape close to the 1988~).
Terzaghi curve associated with the final value of the coeffi-
cient of consolidation. Taylor square root method, it is generally smaller (Lambe and
From what has been said in the previous sections, the soil Whitman 1969; Pelletier et al. 1979; Olson 1985). From data
behaviour is even more complex, since the stress -strain curve reported in the literature and from the author's experience, the
followed by an element depends on its strain rate history and ratio varies between 0.2 and 1. Sridharan and Sreepada Rao
thus on numerous factors such as the permeability, the position (1981) have proposed the "rectangular hyperbola fitting
of the element in the compressible layer, and the thickness of method" to determine c,. From their data, it appears that the
the layer (Fig. 2 , 4 , 5, and 6). In such conditions, the value of values they obtain can also be different from the ones obtained
c, could become extremely variable with the position of the by the Casagrande and Taylor methods. This is not surprising,
element and with time, even if the considered step loading is since all these methods are based on Terzaghi's theory, assum-
entirely in the normally consolidated range. This parameter is, ing a constant c,, which is not the case in reality (Figs. 4-6).
however, very often used in practice to estimate overall rates The values of c, are sometimes computed using [3] with a
of settlement. It is the author's opinion that in these conditions directly measured coefficient of permeability and a modulus
it must be considered as a curve-fitting parameter rather than a obtained from an oedometer curve. These c, values have been
fundamental parameter. found larger than the ones obtained graphically (Leroueil et al.
Consolidation in the laboratory 1 9 8 5 ~Olson
; 1985). This difference can be explained, at least
Depending on the load increment ratio, loading duration, partly, by the stress -strain curve actually followed during
and position of the final stress relative to the preconsolidation consolidation. As shown on Figs. 4 and 5, the stress-strain
pressure, the Ah or E,, - log t curve obtained in an oedometer curve actually followed is such as ODEF (Fig. 5) and most of
test can have a typical S-shape or have a continuously increas- the strain accumulates along DE where the modulus M is much
ing slope (Leonards and Girault 1961; Bjermm 1967). These smaller than the average modulus between points 0 and F.
various transitions between the primary and secondary phases When this last modulus is used in [3], it leads to c, values that
are consistent with time or strain rate effects, as explained by are not representative of behaviour in the laboratory.
Mesri and Godlewski (1977) on the basis of the C,,/C, concept It is generally accepted that the time required to reach a cer-
and by Kabbaj et al. (1985) on the basis of a a{ - E,, - b model. tain degree of consolidation is proportional to the square of the
For most of the clays in the normally consolidated range and drainage length:
in conventional oedometer tests, the Ah - log t curve exhibits
a S-shape and it is possible to define a coefficient of consoli-
dation with the Casagrande log method (c, ,). When this
value is compared with the c,fi value obtained with the Several investigators have camed out oedometer tests on soil
LEROUEIL

Time, s
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

FIG.26. Results of CONMULT program for a stress increment spanning the preconsolidation pressure, and comparison with Terzaghi theory
(from Tavenas et al. 1979).

specimens of various heights or drainage lengths. Figure 27 to5 -


shows the ratio of the times required to reach 50% of primary
consolidation versus the square of the drainage lengths ratio for
the published cases. It can be seen that the time ratios are Intact Norwegian clay ( Berre 1969)
For personal use only.

systematically smaller than (Hlh)2. In other words, the coeffi-


cient of consolidation computed on the basis of test results
obtained on thick specimens is higher than the one deduced 0 Cubzoc- les-Ponts clay (Felix et al. 1981

from a test on thin specimens.


Kabbaj (1985) has simulated a loading on two specimens
having heights of 49.5 and 9.5 mm with a numerical program
taking into account strain rate effects (Kabbaj et al. 1985). He
obtained a consolidation time ratio of 17 instead of the theoret- -
c
,
ical value of 27, which indicates that the results shown in 2
+

Fig. 27 can probably be explained by strain rate effects. How-


ever, as the strain rate effects increase with strain rate (Fig. 3),
their influence on the consolidation time is probably more
important when comparing small specimens, for example, 1
and 5 cm high, than thicker specimens, for example, 10 and
50 cm high.
Consolidation in situ
When the thickness of the compressible layer is not large
compared with the embankment width and the lateral displace-
ments are small, an average coefficient of consolidation can be
computed on the basis of the surface settlement.
A first method consists in applying the graphical Casagrande ( H / h )2

or Taylor method to field conditions. A more attractive method FIG. 27. Results of consolidation tests on specimens of different
has been developed by Asaoka (Asaoka 1978; Magnan and thicknesses.
Deroy 1980). As indicated in Fig. 28, the method consists in
-drawing the settlement of the compressible layer versus the
time on a linear scale (Fig. 28a); where H is the maximum drainage length.
-choosing a time increment At to define the settlements si at A major advantage of the Asaoka method is that it allows an
+
times to iAt (i = 0 , 1, 2, . . . ); estimation of the final settlement and of the coefficient of con-
solidation during the consolidation process.
-plotting si versus siPl (Fig. 28b);
-drawing the line A through the points (si, s ~ - ~ ) . Average in situ c, values have been back-calculated on
The final settlement is theoretically obtained when si = si- numerous sites and compared with c, values obtained in the
at point F in Fig. 28b; the coefficient of consolidation can be laboratory. Table 2 shows such results and ratios between
obtained from the slope of the line A: in situ and laboratory coefficients of consolidation. It can be
seen that these ratios are highly variable, with values ranging
between 3 and 200. If the extremely high values obtained at
Melbourne and Penang are neglected, the average is 20. The c,
102 CAN. GEOTECH. 1. VOL. 25, 1988
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

(o 1 Settlement curve, fbl Asooko construction


FIG. 28. Analysis of settlements with Asaoka method.

TABLE2. Comparison between the coefficients of consolidation determined in the laboratory and
deduced from embankment settlement analysis

Cv in siru
For personal use only.

Cv lab Cv in siru
Site (m2/s) (m2/s) Cv lab Reference
Ska-Edeby IV Holtz and Broms (1972)
Oxford (1) Lewis et al. (1975)
Donnington Lewis et al. (1975)
Oxford (2) Lewis et al. (1975)
Avonmouth Lewis et al. (1975)
Tickton Lewis et al. (1975)
Over Causeway Lewis et al. (1975)
Melbourne Walker and Morgan (1977)
Penang Adachi and Todo (1979)
Cubzac B Magnan et al. (1983)
Cubzac C Leroueil et al. (198%)
A-64 Leroueil et al. (1985~)
Saint-Alban Leroueil et al. (1985~)
R-7 Leroueil et al. (1985~)
Matagami Leroueil et al. (1985~)
Berthierville Kabbaj (1985)
*c, estimated with Asoka method.

values obtained graphically in the laboratory are not truly rep- sentative of field conditions.
resentative of the in situ behaviour. Folkes and Crooks (1985) have suggested a different method
There is a general agreement to the effect that the small for the determination of the coefficient of consolidation based
laboratory specimens are not representative of the clay mass, on pore pressure observations. They consider a pore pressure
which is possibly heterogeneous and often contains thin layers distribution at a given date as an initial distribution (Fig. 29); a
of more pervious materials (Rowe 1972). This was particularly second profile at a later date is then compared with isochrones
true at Melbourne where Walker and Morgan identified silt and generated by finite difference techniques, and an average c,
sand seams. Simons (1975) recommends to calculate cv values value is estimated.
on the basis of permeability tests carried out in the laboratory Unfortunately, the data analysed by Folkes and Crooks do
on large specimens, or in situ. Other causes of error may be the not allow the use of the Asaoka method and both methods
neglect of the two- or three-dimensional aspects of the problem cannot be directly compared. It is worth noting, however, that
(a review of the main theories related to this has been made by in general they will give different results, since one is based on
Murray (1978)) and of the anisotropy of the clay. pore pressure observations and the other is based on settlement
Another possible factor is the scale effect previously observations. In cases approaching Terzaghi hypotheses, with
described (Fig. 27): if the cv parameter obtained in a small spe- a regular increase of the strain with effective stress (soil A in
cimen is not representative of the c, value obtained on a thicker Fig. I), both results should be similar; in soils presenting con-
specimen in laboratory conditions, it could hardly be repre- stant or increasing pore pressures after the end of construction,
sublayers well defined between deep settlement gauges. The
comparison with laboratory behaviour shows that
-the in situ stress-strain curves are very different from the
ones obtained in the laboratory at the end of primary consolida-
tion, which proves that hypothesis A is not correct;
-at large strains, there is a good agreement between in situ
behaviour and the a{ - E, - i, relationship established in the
laboratory. A method is proposed to calculate consolidation
strain in excess of the one predicted by the conventional 24-h
oedometer test;
-at small strains (overconsolidated range and preconsolida-
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

tion pressure), the in situ behaviour is different from the one


deduced from the u{-eV -6, model. Reasons for such a dis-
crepancy are not clear. However, an empirical a; Sill, - uiCcOnv
relation has been established.
One problem remains when the final load does not exceed
the estimated preconsolidation pressure. Could creep deforma-
tions be important in such a case and could the preconsolida-
tion pressure decrease with time? Laboratory test results
indicate that the answer is "yes," since 0; values lower than
the ones obtained in conventional oedometer tests have been
observed in long-term tests. But the representativeness of the
laboratory behaviour at small strains has already been ques-
tioned and such results cannot be considered as proof of the
time or strain rate dependence of the in situ preconsolidation
pressure. The only available detailed field data concerning this
FIG.29. Estimation of in situ c, value (from Folkes and Crooks point seem to be the long-term settlements of buildings in
For personal use only.

1985). Drammen (Norway) reported by Bjermm (1967) and rean-


alysed by Foss (1969) and Larsson (1986). They indicate that
noticeable creep deformations start at effective stresses equal
like soil B (Figs. 1 and lo), the Folkes and Crooks' method to about 80% of the estimated preconsolidation pressure. On
would give zero or a negative cv value, while Asaoka's the other hand, when Morin et al. (1983) compared in situ pre-
method, which is in fact a curve fitting of the settlement curve, consolidation pressure under embankments and conventional
would give a positive value. laboratory preconsolidation pressure, they came to the same
conclusion when a; was determined from field records taken
Conclusions and practical implications up to 16 years after construction and when it was established
The consolidation process is a combination of Darcy's law, from pore pressure generation during construction; this could
which controls the rate at which the water is expelled out of the indicate that the decrease of a; with time is small or not exis-
soil, and compressibility, which controls the evolution of tent in eastern Canada clays. In fact, there are too few data
excess pore pressures and thus the duration of the consolida- concerning this problem to draw a definite conclusion.
tion. The Colloquium has thus been divided into three main The comparison of laboratory and in situ permeability test
parts: compressibility, permeability, and consolidation. results clearly show that
Several rheological models have been proposed in the past, -due to remoulding and clogging phenomena the coefficient
but it seems that the u{ -E, - 6, model is the most representa- of permeability measured in pushed piezometers always under-
tive of the behaviour of natural clays in the laboratory. The estimates the real permeability;
strain rate is thus an important factor, which controls the -the permeability measured in the laboratory seems to be
stress-strain curve followed by the clay, depending on the representative of the in situ permeability in homogeneous
strain rate history and thus on the characteristics of the clay and deposits; otherwise, it could strongly underestimate it;
testing technique. The understanding of these phenomena is -the best tool to measure in situ permeability of natural clays
essential to correctly analyse oedometer test results; for seems to be the self-boring permeameter.
example, strain rate effects have the following implications: It is shown that the coefficients of consolidation determined
-the stress- strain curve followed by a clay specimen during a in oedometer tests are much smaller than the ones actually
stage loading test is very different from the one obtained by mobilized in situ. The average value of the ratio is in excess
simply joining the end points, as is usually done; of 20.
-the stress-strain curves followed during consolidation by From what has been said previously, the following are the
subspecimens within the clay specimen are different, depend- main practical implications:
ing on their position relative to the drainage boundary; -The coefficients of consolidation determined graphically
-the preconsolidation pressure and the entire compression from oedometer tests are not representative and cv values must
curve depends on the testing technique used. This implies that be evaluated by separately determining permeability and com-
before adopting a new testing technique, it is necessary to cali- pressibility components.
brate it with the conventional oedometer test, on which long -Since in the large majority of sites the compressible layer
experience has been gained. cannot be reliably characterized by only one set of parameters
In situ stress-strain curves have been established for clay and since most soil engineers now have computers, the use of a
104 CAN. GEOTECH. J. VOL. 25, 1988

multilayer consolidation program, like C O N M U L T (Magnan GCotechnique, 15: 345 -362.


et al. 1979), is strongly suggested. 1969. Time dependent deformation of normally consolidated
-The permeability parameters can b e determined as suggested clays and peats. ASCE Journal of the Soil Mechanics and Founda-
above from direct measurement in the laboratory o r tests using tions Division, 95(SM1): 1-3 1.
the self-boring permeameter, depending o n the homogeneity of BERRE,T. 1969. Discussion, Proceedings, Specialty Session No. 12
on "Advances in consolidation theories for clays." 7th Intema-
the deposit. T h e variation of the permeability coefficient with
tional Conference on Soil Mechanics and Foundation Engineering,
compression can b e estimated by using a Ck value equal to Mexico, pp. 117-118.
0.5eo. BERRE,T., and IVERSEN, K. 1972. Oedometer tests with different
-As for the compressibility parameters, it is necessary to con- specimen heights on a clay exhibiting large secondary compres-
sider that creep deformations occur during primary consolida- sion. GCotechnique, 22: 53-70.
tion, especially in the normally consolidated range. T h e BERRY,P. L., and POSKITT,T. J. 1972. The consolidation of peat.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

consolidation programs developed by Magnan et al. (1979), GCotechnique, 22: 27-52.


Adachi et al. (1982), and Larsson (1986) take into account BJERRUM, L. 1967. Engineering geology of Norwegian normally con-
creep deformations and can b e considered; another possibility solidated marine clays as related to the settlements of buildings.
would b e to modify the stress-strain curve obtained in 24-h GCotechnique, 17: 83 - 119.
oedometer tests to take into account time o r rate effects, as B o z o z u ~ ,M. 1971. Effect of sampling, size and storage on test
suggested in Figs. 16 and 17, and use a consolidation program results for marine clay. ASTM Symposium on Sampling of Soil
and Rock. American Society for Testing and Materials, Special
without any additional consideration for creep. More research
Technical Publication 483, pp. 121 - 131.
is needed in this field. BUISMAN, A. S. 1936. Results of long duration settlement tests. Pro-
ceedings, 1st International Conference on Soil Mechanics and
Acknowledgements Foundation Engineering, Cambridge, Vol. 1, pp. 103- 107.
BURGHIGNOLI, A. 1979. An experimental study of the structural vis-
T h e author feels it is a great honour to have been invited as cosity of soft clays by means of continuous consolidation tests.
lecturer for the Canadian Geotechnical Colloquium and would Proceedings, 7th European Conference on Soil Mechanics and
like to express his sincere appreciation to the Associate Com- Foundation Engineering, Brighton, Vol. 2, pp. 23-28.
mittee o n ~ e o t e c h n i c a ~
l e s e a r c h which
, is responsible for this BURLAND, J. B., and SYMES,M. 1982. A simple axial displacement
event. gauge for use in the triaxial apparatus. GCotechnique, 32: 62-65.
For personal use only.

T h e author is greatly indebted to his colleague Dr. F. CHAN,H. T., and KENNEY, T. C. 1973. Laboratory investigation of
Tavenas, with whom h e has realized a large part of the research permeability of New Liskeard varved soil. Canadian Geotechnical
presented in this paper. H e also would like to thank Dr. M. Journal, 10: 453-472.
Kabbaj, who has contributed to the development of the
CHANG, Y. C. E. 1981. Long term consolidation beneath the test fills
at Vasby, Sweden. Swedish Geotechnical Institute, Stockholm,
thoughts included, and Drs. M . Bozozuk and P. L a Rochelle,
Sweden, Report 13.
who have acted as advisors during the preparation of this Col- CRAWFORD, C. B. 1964. Interpretation of the consolidation test.
loquium. Most of the research carried out at Universitk Lava1 ASCE Journal of the Soil Mechanics and Foundations Division,
was supported by grants from the Fonds pour la formation d e 90(SM5): 87 - 102.
chercheurs et l'aide B la recherche and by t h e Natural Sciences 1965. The resistance of soil structure to consolidation. Cana-
and Engineering Research Council of Canada. dian Geotechnical Journal, 2: 90-97.
1985. Evaluation and interpretation of soil consolidation
ABOSHI,H. 1973. An experimental investigation on the similitude in tests. ASTM Symposium on Consolidation Behaviour of Soils,
the consolidation of a soft clay, including the secondary creep Fort-Lauderdale. American Society for Testing Materials, Special
settlement. Proceedings, 8th International Conference on Soil Technical Publication 892, pp. 7 1- 103.
Mechanics and ~oundation Engineering, Moscow, Vol. 4(3), CRAWFORD, C. B., and BURN,K. N. 1962. Settlement studies on the
p. 88. Mount Sinai Hospital, Toronto. Engineering Journal, 45(12):
ADACHI,K., and TODO,H. 1979. A case study on settlement of soft 31 -37.
clay in Penang. Proceedings, 6th Asian Regional Conference on CROOKS, J. H. A., BECKER,D. E., JEFFERIES, M. G., and MCKEN-
Soil Mechanics and Foundation Engineering, Vol. 1, pp. ZIE,K. 1984. Yield behaviour and consolidation-1: pore pressure
117- 120. response. Proceedings, ASCE Symposium on Sedimentation Con-
ADACHI, T., OKA,F., and TANGE,Y. 1982. Finite element analysis solidation Models: Predictions and Validation, pp. 356-381.
of two dimensional consolidation using an elasto-viscoplastic con- EDEN,W. J., and POOROOSHASB, H. B. 1968. Settlement observation
stitutive equation. Proceedings, 4th International Conference on at Kars Bridge. Canadian Geotechnical Journal, 5: 28-45.
Numerical Methods in Geomechanics, Edmonton, Vol. 1, pp. FELIX,B., VAUILLAT, P., DARVE,F., and FLAVIGNY, E. 1981.
287-296. Comportement visqueux et consolidation des argiles. Proceedings,
ARONOWITZ, A. 1986. Presentation at the annual Meeting of the 10th International Conference on Soil Mechanics and Foundation
Transportation Research Board, Washington. Engineering, Stockholm, Vol. 1, pp. 597 -602.
ASAOKA, A. 1978. Observational procedure of settlement prediction. FOLKES,D. J., and CROOKS, J. H. A. 1985. Effective stress paths and
Soils and Foundations, 18(4): 87 - 101. yielding in soft clays below embankments. Canadian Geotechnical
BAGUELIN, F., JEZEQUEL, J. F., and LE MEHAUTE, A. 1974. Self- Journal, 22: 357 -374.
boring placement method of soil characteristics measurements. Foss, I. 1969. Secondary settlements of buildings in Drammen,
Proceedings, Engineering Foundation Conference on Subsurface Norway. Proceedings, 7th International Conference on Soil
Exploration for Underground Excavation and Heavy Construction, Mechanics and Foundation Engineering, Mexico, Vol. 2, pp.
Henniker, pp. 312-332. 99- 106.
BALASUBRAMANIAN, A. S., and BRENNER, R. P. 1981. Consolidation GIBSON,R. E., and Lo, K. Y. 1961. A theory consolidation for soils
and settlement of soft clay. In Soft clay engineering. Edited by exhibiting secondary compression. Norwegian Geotechnical Insti-
E. W. Brand and R. P. Brenner. Asian Institute of Technology, tute, Oslo, Norway, Report 41.
Bangkok, Thailand, chap. 7. GRAHAM, J., CROOKS,J. H. A., and BELL,A. L. 1983. Time effects
BARDEN,L. 1965. Consolidation of clay with non-linear viscosity. on the stress-strain behaviour of natural soft clays. GCotechnique,
LEROUEIL 105

33: 327 -340. structuree. Proceedings, 32nd Canadian Geotechnical Conference,


HANSBO, S. 1960. Consolidation of clays with special reference to the Qutbec, pp. 2.61 -2.75.
influence of vertical sand drains. Swedish Geotechnical Institute, LEONARDS, G. 1977. Panel discussion, Proceedings International
Stockholm, Sweden, Proceedings 18. Conference on Soil Mechanics and Foundation Engineering,
HOLTZ,R. D., and BROMS,B. 1972. Long term loading tests at Ska- Tokyo, Vol. 3, pp. 384-386.
Edeby, Sweden. ASCE Conference on Performance of Earth and LEONARDS, G. A,, and GIRAULT, P. 1961. A study of the one dimen-
Earth-Supported Structures, Purdue, Vol. 1.1, pp. 435 -464. sional consolidation test. Proceedings, 5th International Confer-
HOLTZ,R. D., JAMIOLKOWSKI, M. B., and LANCELLOTTA, R. 1986. ence on Soil Mechanics and Foundation Engineering, Paris, Vol.
Lessons from oedometer tests on high quality samples. ASCE Jour- 1, pp. 213-218.
nal of Geotechnical Engineering, 112: 768-776. LEROUEIL, S., KABBAJ, M. 1987. Discussion, Settlements analysis of
HVORSLEV, M. J. 1951. Time lag and soil permeability in ground- embankments on soft clays. ASCE Journal of Geotechnical Engi-
water observations. United States Army Waterways Experiment neering, 113: 1067- 1070.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

Station, Vicksburg, MS, Bulletin 36. LEROUEIL, S., and TAVENAS, F. 1981. Pitfalls of back-analyses. Pro-
JAMIOLKOWSKI, M., LADD,C. C., GERMAINE, J. T., and LANCEL- ceedings, 10th International Conference on Soil Mechanics and
LOTTA,R. 1985. New developments in field and laboratory testing Foundation Engineering, Stockholm, Vol. 1, pp. 185 - 190.
of soils. Proceedings, 11th International Conference on Soil 1986. Effective stress paths and yielding in soft clays below
Mechanics and Foundation Engineering, San Francisco, Vol. 1, embankments: Discussion. Canadian Geotechnical Journal, 23:
pp. 57-153. 410-413.
JARRETT, P. M. 1967. Time-dependent consolidation of a sensitive LEROUEIL,S., TAVENAS, F., MIEUSSENS, C., and PEIGNAUD, M.
clay. ASTM Bulletin, 7(7): 300 - 304. 1978. Construction pore pressures in clay foundations under
KABBAJ, M. 1985. Aspects rhCologiques des argiles naturelles en con- embankments. Part 11: generalized behaviour. Canadian Geotech-
solidation. Ph.D. thesis, UniversitC Laval, QuCbec, Que. nical Journal, 15: 66-82.
KABBAJ, M., OKA,F., LEROUEIL, S., and TAVENAS, F. 1985. Con- LEROUEIL, S., SAMSON, L., and B o z o z u ~ ,M. 1983a. Laboratory
solidation of natural clays and laboratory testing. ASTM Sympo- and field determination of preconsolidation pressures at Gloucester.
sium on Consolidation Behaviour of Soils, Fort Lauderdale. Canadian Geotechnical Journal, 20: 477 -490.
American Society for Testing and Materials, Special Technical LEROUEIL,S., TAVENAS, F., SAMSON, L., and MORIN,P. 1983b.
Publication 892, pp. 378-403. Preconsolidation pressure of Champlain clays. Part 11. Laboratory
KABBAJ, M., TAVENAS, F., and LEROUEIL, S. 1988. In situ and lab- determination. Canadian Geotechnical Journal, 20: 803 -816.
oratory stress-strain relations. GCotechnique, in press. LEROUEIL, S., KABBAJ, M., and TAVENAS, F. 1985a. Discussion on
~ R I S E LJ., 1985. Histoire de I'ingCniCrie gComCcanique jusqu'h theme lecture 2-B on laboratory testing. Proceedings, 1lth Interna-
For personal use only.

1700. Proceedings, 1lth International Conference on Soil Mechan- tional Conference on Soil Mechanics and Foundation Engineering,
ics and Foundation Engineering, San Francisco, Jubilee volume, San Francisco.
pp. 3-90. LEROUEIL,S., KABBAJ,M., TAVENAS,F., and BOUCHARD, R.
KOPPEJAN,A. W. 1948. A formula combining the Terzaghi load 19856. Stress - strain - strain rate relation for the compressibility
compression relationship and the Buisman secular time effect. Pro- of sensitive natural clays. GCotechnique, 35: 159- 180.
ceedings 2nd International Conference on Soil Mechanics and LEROUEIL, S., MAGNAN, J. P., and TAVENAS, F. 1 9 8 5 ~ Remblais
.
Foundation Engineering, Rotterdam, Vol. 3, pp. 32-37. sur argiles molles. Technique et Documentation, Paris, France.
LACASSE, S., BERRE,T., and LEFEBVRE, G. 1985. Block sampling of LEROUEIL, S., KABBAJ, M., TAVENAS, F., and BOUCHARD, R. 1986.
sensitive clays. Proceedings, 1lth International Conference on Soil Reply to Discussion, Stress - strain - strain rate relation for the
Mechanics and Foundation Engineering, San Francisco, Vol. 2, compressibility of sensitive natural clays. GCotchnique, 36:
pp. 887-892. 288-290.
LADD, C. C., FOOTT, R., ISHIHARA, K., SCHLOSSER, F., and LEROUEIL, S., DIENE,M., TAVENAS, F., KABBAJ,M., and LA
P o u ~ o s H.
, G. 1977. Stress-deformation and strength characteris- ROCHELLE, P. 1988a. Direct determination of the permeability of
tics. State-of-the-Art Report, Proceedings, 9th International Con- clay under embankments. ASCE Journal of Geotechnical Engineer-
ference on Soil Mechanics and Foundation Engineering, Tokyo, ing, 114, in press.
Vol. 2, pp. 421-494. LEROUEIL, S., KABBAJ,M., and TAVENAS, F. 1988b. Study of the
LAMBE,T. W., and WHITMAN, R. V. 1969. Soil mechanics. John validity of a 0: -E, - i, model in in situ conditions. Soils and Foun-
Wiley and Sons, New York, NY. dations, in press.
LA ROCHELLE, P., and LEFEBVRE, G. 1971. Sampling disturbance in LEWIS,W. A., MURRAY, R. T., and SYMONS, I. F. 1976. Settlement
Champlain clays. ASTM Symposium on Sampling of Soil and and stability of embankments constructed on soft alluvial soils.
Rock, American Society for Testing and Materials, Special Techni- Proceedings-the Institution of Civil Engineers, 59: 57 1-593.
cal Publication 483, pp. 143- 163. MAGNAN, J. P. 1983. ThCorie et pratique des drains verticaux. Tech-
LA ROCHELLE,P., SARRAILH, J., TAVENAS, F., ROY, M., and nique et Documentation, Paris, France.
LEROUEIL, S. 1981. Causes of sampling disturbance and design of MAGNAN, J. P., and DEROY,J. M. 1980. Analyse graphique des tas-
a new sampler for sensitive soils. Canadian Geotechnical Journal, sements observCs sous les ouvrages. Bulletin de Liaison des Labo-
Vol. 18: 52-66. ratoires des Ponts et Chausstes, no. 109: 45-52.
LARSSON, R. 1981. Drained behaviour of Swedish clays. Swedish MA;GNAN, J. P., BAGHERY, S., BRUCY,M., and TAVENAS, F. 1979.
Geotechnical Institute, Stockholm, Sweden, Report 12. Etude numCrique de la consoIidation unidimensionnelle en tenant
1986. Consolidation of soft soils. Swedish Geotechnical Insti- compte des variations de la permCabilitt et de la compressibilitC du
tute, Stockholm, Sweden, Report 29. sol, du fluage et de la non-saturation. Bulletin de Liaison des Labo-
LARSSON, R., and SALLFORS, G. 1985. Automatic continuous consol- ratoires des Ponts et ChaussCes, no. 103: 83 -94.
idation testing in Sweden. ASTM Symposium on Consolidation MAGNAN, J. P., MIEUSSENS, C., and QUEYROI, D. 1983. ~ t u d d'un
e
Behaviour of Soils, Fort Lauderdale, American Society for Testing remblai sur sols compressibles: le remblai B du site expCrimenta1
and Materials, Special Technical Publication 892, pp. 299-328. de Cubzac-les-Ponts. Laboratoire Central des Ponts et ChaussCes,
LAW,K. T., and LEE,C. F. 1981. Initial gradient in a dense glacial Paris, France, Research Report LPC 127.
till. Proceedings, 10th International Conference on Soil Mechanics MESRI,G., and CASTRO,A. 1987. C,/C, concept and KOduring sec-
and Foundation Engineering, Stockholm, Vol. 1, pp. 441 -446. ondary compression. ASCE Journal of Geotechnical Engineering,
LEFEBVRE,G., and PHILIBERT, A. 1979. Mesure des pressions 113: 230-247.
laterales durant la consolidation unidimensionnelle d'une argile MESRI,G., and CHOI,Y. K. 1979. Strain rate behaviour of the Saint-
106 CAN. GEOTECH. J. VOL. 25, 1988

Jean Vianney clay: Discussion. Canadian Geotechnical Journal, clays. Ph.D. thesis, Chalmers University of Technology, Gothen-
16: 831 -834. burg, Sweden.
1984. Discussion, Time effects on the stress-strain behav- SAMSON, L., LEROUEIL, S., MORIN,P., and LE BIHAN,J. P. 1981.
iour of natural soft clays. GCotechnique, 34: 439-442. Pressions de prkconsolidation des argiles sensibles. DSS Contract
1985a. Settlement analysis of embankments on soft clays. lSX79-00026. Division of Building Research, National Research
ASCE Journal of Geotechnical Engineering, 111: 441 -464. Council of Canada.
19856. The uniqueness of the end-of-primary (EOP) void SCHIFFMAN, R. L., PANE,V., and GIBSON, R. E. 1984. The theory of
ratio - effective stress relationship. Proceedings, 1lth Interna- one-dimensional consolidation of saturated clays. IV. An overview
tional Conference on Soil Mechanics and Foundation Engineering, of non-linear finite strain sedimentation and consolidation. ASCE
San Francisco, Vol. 2, pp. 587-590. Symposium on Sedimentation Consolidation Models-Prediction
MESRI,G., and FENG,T. W. 1986. Discussion, Stress - strain - and Validation, San Francisco, pp. 1-29.
strain rate relation for the compressibility of sensitive natural clays. SEKIGUCHI, H., and TORIIHARA, M. 1976. Theory of one-dimen-
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13

Gkotechnique, 36: 283 -287. sional consolidation of clays with consideration of their rheological
MESRI,G., and GODLEWSKI, P. M. 1977. Time and stress compres- properties. Soils and Foundations, 16(1): 27 -44.
sibility interrelationship. ASCE Journal of the Geotechnical Engi- SHIELDS,H. 1974. Discussion on session on Normally consolidated
neering Division, 103(GT5): 417 -430. and lightly overconsolidated cohesive materials. British Geotechni-
MESRI,G., and ROKHSAR, A. 1974. Theory of consolidation for cal Society Conference on Settlement of Structures, Cambridge,
clays. ASCE Journal of the Geotechnical Engineering Division, pp. 698-700.
lOO(GT8): 889-904. SHOJI,M., and MATSUMOTO, T. 1976. Consolidation of embankment
MIEUSSENS, C., and DUCASSE, P. 1977. Mesure en place des coeffi- foundation. Soils and Foundations, 16(1): 59-74.
cients de permCabilitB et des coefficients de consolidation horizon- SILVESTRI, V. 1981. Behavior of an overconsolidated sensitive clay in
taux et verticaux. Canadian Geotechnical Journal, 14: 76-90. drained KOtriaxial tests. Proceedings ASTM Symposium on Labor-
MILLER,R. J., and Low, P. F. 1963. Threshold gradient for water atory Shear Strength of Soil. American Society for Testing and
flow in clay system. Soil Science Society of America Proceedings, Materials, Special Technical Publication 740, pp. 619-630.
27(6): 605 -609. SILVESTRI, V., YONG,R. N., SOULIB,M., and GABRIEL, F. 1985.
MORIN,P., LEROUEIL, S., SAMSON, L. 1983. Preconsolidation pres- Controlled-strain, controlled-gradient and standard consolidation
sure of Champlain clays. Part I. In situ determination. Canadian testing of sensitive clays. ASTM Symposium on Consolidation
Geotechnical Journal, 20: 782-802. Behaviour of Soils, Fort Lauderdale, American Society for Testing
MURRAY, R. T. 1978. Developments in two and three-dimensional and Materials, Special Technical Publication 892, pp.433 -450.
consolidation theory. Developments in soil mechanics-1 . Applied SIMONS,N. E. 1975. General report "Normally consolidated and
For personal use only.

Science Publishers Ltd., London, England. highly over-consolidated cohesive materials." Conference of the
OKA,F. 1981. Prediction of time-dependent behaviour of clay. Pro- British Geotechnical Society on Settlements of Stmctures, Cam-
ceedings, 10th International Conference on Soil Mechanics and bridge, pp. 500-530.
Foundation Engineering, Stockholm, Vol. 1, pp. 215 -218. S O U L I M.,
~ , and SILVESTRI, V. 1984. Consolidation performance of
OLSEN,H. W. 1985. Osmosis: a cause of apparent deviations from soft clays-Part Il-Predictions. ASCE Symposium on Sedimenta-
Darcy's law. Canadian Geotechnical Journal, 22: pp. 238 -241. tion Consolidation Models-Predictions and Validation, San Fran-
OLSON,R. E. 1985. State-of-the-art: consolidation testing. ASTM cisco, pp. 197-215.
Symposium on Consolidation of Soils, Fort Lauderdale. American SRIDHARAN, A,, and SREEPADA RAO, A. 1981. Rectangular hyper-
Society for Testing and Materials, Special Technical Publication bola fitting method for one dimensional consolidation. Geotech-
892, pp. 7-68. nical Testing Journal, 4(4): 161 - 168.
OLSON,R. E., and DANIEL,D. E. 1981. Measurement of the hydrau- SUKLJE,L. 1957. The analysis of the consolidation process by the
lic conductivity of fine-grained soils. ASTM Symposium on per- isotache method. Proceedings, 4th International Conference on
meability and groundwater containment transport. American Soil Mechanics and Foundation Engineering, London, Vol. 1, pp.
Society of Testing and Materials, Special Technical Publication 200-206.
746, pp. 18-64. 1969. Rheological aspects of soil mechanics. Wiley-Intersci-
PELLETIER, J. H., OLSON,R. E., and RIXNER, J. J. 1979. Estimation ence, London, England.
of consolidation properties of clay from field observations. Geo- TAVENAS, F., and LEROUEIL, S. 1977. Effects of stresses and time on
technical Testing Journal, 2(1): 34-43. yielding of clays. Proceedings, 9th International Conference on
POOROOSHASB, H. B., and SIVAPATHAM, T. 1969. Consolidation of Soil Mechanics and Foundation Engineering, Tokyo, Vol. 1, pp.
sensitive clays exhibiting strong structural breakdown. Proceed- 319-326.
ings, 7th International Conference on Soil Mechanics and Founda- TAVENAS, F., BRUCY,M., MAGNAN, J. P., LA ROCHELLE, P., and
tion Engineering, Mexico City, Specialty Session 12, pp. 27 -37. ROY,M. 1979. Analyse critique de la thkorie de consolidation uni-
POOROOSHASB, H. B., LAW,K. T., BOZOZUK, M., and EDEN,W. J. dimensionnelle de Terzaghi. Revue Fran~aisede GCotechnique,
1981. Consolidation of sensitive clays. Proceedings, 10th Interna- no. 7: 29-43.
tional Conference on Soil Mechanics and Foundation Engineering, TAVENAS, F., JEAN,P., LEBLOND, P., and LEROUEIL, S. 1983a. The
Stockholm, Vol. 1, pp. 219-223. permeability of natural soft clays. Part 11: Permeability characteris-
POSKITT,T. J., and BIRDSALL, R. 0. 1971. A theoretical and experi- tics. Canadain Geotechnical Journal, 20: 645 -660.
mental investigation of mildly nonlinear consolidation behavior in TAVENAS, F., LEBLOND, P., JEAN,P., and LEROUEIL, S. 19836. The
saturated soil. Canadian Geotechnical Journal, 8: 182 -2 16. permeability of natural soft clays. Part I: Methods of laboratory
RANDOLPH, M. F., and BOOKER, J. R. 1982. Analysis of seepage into measurement. Canadian Geotechnical Journal, 20: 629-644.
a cylindrical permeameter. Proceedings, 4th International Confer- TAVENAS, F., TREMBLAY, M., and LEROUEIL,S. 1 9 8 3 ~ .Mesure
ence on Numerical Methods in Geomechanics, Edmonton, Vol. 1, in situ de la permCabilitC des argiles. Bulletin de 1'Association
pp. 349-357. Internationale de GCologie de I'Ingknieur, no. 26 -27: 509 -513.
RIZKALLAH, V., BLUMEL, W., and RICHWIEN, W. 1977. Consolida- TAVENAS, F., DIENE,M. and LEROUEIL, S. 1986a. Analysis of the
tion and shear strength of soft organic clay of Bremerhaven. Pro- in situ constant head permeability test. Proceedings, 39th Canadian
ceedings, Symposium on Soft Clay, Bangkok, pp. 133- 145. Geotechnical Conference, Ottawa, pp. 71 -77.
ROWE,P. W. 1972. The relevance of soil fabric to site investigation TAVENAS, I?., TREMBLAY, M., LAROUCHE, G., and LEROUEIL, S.
practice. GCotechnique, 22: 195-300. 19866. In situ measurement of permeability in soft clays. ASCE
SALLFORS, G. 1975. Preconsolidation pressure on soft high plastic Specialty Conference In situ '86, Blacksburg, pp. 1034- 1048.
UEIL 107

TAYLOR, D. W., and MERCHANT, W. 1940. A theory of clay consoli- WALKER, L. K., and MORGAN, J. R. 1977. Field performance of a
dation accounting for secondary compression. Journal of Mathema- firm silty clay. 9th International Conference on Soil Mechanics and
tics and Physics (Cambridge, Mass), 19: 167- 185. Foundation Engineering, Tokyo, Vol. 1, pp. 341 -346.
TERZAGHI, K. 1923. Die Berechnung de Durchlassig keit des Tones WILKINSON, W. B. 1968. Constant head in situ permeability tests in
aus dem Verlauf de hydrodynamischen Spannungserscheinungen. clay strata. GCotechnique, 18: 172 - 194.
Akademie de Wissenschaften, Wien, Sitzungsberichte, Mathe- Wu, T. H., RESENDIZ, D., and NEUKIRCHNER, R. J. 1966. Analysis
matisch-Natunvissenschaffliche Klasse, part IIa 132(3/4): of consolidation by rate process theory. ASCE Journal of the Soil
125- 138. Mechanics and Foundations Division, 92(SM6): 229-248.
VAID,Y. P., ROBERTSON, P. K., and CAMPANELLA, R. G. 1979. YAMAMOTO, S. 1977. Land subsidence in Japan. Japanese Society of
Strain rate behaviour of Saint-Jean-Vianney clay. Canadian Geo- Soil Mechanics and Foundation Engineering, Tsuchi-to-kiso, pp.
technical Journal, 16: 34 -42. 13- 19.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by NC STATE UNIVERSITY on 01/17/13
For personal use only.

You might also like