Sie sind auf Seite 1von 23

513

Bearing capacity under cyclic loading — offshore,


along the coast, and on land. The 21st Bjerrum
Lecture presented in Oslo, 23 November 20071
Knut H. Andersen

Abstract: Cyclic loading can be important for the foundation design of structures, both offshore, along the coast, and on
land, and for the stability of slopes. This is illustrated by several examples. The paper discusses how soil behaves under
cyclic loading, both for structures and for slopes, and shows that the cyclic shear strength and the failure mode under cy-
clic loading depend strongly on the stress path and the combination of average and cyclic shear stresses. Diagrams with
the cyclic shear strength of clay, sand, and silt that can be used in practical design are presented. Comparisons between
calculations and model tests indicate that foundation capacity under cyclic loading can be determined on the basis of cy-
clic shear strength determined in laboratory tests.
Key words: cyclic shear strength, bearing capacity, clay, sand, silt, design diagrams.
Résumé : Le chargement cyclique est important pour la conception de fondations de structures, autant en mer, sur la côte
et sur la terre, et pour la stabilité des pentes. Ceci est illustré à l’aide de plusieurs exemples. Cet article discute du compor-
tement du sol soumis à un chargement cyclique, pour la structure et les pentes, et démontre que la résistance au cisaille-
ment cyclique et le mode de défaillance sous chargement cyclique dépendent fortement du cheminement des contraintes et
de la combinaison des contraintes en cisaillement moyennes et cycliques. Des diagrammes de résistance au cisaillement
cyclique pour l’argile, sable et silt, qui peuvent être utilisés pour la conception, sont présentés. Des comparaisons entre les
calculs et les essais modélisés indiquent que la capacité de la fondation soumise à un chargement cyclique peut être déter-
minée à partir de la résistance au cisaillement cyclique obtenu par des essais en laboratoire.
Mots-clés : résistance au cisaillement cyclique, capacité portante, argile, sable, silt, diagrammes de conception.
[Traduit par la Rédaction]

Introduction practical foundation design and research in this area (e.g.,


Bearing capacity under cyclic loading governs the founda- Bjerrum 1973). Since then, the cyclic behaviour of soil has
tion design of offshore installations and cyclic soil behav- been an important topic at NGI.
iour has received significant attention over the last 30– This paper first gives examples of cases where cyclic
40 years. However, there are also installations and slopes loading is important for foundation design. Then it explains
along the coast and on land where it is important to consider how soil behaves under cyclic loading, and diagrams with
the effects of cyclic loading on bearing capacity. cyclic shear strength of clay, silt, and sand that can be used
At the Norwegian Geotechnical Institute (NGI), the be- in practical design are provided. Finally, comparisons of cal-
haviour of soils under cyclic loading started in the early culated to measured bearing capacity in model tests show
1970s in connection with the first exploitation of oil and that bearing capacity under cyclic loading can be reliably
gas in the North Sea. Dr. Laurits Bjerrum, NGI’s director at predicted based on cyclic strength from laboratory tests.
that time, saw the geotechnical challenges in connection
with the foundation design of the fixed platforms that were Examples of cyclic loading
to be constructed, and he enthusiastically involved NGI in
The first gravity platform in the North Sea was the Eko-
Received 14 May 2008. Accepted 14 January 2009. Published fisk oil storage tank in Fig. 1, which was installed in June
on the NRC Research Press Web site at cgj.nrc.ca on 1 May 1973 (Clausen et al. 1975). The tank was designed for a
2009. 100 year storm with a maximum wave of 24 m. In addition
K.H. Andersen. Norwegian Geotechnical Institute (NGI),
to the maximum wave, the foundation had to withstand the
Sognsveien 72, P.O. Box 3930, Ullevaal Stadion, Oslo 0806, forces from a large number of other waves in the storm.
Norway (e-mail: Knut.h.andersen@ngi.no). These forces may generate excess pore pressure and degrade
1This
the soil prior to the impact of the maximum wave.
paper represents the written version of the 21st Bjerrum
Lecture. While it has been edited for the present publication, it
Since the installation of the Ekofisk tank, a large number
retains the general structure of the original lecture, which was of gravity platforms have been installed in the North Sea
intended for a general geotechnical audience. The Bjerrum and other parts of the world. In several cases, the platforms
Lecture is presented in Oslo in alternate years by the have been designed for severe storms with a maximum
Norwegian Geotechnical Society with the support of the 100 year wave in excess of 30 m. One example is the Frigg
Bjerrum Memorial Fund (Laurits Bjerrums Minnefond). TCP2 Condeep platform in Fig. 2.

Can. Geotech. J. 46: 513–535 (2009) doi:10.1139/T09-003 Published by NRC Research Press
514 Can. Geotech. J. Vol. 46, 2009

Fig. 1. Ekofisk oil storage tank, installed in the North Sea in June 1973 (from Clausen et al. 1975; # 1975 Offshore Technology Confer-
ence).

Fig. 2. Frigg TCP2 Condeep platform in the North Sea (adapted from Andersen and Høeg 1991).

In recent years, gravity platforms have also been designed ample is the Snorre tension leg platform (TLP), which in
to serve as liquified natural gas (LNG) terminals. These ter- 1991 was the first tension leg installation to be anchored by
minals have been situated closer to land than the installa- means of suction anchors (Christophersen et al. 1992)
tions for oil and gas production. Geotechnically, this has (Fig. 3). Due to the novelty of the suction anchor concept at
presented new challenges with soil profiles consisting of that time, extensive studies were made to ensure that the
mixed layers of loose silt, sand, and soft clay often signifi- foundation design methods were sound. This included field-
cantly less uniform than encountered offshore. model testing that is described towards the end of this paper.
Essentially, all offshore installations will experience wave Suction anchors have since been extensively used to anchor
loading and need to be designed for cyclic loading. In addi- various types of floating installations; an example is shown
tion to the gravity platforms mentioned above, this includes in Fig. 4. By the end of 2004, about 500 suction anchors
piled installations, anchors, and skirted installations. One ex- were installed at 50 different offshore locations worldwide

Published by NRC Research Press


Andersen 515

Fig. 3. Snorre TLP platform installed in the North Sea in 1991. Fig. 5. Oosterschelde storm-surge barrier. (www.deltawerken.com/
Deltaworks, photo by the Ministry of Public Works, Transport and
Water Management (the Netherlands)).

Fig. 6. Offshore wind turbines (www.vestas.com).


Fig. 4. Suction anchors installed at the Horn Mountain field in the
Gulf of Mexico in 2002 (photo by E.C. Clukey, BP America Pro-
duction Company).

Fig. 7. Vibrations from traffic (photo by Christian Madshus, NGI).

failures of buildings and bridges (Fig. 8) and induce slope


in water depths reaching more than 2000 m (Andersen et al. failures (Fig. 9). Cyclic loading from traffic and earthquakes
2005). have lower cyclic load periods (*1 s) than wave loading
Structures along the coast may also be subjected to severe (*10–20 s) and this needs to be taken into account.
wave loading. Such structures include harbours, breakwaters, Ice loading can cause cyclic loads on structures that are
and storm-surge barriers. One example is the Oosterschelde placed in the sea. The Great Belt bridge in Denmark, for in-
storm-surge barrier in the Netherlands that was completed in stance (Fig. 10), was designed for cyclic loading from ice
1986 (Fig. 5). sheets that break when they float past and impact on the
Many wind turbines are now placed offshore (Fig. 6), and bridge piers (Andersen et al. 1991). Andersen et al. (1991)
their foundations need to be designed for cyclic loading reported that the breaking of the ice sheets would cause an
from both wind and waves. Even if placed on land, they do impulse that would set the bridge piers in motion with a fre-
of course need to be designed for cyclic wind loading. quency determined by their resonance periods. This would
On land there are also situations with cyclic loading. This lead to a two-frequency cyclic loading on the soil beneath
includes design of road and railway embankments and cuts the piers (Fig. 10). The importance of ice loading on off-
to withstand the effects of traffic-induced cyclic loading shore structures will increase as oil and gas production
(Fig. 7). In seismic areas, earthquakes may cause foundation moves into arctic areas.

Published by NRC Research Press


516 Can. Geotech. J. Vol. 46, 2009

Fig. 8. Earthquake damage to buildings (photo by Amir Kaynia, placements and fails at a cyclic load that is lower than the
NGI). failure load under monotonic loading.
In this paper, the failure load and the shear stress at fail-
ure under monotonic loading will be called static capacity
and static shear strength, respectively.
The reason why the cyclic capacity may be smaller than
the static capacity is that the cyclic loading tends to break
down the soil structure and cause a tendency for volumetric
reduction in the soil. If the conditions are undrained, volu-
metric changes will be prevented by the low volumetric
compressibility of the water. The normal stresses that were
carried by the soil will then be transferred to the pore water
and the effective stresses in the soil will decrease accord-
ingly. This is illustrated by the effective stress paths for
monotonic and cyclic soil elements in Fig. 12.
The development of pore pressure and shear strain with
time for a soil element subjected to undrained cyclic loading
with a constant cyclic shear stress is illustrated in Fig. 13. The
load cycles with a single-amplitude shear stress, tcy, around a
constant shear stress, ta. The cyclic loading generates a pore
pressure characterized by a permanent pore-pressure compo-
nent, up, and a cyclic pore pressure component, ucy. The in-
creased pore pressure reduces the effective stresses in the
Fig. 9. Earthquake-induced landslide, El Salvador (U.S. Geological
soil, resulting in increased permanent, gp, and cyclic, gcy,
Survey, photo by Ed Harp, USGS).
shear strains with time.
The stress–strain behaviour of a soil element under the
cyclic loading in Fig. 13 is illustrated in Fig. 14.

Effect of average shear stress and test type


The soil beneath a structure is subjected to various stress
conditions where the average and cyclic shear stresses and
the type of loading (e.g., triaxial versus direct simple shear,
DSS) vary from one point to another. This is illustrated in
Fig. 15, which shows a simplified image of the stress condi-
tions along a potential failure surface beneath a gravity
structure under cyclic loading.
The behaviour of soil elements under various loading con-
ditions is illustrated in Figs. 16 and 17 with results from lab-
oratory tests on Drammen clay, a marine clay with a
plasticity index of Ip = 27%.
The first two tests in Fig. 16, Figs. 16a and 16b, show
that the response to symmetrical cyclic loading is different
in DSS and triaxial tests. In the DSS tests, the shear strain
develops relatively symmetrically, apart from the first quar-
ter cycle. In the triaxial test, the shear strain is nonsymmet-
rical with a permanent shear strain of the same magnitude as
the cyclic shear strain. This is related to the strength aniso-
What happens to soil subjected to cyclic
tropy under triaxial loading, with an extension strength that
loading? is smaller than the compression strength.
General behaviour The third test, Fig. 16c, has a shear stress where the aver-
Cyclic loading may reduce the bearing capacity of a soil age and the cyclic components are the same. The result is a
and the bearing capacity under cyclic loading may be lower shear-strain development with dominant permanent shear
than the capacity under monotonic loading. This is illus- strain and small cyclic shear strain that does not increase
trated by the results of model tests of a gravity platform on significantly with number of cycles.
clay in Fig. 11. The model has a submerged weight, W’, and Figure 17 shows that the cyclic behaviour is not governed
is loaded with a cyclic horizontal load, Hcy, a distance z by the maximum shear stress (tmax). The three tests in Fig. 17
above the sea floor. The results show that the displacements have the same maximum shear stress, but different average
under cyclic loading increase with the number of cycles and and cyclic shear stress components. The test with ta = 0 fails
become larger than the displacements under monotonic load- after 10 cycles, whereas the tests with ta = 0.5tmax and ta =
ing at the same load. The model develops excessive dis- 0.85tmax have developed only small shear strains after 2500

Published by NRC Research Press


Andersen 517

Fig. 10. Ice load history for Great Belt bridge piers (photo by Henrik Sendelbach; commons.wikimedia.org/wiki/File:Storeb%C3%A6lts-
broen_from_Sj%C3%A6lland.jpg).

Fig. 11. Results of model tests with monotonic and cyclic loading Fig. 12. Effective stress paths for undrained tests with monotonic
on gravity platform on clay (adapted from Dyvik et al. 1989). and cyclic loading. up, permanent pore-pressure component; s’, ef-
fective normal stress; t, shear stress. ta, average shear stress; tcy,
single-amplitude cyclic shear stress.

cycles, and the test with the highest ta has the smallest shear Drammen clay (e.g., Andersen et al. 1988) were plotted as
strains. function of average and cyclic shear stresses.
These examples illustrate that the cyclic behaviour de- The plot for DSS tests on normally consolidated clay is
pends on both average and cyclic shear stresses, and that presented in Fig. 18. Fig. 18a shows the results from one
the behaviour is different in triaxial and DSS tests. The monotonic and nine cyclic tests. The location of the various
next section describes how this knowledge can be utilized points is determined by the average and cyclic shear stresses
to establish design diagrams for practical applications. under which the tests were run, and the numbers along each
point give the number of cycles to failure, Nf, and the per-
Number of cycles to failure as function of average and manent and cyclic shear strains at failure, gp/gcy. Failure
cyclic shear stresses was defined as either a permanent or a cyclic shear strain of
Given the importance of the average and cyclic shear 15%. The shear stresses in Fig. 18 are normalized with re-
stress components for the cyclic behaviour, the results from spect to the static shear strength, su, in a DSS test run to
a large DSS and triaxial cyclic testing programme on plastic failure with a constant rate of shear strain of *4.5%/h.

Published by NRC Research Press


518 Can. Geotech. J. Vol. 46, 2009

Fig. 13. Pore-pressure and shear strain as function of time under Fig. 15. Simplified stress conditions along a potential failure sur-
undrained cyclic loading. u, pore pressure; g, shear strain; t0, initial face in the soil beneath a gravity structure under cyclic loading. H,
consolidation shear stress. resultant horizontal load; h, height above seafloor of resultant hori-
zontal load.

Fig. 14. Stress–strain behaviour under cyclic loading.


stresses approaching the compression shear strength, the
failure mode will be dominated by large permanent com-
pression shear strains, whereas for average shear stresses ap-
proaching the extension strength, the failure mode will be
dominated by large permanent extension shear strains.
Strains in compression are defined as positive and strains in
extension are defined as negative in Fig. 19 and later fig-
ures.
The intersection of the contours with the ta–axis and the
location of the contours for values of ta approaching the
static undrained shear strength depend upon the duration of
ta because of the undrained creep that occurs under high
average shear stress. In the diagrams in this paper, the con-
tours have been drawn to intersect the horizontal axis at the
undrained shear strength in standard laboratory tests. If the
duration of ta deviates from the approximately 2 h time to
By inspection of the data in Fig. 18a, it was possible to failure in a standard monotonic laboratory test, this can be
draw contours through points with the same number of corrected for when applying the diagrams by using a shear
cycles to failure. This gave the contour diagram in Fig. 18b. strength corrected for load duration when denormalizing ta/
Curves that define the failure mode, i.e., the permanent and su. The shear strength correction can be based on the data in
cyclic shear strains at failure, are drawn in addition. Fig. 20, which shows the shear strength of several clays as a
The contour diagram defines the number of cycles to fail- function of time to failure.
ure as a function of average and cyclic shear stresses. The
diagram also defines the failure mode; i.e., the combination
of permanent and cyclic shear strains at failure, gp and gcy. Cyclic shear strength
The diagram shows that the failure will occur as large cyclic The diagrams in Figs. 18 and 19 show the combinations
shear strains when the average shear stress is small, and as of average and cyclic shear stresses that will cause failure
large permanent shear strains when the average shear stress in a given number of cycles. In a stability analysis, one
approaches the static shear strength. For average shear needs to know the maximum shear stress that can be mobi-
stresses between zero and the static shear strength, the fail- lized. As illustrated in Fig. 21, the maximum shear stress
ure mode will be a combination of permanent and cyclic that can be mobilized is the sum of average and cyclic shear
shear strains. stresses at failure, i.e.,
A contour diagram similar to the one for DSS tests is ½1 t f;cy ¼ t a;f þ t cy;f
given for triaxial tests in Fig. 19. The diagram is established
based on a number of cyclic triaxial tests, each with differ- where tf,cy is the cyclic shear strength, ta,f is the average
ent average and cyclic shear stresses, such as for the DSS shear stress at failure, and tcy,f is the cyclic shear stress at
tests. The diagram in Fig. 19 shows that the triaxial tests failure. ta,f is composed of the initial consolidation shear
will fail with large cyclic shear strains for a range of aver- stress, t0, plus the change in average shear stress, Dta, due
age shear stresses about halfway between the compression to an additional average load or redistribution of average
and extension static shear strengths. For average shear stresses, i.e.,

Published by NRC Research Press


Andersen 519

Fig. 16. Stress–strain behaviour under various cyclic loading conditions.

Fig. 17. Results from cyclic triaxial tests with the same maximum Fig. 20). As the cyclic tests are run with a load period of
shear stress, tmax. 10 s and the monotonic tests are brought to failure in about
2 h, the cyclic strength may thus be higher than the conven-
tional static shear strength.
The diagrams in Figs. 22 and 23 provide the data required
to define the cyclic shear strength for the various stress con-
ditions shown in Fig. 15.
All the diagrams in this paper assume that Dta is applied
undrained. A diagram showing the effect of applying Dta
drained is presented for normally consolidated Drammen
clay in Andersen (1988). The effect of drained versus un-
drained Dta for very dense sand in presented in Andersen
and Berre (1999).
The diagrams also assume that the soil is undrained dur-
ing a cycle. If it is possible for drainage or pore-pressure re-
distribution to occur during a cycle, as it may be for sand in
some cases, one should consider limiting the cyclic strength
to the drained shear strength.

Equivalent number of cycles


½2 t a;f ¼ t 0 þ t a The diagrams in Figs. 22 and 23 give the cyclic shear
For clays, Dta would normally be undrained, but for sands, strength for cases with a constant cyclic shear stress during
Dta may be drained. the cyclic load history. In a storm, however, the cyclic shear
Equation [1] means that the cyclic shear strength for a stress is likely to vary from one cycle to the next. The
given number of cycles to failure can be established by sim- equivalent number of cycles of the maximum cyclic shear
ply adding average and cyclic shear stresses for various stress, Neqv, that would give the same effect as the real cy-
given values of average shear stress in the diagrams in clic load history must therefore be determined to be able to
Figs. 18 and 19, and then plotting this sum as a function of use the diagrams.
the average shear stress. This is done as illustrated for DSS Procedures to determine Neqv are presented by Andersen
tests in Fig. 22 and for triaxial tests in Fig. 23. The reason (1976) and Andersen et al. (1992). The procedures use
for the discontinuity in the curves for triaxial tests in shear-strain or pore-pressure contour diagrams of the type
Fig. 23 is that the average stress has opposite sign for com- presented in Figs. 24 and 25. The diagrams are established
pression and extension. Compression is defined as a failure from the same tests as used to establish the diagrams in
mode where the permanent shear strain is positive, and ex- Figs. 18 and 19. The diagrams in Figs. 24 and 25 are for
tension is defined as the failure mode where the permanent DSS tests with an average shear stress of t a = 0. Similar di-
shear strain is negative. agrams can be established for other average shear stresses
The cyclic shear strength diagrams show that the cyclic and for triaxial tests. Various such diagrams are given in
shear strength can be higher than the static shear strength, Andersen (2004).
as tf,cy/su > 1 for a low number of cycles in some cases. For undrained conditions, Neqv may be determined by the
This is because the clay strength is rate dependent (e.g., ‘‘strain accumulation’’ procedure where the cyclic shear

Published by NRC Research Press


520 Can. Geotech. J. Vol. 46, 2009

Fig. 18. Number of cycles to failure and failure mode as a function of average and cyclic shear stresses for cyclic DSS tests on normally
consolidated Drammen clay. (a) Results from one monotonic and nine cyclic tests; (b) contour diagram with the same number of cycles to
failure based on the data in part (a). ta,f, average shear stress at failure; tcy,f, cyclic shear stress at failure; suDSS, static DSS shear strength.

Fig. 19. Number of cycles to failure and failure mode as a function of average and cyclic shear stresses for cyclic triaxial tests on normally
consolidated Drammen clay. gp, permanent shear strain; suC, static compression shear strength.

Fig. 20. Static shear strength of several clays as a function of time to Fig. 21. Definition of cyclic shear strength, tf,cy.
failure (adapted from Lunne and Andersen 2007). tf, time to failure.

storm, it is necessary to account for the effect of drainage.


The equivalent number of cycles, Neqv, can then be deter-
mined by keeping track of the permanent pore pressure ac-
cumulated during the cyclic load history, accounting for the
simultaneous generation and dissipation of the excess pore
pressure. Andersen et al. (1994) describes the pore-pressure
accumulation procedure and how the pore-pressure dissipa-
tion can be taken into account.
In principle, the pore-pressure accumulation procedure
could also be used for clays. In practice, however, accurate
laboratory measurement of pore pressure is more difficult to
strain is used as a memory of the cyclic effect by keeping perform in clays than in sand. As drainage is not likely to oc-
track of the cyclic shear strain during the cyclic load history. cur during the cyclic load history in clays, it may be prefera-
For sands or situations where drainage can occur during a ble to use the cyclic strain-accumulation procedure for clays.

Published by NRC Research Press


Andersen 521

Fig. 22. Cyclic DSS shear strength. Normally consolidated Drammen clay.

Fig. 23. Cyclic triaxial compression and extension shear strengths. Normally consolidated Drammen clay.

Fig. 24. Cyclic shear strain as function of number of cycles (DSS Fig. 25. Permanent pore pressure as function of number of cycles
tests with ta = 0 on Drammen clay with overconsolidation ratio (DSS tests with ta = 0 on Drammen clay with OCR = 1).
(OCR) = 1). svc’, vertical effective consolidation stress.

Published by NRC Research Press


522 Can. Geotech. J. Vol. 46, 2009

Cyclic strength diagrams for practical Fig. 26. Normalized average and cyclic shear stresses for Nf = 10 in
applications DSS tests, for different normally consolidated clays (adapted from
Andersen 2004).
Importance of databases
Diagrams with cyclic shear strength of clay, silt, and sand
that have been established from DSS and triaxial laboratory
tests are presented in this and the preceding sections. The
data are compiled from NGI files and from published litera-
ture sources that are included in the list of references at the
end of the paper. The literature sources are Lee and Seed
(1967); Mulilis et al. (1975, 1977a, 1977b); Wong et al.
(1975); Silver et al. (1976); Seed et al. (1977); Lee and
Vernese (1978); Ishihara and Takatsu (1979); Vaid and
Finn (1979); Finn (1981); Vaid and Chern (1983); Siddiqi
(1984); Sakai and Ochiai (1986); Tatsuoka et al. (1986a,
1986b, 1988); Toki et al. (1986); Tanaka et al. (1987); Hata-
naka et al. (1988); Hyodo et al. (1991, 1994, 1996); Yoshimi
et al. (1994); Evans and Zhou (1995); Vaid et al. (1999);
Koseki and Ohta (2001); Oda et al. (2001); Hosono and
Yoshimine (2004); Kokusho et al. (2004); Park and Byrne
(2004); Porcini et al. (2004); Sivathalayan and Ha (2004);
Triantafyllidis et al. (2004); Wijewickreme et al. (2004); Fig. 27. Normalized average and cyclic shear stresses for Nf = 10 in
Koseki et al. (2005); Wijewickreme and Sanin (2005); and triaxial tests, for different normally consolidated clays (adpated
Hyde et al. (2006). from Andersen 2004).
The diagrams can be used to determine parameters for
feasibility studies before site-specific data are available.
The diagrams can also be used as guides when specifying
and interpreting site-specific cyclic laboratory tests. This
will help to reduce the required number of site-specific cy-
clic tests and to evaluate their quality. Many of the diagrams
presented in the paper have already successfully served
these purposes in practical projects.
For final design, the parameters from the diagrams should
be verified by site-specific laboratory tests.
Diagrams for normally consolidated Drammen clay were
presented in the preceding section. These diagrams have the
format that the diagrams should ideally have. Additional di-
agrams for clays are presented in the first part of this sec-
tion. Thereafter, diagrams for sands and silts are presented.
A database similar to the one for Drammen clay is available
for very dense clean Baskarp sand (Andersen and Berre
1999). Baskarp sand is, however, more angular than most Diagrams for overconsolidated Drammen clay with over-
natural sands and may give strengths and moduli on the consolidation ratios of 4 and 40 can be found in Andersen
high side. The diagrams for silts and sands are less complete (2004).
than for clay due to a lack of a general systematic testing
programme similar to the ones on Drammen clay and Bas- Cyclic shear strength of sands and silts
karp sand.
The main diagrams are for normally consolidated soils Sample preparation
with moderate or no preshearing. The effect of overconsoli- High-quality samples of sands and silts that are represen-
dation and (or) preshearing can be evaluated from additional tative of the in situ conditions are difficult to obtain. In
diagrams that show the effect of these parameters. cases where ‘‘intact’’ samples have been disturbed signifi-
All the diagrams in this paper are from stress-controlled cantly, one often reconstitutes the samples, but this may not
cyclic tests with a 10 s load period on noncemented soils. give samples that are fully representative of the in situ con-
ditions.
Cyclic shear strength of clays Silver et al. (1976) and Mulilis et al. (1977a) have shown
Figures 18, 19, 22, and 23 present diagrams for normally that the cyclic shear strength of sands and silts depends on
consolidated Drammen clay. The cyclic shear strength for a the sample preparation method. Vaid et al. (1999) showed
number of other normally consolidated clays are compared that water-deposited sand closely simulated static and cyclic
with Drammen clay in Figs. 26 and 27. There is a tendency behaviour of frozen sand samples of natural alluvial and hy-
for the normalized cyclic shear strength to increase with in- draulic fill. They also found that loose samples prepared by
creasing plasticity index. moist tamping gave a weaker response than water-deposited

Published by NRC Research Press


Andersen 523

Fig. 28. Cyclic shear stress to reach g = 3.75% after 10 cycles in Fig. 30. Shear strength for 10 cycles of symmetrical cyclic loading
isotropically consolidated triaxial tests on different sands as a func- (ta = 0) in DSS tests on different normally consolidated sands as a
tion of relative density. Comparison of in situ frozen samples and function of relative density.
samples prepared by wet tamping, water deposition, and water vi-
bration.

the strength of high-quality frozen samples. A few results


from water-deposited and water-vibrated samples do not de-
viate significantly from wet tamping and frozen samples.
The tests on reconstituted samples are from various sources
and thus not on the same soil as the frozen samples. The
data in Fig. 28 show that the samples reconstituted by wet
tamping have similar cyclic shear strengths as the frozen
Fig. 29. Shear strength for 10 cycles of symmetrical cyclic loading samples. These data thus indicate that wet tamping may
(ta = 0) in DSS tests on different sands and silts with relative den- give a reasonably realistic in situ cyclic shear strength. As
sity (Dr) < 65% as a function of vertical consolidation stress. shown later, wet tamping also seems to give similar cyclic
shear strengths as ‘‘intact’’ samples (Fig. 29), whereas dry
tamping seems to give a tendency for lower strengths than
wet tamping and ‘‘intact’’ samples (Figs. 29 and 30).
A fines content of more than a few percent will make re-
constitution more difficult, especially if the fines also con-
tain clay. Use of ‘‘intact’’ samples may then be possible and
preferable, even if such specimens will also be disturbed. It
is uncertain whether the disturbance may give a cyclic
strength that is too high or too low. This will depend on the
relative density. If samples have to be reconstituted, they
should be prepared wet, by tamping or water deposition, un-
less the in situ soil has been deposited under dry conditions.
The samples should be presheared if the soil is subjected
to drained preshearing prior to and (or) during the main de-
sign event. Herein, preshearing means cyclic loading accom-
panied by drainage during or after the cyclic loading. For
large offshore gravity platforms on dense sand, preshearing
has typically been estimated to be about 400 cycles at tcy/
0
s vc = 0.04, as this has been estimated to occur during the
build-up period of the design storm or during smaller storms
prior to the design storm. In earthquake areas, the soil may
have experienced smaller earthquakes during its history. Pre-
samples. Høeg et al. (2000) showed results for the static shearing will increase the cyclic shear strength. Preshearing
shear strength of silty sand tailings, with lower static may also reduce the effect of the sample preparation
strength for samples prepared by wet tamping than for method, but this has not been systematically studied.
‘‘intact’’ samples at the same void ratio.
In several cases, however, samples prepared by wet tamp- Effect of consolidation time
ing have given similar strengths as ‘‘intact’’ samples. This is There seems to be a long-term effect of the consolidation
supported by Fig. 28, which compares the cyclic triaxial time of reconstituted samples, even for clean sand. Tatsuoka
shear strength of samples reconstituted by wet tamping with et al. (1986a) found that the cyclic strength in isotropically

Published by NRC Research Press


524 Can. Geotech. J. Vol. 46, 2009

consolidated triaxial tests on air-pluviated Toyoura sand Fig. 31. Cyclic shear stress to failure in 10 cycles in anisotropically
with no fines content increased by 15%–20% after 68 days, consolidated triaxial tests with ta = t0 compared with DSS tests.
for relative densities of both 50% and 80%. Tatsuoka et al. Different normally consolidated sands prepared by wet tamping, as
(1988) found that the cyclic shear strength of isotropically a function of relative density. t0 is about 0.25svc’ to 0.275svc’.
consolidated triaxial tests on water-vibrated Sengenyama
sand with 2.4% fines and a relative density, Dr, of 80% in-
creased by about 25% after 64 h and 40% after 68 days.
Mulilis et al. (1977b) found that the cyclic strength of iso-
tropically consolidated triaxial tests on air-pluviated Monte-
rey No. 0 sand with no fines at Dr = 50% was essentially the
same after 20 min and 1 day consolidation, whereas it in-
creased by 12% and 26% after 10 and 100 day consolidation
times, respectively. The cyclic strength in DSS tests at NGI
on air-pluviated Oosterschelde sand with no fines and Dr =
62%, however, gave the same cyclic shear strength after
16 h and 4 day consolidation.
For conventional projects, it may be impractical to repro-
duce the long-term effect in the laboratory, but the consoli-
dation stress should act overnight for samples with some
fines content. The effect of a longer consolidation time
must be evaluated by judgement and experience as refer-
enced above. The experience above is from reconstituted
samples with no preshearing, and one should be aware that samples. The samples in Fig. 29 have been presheared with
the effect is likely to be less on intact samples and on sam- 0
400 cycles at tcy/s vc = 0.05 or less.
ples with preshearing. As seen earlier, the samples that were reconstituted dry
tend to give the lowest cyclic shear strengths, and these are
Cyclic shear strength in symmetrical DSS loading
believed to underestimate the in situ shear strength. The
The cyclic DSS shear strength for 10 load cycles of dif- ‘‘intact’’ samples and the samples reconstituted wet tend to
ferent normally consolidated sands prepared by wet and dry give similar results. It is interesting to note that the cyclic
reconstitution is plotted in Fig. 30. shear strength of normally consolidated Drammen clay plots
The cyclic shear strength depends on the shear strain used in the high range of the sand and silt data. There is a ten-
to define failure, especially for densities above Dr = 65%. dency for sand to have smaller cyclic shear strength than
Failure is defined as large shear strains of g = 7.5%–15% clay, and the cyclic shear strengths are generally lower than
for most of the points in Fig. 30, but in a few cases failure the cyclic shear strength of Drammen clay. This is in line
is defined at a lower strain. This includes the data point for with the results for clays in Figs. 26 and 27, which seem to
0
Baskarp sand at tcy/s vc = 1.2, which is for g = 3% and indicate that the cyclic shear strength decreases with de-
which would plot even higher if a higher failure strain had creasing plasticity. Attempts to plot the cyclic shear strength
been used. of sand and silt as a function of grain size or the grain size
The samples with Dr < 85% in Fig. 30 have been pre- uniformity coefficient did not show any clear trend.
0
sheared with 400 cycles at tcy/s vc = 0.04 or less. The sam-
The data in Fig. 29 show that there is some tendency for
ples with Dr ‡ 85% have been presheared with 400 cycles at
0 the normalized cyclic shear strength to increase with de-
tcy/s vc between 0.075 and 0.1.
creasing consolidation stress. This tendency may be more
The plot in Fig. 30 shows that the cyclic shear strength pronounced below s vc 0
= 100 kPa, where there is no data for
increases with relative density, especially for relative den- intact samples and samples prepared by wet reconstitution.
sities above Dr = 65%. The cyclic shear strength is less de-
pendent on relative density below Dr = 65%. The plot in Dense soils (Dr > 65%) may also deviate from the trend
Fig. 30 contains results from tests on both wet and dry re- in Fig. 29 and give a greater increase in strength ratio with
0
constituted samples. The results from the tests on samples decreasing tcy,f/s vc . For instance, DSS tests on air-pluviated
with dry reconstitution generally tend to plot lower than Fraser River sand with Dr*80% gave about 25% increase in
0 0 0
those with wet reconstitution. Based on the earlier discus- tcy,f/s vc for s vc = 100 kPa compared with tests with s vc =
sion, one should disregard the results from reconstituted dry 200 kPa. Tests with Dr*40%–44% gave a smaller increase
0 0
specimens. (about 10%) in tcy,f/s vc when decreasing s vc from 200 to
Figure 30 contains data from tests with consolidation 100 kPa (Park and Byrne 2004; Wijewickreme et al. 2004).
0 Calcareous soils may also deviate from the trend in
stresses in a range of s vc = 10–710 kPa. The cyclic shear
strengths for relative densities Dr < 65% was plotted in Fig. 29. Experience from testing of calcareous silts and
Fig. 29 as function of consolidation stress to see what effect sands shows that calcareous soils tend to give similar or
0
the consolidation stress may have. Figure 29 also contains slightly higher normalized cyclic shear strength, tcy,f/s vc ,
data for normally consolidated Drammen clay and some than the noncalcareous soils in Fig. 29 at vertical consolida-
loose to medium-dense silty soils that were not included in tion stresses above 250 kPa. For lower consolidation
Fig. 30 because of the problems of assessing relative density stresses, however, laboratory tests have shown a significant
0 0
with silty soils. Some of these tests were run on ‘‘intact’’ increase in tcy,f/s vc with decreasing s vc for calcareous soils.

Published by NRC Research Press


Andersen 525

Fig. 32. Shear strength for 10 cycles of symmetrical cyclic loading (ta = 0) in DSS tests on different normally consolidated soils, as a
function of water content.

Fig. 33. Cyclic shear stress to failure in 10 cycles in anisotropically consolidated triaxial tests with ta = t0 compared with DSS tests. Dif-
ferent normally consolidated sands prepared intact and by wet tamping, as a function of relative density.

The tests on calcareous soils have generally been on intact Cyclic strength in triaxial tests
samples from a relatively shallow depth. Reasons for the The cyclic shear stress that will bring anisotropically con-
0 0
significant increase in tcy,f/s vc with decreasing s vc may be solidated triaxial samples to failure in 10 cycles is compared
that calcareous soil has high angularity, but may also be with DSS tests in Fig. 31. There are only three triaxial tests.
due to cementation or overconsolidation, even if they are One should therefore be cautious about drawing general
not characterized as cemented or overconsolidated. conclusions. However, the normalized cyclic shear stress to

Published by NRC Research Press


526 Can. Geotech. J. Vol. 46, 2009

0
failure, tcy,f/s vc , is about 75% higher for the triaxial tests Fig. 34. Cyclic shear stress to failure as a function of number of
with ta = t0 than for the DSS tests with ta = 0. cycles. DSS tests with symmetrical cyclic loading (ta = 0) on nor-
mally consolidated sands and silts: Dr-range 40%–100%; consoli-
Cyclic shear strength and water content dation stress range 85–710 kPa. (a) tcyf/svc’ versus Nf;
The cyclic shear stress that gives failure in 10 cycles in (b) normalized to cyclic shear stress to failure in 10 cycles.
DSS tests with symmetrical cyclic loading is plotted as func-
tion of the water content in Fig. 32. The samples prepared
by dry reconstitution tend to give the lowest cyclic shear
strength, as in previous figures.
For the intact samples and the samples prepared by wet
0
reconstitution, the normalized cyclic shear strength, tcy,f/s vc ,
increases with decreasing water content, especially as the
water content decreases below a water content of about
25%. The effect of the water content is less significant
above 25%.
0
Some of the scatter in tcy,f/s vc below a water content of
25% is due to differences in consolidation stress, with a ten-
0
dency for tcy,f/s vc to increase with decreasing consolidation
stress. The effect of consolidation stress was not large in the
tests on sand with a relative density below *65% (Fig. 29),
but as mentioned before the effect of the consolidation stress
may be more pronounced at higher relative densities, i.e., at
lower water content. Some scatter may also be due to uncer-
tainty in the water content for some of the tests, as it was
not clear in all the literature cases whether the reported
water content was before or after consolidation. The water
content after consolidation was used when it was available.
The cyclic shear stress that will bring anisotropically con-
solidated triaxial samples to failure in 10 cycles is compared
with DSS tests in Fig. 33. For the samples prepared by wet
reconstitution, the normalized cyclic shear stress to failure,
0
tcy,f/s vc , is higher in the triaxial tests with ta = t0 than in
the DSS tests with ta = 0. This is the same tendency seen
0
when tcy,f/s vc was plotted as a function of relative density
in Fig. 31. The two triaxial tests on intact silt–clay samples
show the opposite trend, but one should be cautious about
drawing general conclusions with the limited number of tri-
axial tests available.

Effect of number of cycles


The cyclic shear stress that will bring DSS samples and
anisotropically consolidated triaxial samples to failure for a
number of cycles different than 10 are plotted in Figs. 34
and 35. Figs. 34a and 35a show the cyclic shear stress to
failure normalized to the vertical effective consolidation
0
stress, tcy,f/s vc . Figs. 34b and 35b show the cyclic shear
stress to failure relative to the cyclic shear stress to failure consolidated soils indicate that they possibly follow the
in 10 cycles. The diagrams contain results from both intact trend in Figs. 34 and 35.
samples and wet and dry sample preparation. Normally con-
solidated Drammen clay is included for reference. Effect of average shear stress
The data show that in DSS tests the cyclic shear stress to The diagrams in Figs. 29 to 35 are for DSS tests with ta =
0
give failure in three cycles is about 20% to 40% higher than 0 and triaxial tests with ta = t0, where t0/s vc *0.25–0.275.
for 10 cycles. The difference between three and 10 cycles to The cyclic shear strength is tf,cy = ta,f + tcy,f, and the cyclic
failure shows more scatter for the triaxial tests. There are no shear strength depends on the average shear stress. To fully
systematic differences between intact samples and samples define the cyclic shear strength, one therefore needs dia-
with wet and dry sample preparation. Drammen clay tends to grams of the type presented for clays in Figs. 18, 19, 26,
be somewhat less sensitivite to the number of cycles to failure and 27. Such diagrams are given for the very dense Baskarp
for DSS tests, but is closer to the average for triaxial tests. sand in Andersen and Berre (1999) for both undrained and
The data in Figs. 34 and 35 are for normally consolidated drained Dta.
and close to normally consolidated soils. Less data are avail- In triaxial tests with drained average shear stress, the cy-
able for overconsolidated soils, but the data for a few over- clic shear strength will depend on the stress path for the

Published by NRC Research Press


Andersen 527

Fig. 35. Cyclic shear stress to failure as function of number of Fig. 36. Cyclic shear stress to failure in 10 cycles in isotropically
cycles. Anisotropically consolidated triaxial tests with ta = t0 = consolidated triaxial tests compared with triaxial tests with ta = t0
0.21–0.3svc’ on normally consolidated sands and silts: Dr-range and DSS tests as a function of Dr. Different normally consolidated
35%–96%; consolidation stress range 100–710 kPa. (a) tcyf/svc’ sands prepared by wet tamping.
versus Nf; (b) normalized to cyclic shear stress to failure in 10 cy-
cles.

have sometimes been used to estimate the cyclic shear


strength for DSS stress conditions, even if the isotropically
consolidated triaxial tests do not model a representative
stress path. Isotropically consolidated triaxial tests generally
0
show higher cyclic shear stresses at failure, tcy,f/s vc , than
both the DSS and anisotropically consolidated triaxial tests
when they are normalized to the vertical effective consolida-
tion stress (Fig. 36). To correct for this, one may assume
that the strengths of DSS and isotropically consolidated tri-
axial tests will be the same when normalized to the octahe-
dral effective consolidation stress. A comparison of
available data show that normalization to the octahedral
consolidation stress appears to give reasonable estimates of
the cyclic DSS strength for Dr < 65%. However, the hori-
zontal consolidation stress needed to calculate the octahedral
stress in the DSS test is uncertain. At higher relative den-
sities, normalization to the octahedral consolidation stress
seems to overestimate the cyclic DSS strength.
One should generally be cautious about using isotropically
consolidated triaxial tests, as the stress conditions in this test
are normally not representative of the stress conditions be-
neath a structure.

Effect of preshearing
average load, i.e., whether Dta is applied by changing the Sand deposits are often subjected to preshearing, i.e.,
vertical or the horizontal normal stress. small cyclic shear stresses accompanied by drainage, prior
In dense soils that dilate under a change in average shear to the main design event, as discussed earlier. Preshearing
stress, the cyclic shear strength will generally be higher if may influence the undrained cyclic shear strength of a sand.
Dta is applied undrained than if it is applied drained. In Figure 37 shows how the undrained cyclic shear strength for
10 load cycles depends on the normalized cyclic shear
soils that contract under a change in average shear stress, 0
stress, tcy/s vc , and the number of cycles during preshearing.
the cyclic shear strength may be lowest when Dta is applied
undrained. The figure contains DSS, shaking table, and triaxial tests
subjected to various degrees of preshearing.
In DSS tests, the preshearing may
Comparison between DSS and isotropically consolidated
triaxial tests  improve the seating between the sand and the horizontal
The DSS test represents the best simulation of the in situ end plates
stress conditions for many situations. When DSS tests have  level out stress concentrations from the consolidation
not been available, isotropically consolidated triaxial tests  increase the horizontal effective stress

Published by NRC Research Press


528 Can. Geotech. J. Vol. 46, 2009

Fig. 37. Effect of preshearing on undrained cyclic shear stress at a cyclic shear stress of half the static shear strength gives an
failure in triaxial and DSS tests. increased resistance to subsequent undrained cyclic loading
in normally consolidated clay, but a reduced resistance in
clay with an overconsolidation ratio of 4 (e.g., Andersen
1988). Similar effects of overconsolidation may also exist
for cohesionless soil, but no data have been available for
overconsolidated sand.

Effect of overconsolidation
The in situ soil may be overconsolidated due to removal
of previous overburden, variation in the weight of the struc-
ture or preloading by temporary weight or underpressure.
Overconsolidation will increase the horizontal effective nor-
mal stress, increase the relative density, and possibly change
the structure. These changes will tend to increase the cyclic
resistance of the soil.
The effect of overconsolidation on the cyclic shear stress
0
at failure, tcy,f/s vc , was compiled for different cohesionless
soils in Fig. 38. Both loose and dense soils are included.
Drammen clay and an empirical equation from Ishihara and
Takatsu (1979) are included as references. The equation
from Ishihara and Takatsu (1979) is applied for two condi-
tions: (i) K0-value of 0.5, 0.7, 1.0, and 1.35 for OCR = 1, 2,
4, and 8, respectively; and (ii) K0 = 1.0 for all OCRs. The
first condition simulates one-dimensional consolidation,
such as in the DSS test. The second condition represents iso-
tropically consolidated triaxial conditions where K0 does not
change because of overconsolidation.
The data in Fig. 38 show that
 Overconsolidation increases the resistance to cyclic load-
ing significantly.
 The effect of overconsolidation on the cyclic shear stress
 change the soil structure at failure is typically about 50% higher in DSS than in
The volume reduction is generally small during preshear- triaxial tests. One reason for this is likely the increase in
ing, and the increase in density is by far not large enough to K0 with OCR in the DSS tests. If the triaxial tests are run
explain the increased cyclic resistance. One may expect pre- by adjusting K0 as function of OCR, one would expect
shearing to have less effect in triaxial tests than in DSS the same effect of OCR as in the DSS tests.
tests, as seating and stress concentrations are likely to be  The increase in the cyclic DSS strength for the noncohe-
less important and the horizontal stress is kept constant in sive soils is close to or slightly smaller than the increase
triaxial tests. measured on Drammen clay.
The data in Fig. 37 show that
 The empirical equations from Ishihara and Takatsu
 the cyclic shear stress at failure increases with increasing (1979) give a somewhat greater effect of overconsolida-
preshearing, i.e., both with increasing cyclic shear stress tion than the data points, but also indicate a difference
and increasing number of cycles during preshearing between DSS and isotropically consolidated triaxial tests.
 preshearing has an important effect for both low and high
relative density; no clear trend is observed as function of Effect of load period
Dr All results presented in this paper are from tests with a
 there are less data with preshearing in triaxial than in 10 s load period. There is scatter in the data, but experience
DSS tests, and the data do not show clear differences be- from the testing of several offshore clays is that one 100 s
tween triaxial and DSS tests load cycle has the same effect as 1.5 to 5 cycles with a 10 s
0
 preshearing with 400 cycles at tcy/s vc = 0.04 may give a load period. This observation is for clays with a plasticity
cyclic shear strength increase between about 5% and 25% index, Ip, in the range of 40%–100%. There is not a clear
The effect of preshearing may be the opposite of what is trend, but the effect of a longer load period seems to be
presented above if the preshearing causes large shear strains highest for tests with a high Ip.
that may break down the structure, e.g., Oda et al. (2001) The tests on quick clay in the subsequent section show
and Wijewickreme and Sanin (2005). that one 10 s cycle seems to have the same effect as about
The data above are for normally consolidated soil. Pre- eight 1 s cycles. Cyclic tests on sand seem to indicate no
shearing may give a less positive effect on overconsolidated significant effect of load period (Lee and Vernese 1978;
soil. Cyclic tests on Drammen clay show that preshearing at Tatsuoka et al. 1986a). No systematic data has been found

Published by NRC Research Press


Andersen 529

Fig. 38. Effect of overconsolidation on cyclic shear stress at failure Fig. 39. Simplified stress conditions in infinite slope. sn’, effective
for 10 load cycles in DSS tests with symmetrical loading (ta = 0) stress normal to the failure surface.
and in triaxial tests. CIU, isotropically consolidated triaxial test.

The capacity of piles under vertical cyclic loading can be


calculated by modelling the pile–soil interface with nonlinear
‘‘t–z’’ springs (Karlsrud and Nadim 1990), where t and z rep-
resent mobilized skin friction and vertical displacement, re-
spectively, with cyclic and permanent components
determined through strain contour diagrams of the type men-
tioned in the section ‘‘Cyclic data for displacement analyses’’
above.

Clay slopes under earthquake loading


for silt, but it seems reasonable to expect an effect between Special considerations for quick clay and slopes
clay and sand, depending on the silt, sand, and clay content. subjected to earthquake loading
For slopes subjected to earthquake loading, the stress con-
Cyclic data for displacement analyses ditions and the failure mechanism require special considera-
Data for calculation of cyclic displacements, equivalent tions. A permanent slope will consolidate under a shear
soil stiffnesses for structural dynamic analyses, permanent stress from its own weight, tc, as illustrated in Fig. 39. For
displacements due to cyclically induced shear strains, and relatively steep slopes, this average shear stress may be of
permanent displacement due to dissipation of cyclically in- about the same magnitude as the cyclic shear stress due to
duced pore pressure are presented in Andersen (2004) for the earthquake. The earthquake has a cyclic load period
clay and in Andersen et al. (1994) and Andersen and Berre about one-tenth of offshore wave loading.
(1999) for sand. This section presents preliminary results from an on-going
The data are presented in the same format as for cyclic laboratory testing programme where such conditions are
stresses to failure in Figs. 18 and 19, but with contours of modelled. The tests were run on high-quality block samples
permanent and cyclic shear strains and permanent pore pres- of a Norwegian quick clay with a sensitivity of more than
sure for a given number of cycles instead of contours of the 75. The samples were normally consolidated with water con-
number of cycles to failure. tent of about 39%, clay content of about 40%, plasticity in-
Procedures to calculate the different displacement compo- dex of about 10%–12%, and an apparent overconsolidation
nents are proposed in Andersen (1991) and Andersen and ratio due to secondary consolidation less than 1.5. The pre-
Høeg (1991). consolidation stress was obtained from constant rate of
strain oedometer tests.
Capacity calculation Quick clay was used for the testing because new Norwe-
The capacity of a foundation under combined static and gian regulations may require slopes to be checked for earth-
cyclic loading can be calculated by limiting equilibrium quake loading, and some slopes in Norway contain quick
analyses with the procedure proposed by Andersen and clay. There is little experience with the cyclic behaviour of
Lauritzsen (1988). The cyclic shear strength of the clay is quick clay and it was believed that quick clay may be more
determined from diagrams such as in Figs. 22 and 23. The vulnerable to cyclic loading than ordinary clays.
cyclic shear strength at a given point on the failure surface The purpose of the testing was (i) to check cyclic behav-
(see Fig. 15) is interpolated between the strengths from iour under the stress conditions in a slope and (ii) to study if
compression, DSS, and extension tests. The procedure ac- quick clay behaves differently from ordinary clays.
counts for the redistribution of average soil stresses during
cyclic loading and determines whether the failure mode will Cyclic behaviour of quick clay versus ordinary clays
be large cyclic displacements, large average displacements The behaviour of the quick clay under cyclic loading is
or a combination of the two. The procedure is based on the compared with normally consolidated Drammen clay in
assumption that the combination of average and cyclic shear Fig. 40. The data in Fig. 40 are for DSS tests with symmet-
strains is the same along the potential failure surface (strain rical stress-controlled 10 s load cycles. The number of
compatibility) and on the condition that the average shear cycles to failure is slightly lower for the quick clay than for
stresses along the potential failure surface is in equilibrium Drammen clay. The relatively small difference may be ex-
with the average loads. plained by differences from one soil to another. In particu-

Published by NRC Research Press


530 Can. Geotech. J. Vol. 46, 2009

Fig. 40. Number of cycles to failure for quick clay compared with Fig. 42. Stress–strain behaviour in monotonic, cyclic, and post-
Drammen clay. DSS tests with symmetrical loading (ta = 0) and cyclic monotonic DSS tests with ta = tc = 20.8 kPa = 0.16svc’. th,
10 s load period. St, sensitivity. horizontal shear stress.

Fig. 41. Undrained static DSS shear strength of the quick clay as a
function of normalized consolidation shear stress, tc/svc’.

otonic and cyclic testing. If this strength increase is


neglected, the stability of a slope under additional undrained
loading may be significantly underestimated. On the other
hand, a laboratory test is normally run to failure in less than
2 h, and the slope will remain undrained for a much longer
period than this after additional loading. The undrained
shear strength may then be reduced by about 20%, accord-
ing to Fig. 20. Monotonic tests on the quick clay sheared
with different strain rates agree well with the diagram in
Fig. 20.

Effect of cyclic loading on static shear strength


A series of tests were performed to see the effect of cyclic
lar, the quick clay had somewhat smaller clay content (40%) loading on the static shear strength of a slope. Examples of
0
than Drammen clay (45%–55%). Experience has shown that cyclic tests consolidated with tc/s vc = 0.16, corresponding to
the cyclic shear strength may tend to decrease with decreas- a slope of about 108, are presented in Fig. 42. The figure
ing plasticity (e.g., Figs. 26 and 27). The data thus show that shows one reference monotonic test and two cyclic tests
the quick clay tested seems to behave similarly to other that were run with monotonic loading to failure after cy-
clays under cyclic loading. cling. The monotonic tests were run strain-controlled with a
rate of shear strain of *4.5%/h.
Effect of load period The reason why the cyclic stress–strain curves go beyond
The tests in Fig. 40 were run with a load period of 10 s. the monotonic stress–strain curve is that the cyclic tests are
Earthquakes have a period of about 1 s, and preliminary run stress-controlled and that the rate of strain is signifi-
tests on the quick clay with different load periods indicate cantly higher in the cyclic tests when they develop large
that the number of cycles to failure increases about 8 times strains than the rate of strain in the monotonic tests.
when the load period is reduced from 10 to 1 s. The cyclic loading was stopped when a permanent shear
strain of gp = 2% was reached in the first cyclic test (DSS8)
Effect of shear stress during consolidation and when gp = 12% was reached in the other cyclic test
The undrained shear strength will increase if the clay is (DSS6). The results in Fig. 42 show that the monotonic
consolidated under a shear stress, as illustrated in Fig. 41. peak shear strength is reduced by the cyclic loading and
The effect may be smaller in overconsolidated clay and that the post-cyclic monotonic stress–strain curves rapidly
may depend on plasticity. join the virgin monotonic stress–strain curve. Differences in
The data in Fig. 41 illustrates the importance of consoli- the monotonic curves are believed to be due to soil variabil-
dating the clay under the appropriate stresses, both for mon- ity. The series of tests consolidated under other shear

Published by NRC Research Press


Andersen 531

stresses show better agreement. It thus appears that the post- Fig. 43. Development of shear strain in undrained DSS tests with
cyclic static shear strength is governed by the virgin mono- constant shear stress.
tonic stress–strain curve and the permanent shear strain that
is developed during cyclic loading. Similar results have been
found on Drammen clay (Andersen 1988), but the behaviour
is even more evident for the quick clay due to the more pro-
nounced strain-softening.
Tests with symmetrical cyclic loading (ta = tc = 0) show
a somewhat different picture, where the post-cyclic mono-
tonic stress–strain curve does not reach the monotonic
stress–strain curve. This more severe effect of cyclic loading
was seen both in tests on quick clay and Drammen clay
(Andersen 1988).

Effect of cyclic loading on undrained creep


A slope subjected to an earthquake will experience both
cyclic and permanent shear strains and displacements. How-
ever, the failure mode is not likely to be large cyclic shear
strains and displacements because of the relatively signifi- be taken into consideration when applying the shear strength
cant average shear stress in a slope. The cyclic loading will discussed above.
instead cause large permanent shear strains and displace-
ments, as in the tests in Fig. 42. The failure is not likely to
occur during the peak load either, because the duration of Verification of calculated foundation
the peak load is not long enough to accelerate the soil capacity by means of model tests
mass. The critical mechanism is therefore likely to be devel- The use of laboratory test data, as presented above, to cal-
opment of large permanent shear strains and displacements culate the foundation capacity of structures subjected to cy-
during or after the earthquake. The critical period may be clic loading was verified by predicting or back-calculating
some time after the earthquake, before the excess pore pres- various series of model tests. The calculation method is
sure generated by the cyclic loading has dissipated. During briefly described earlier in the paper.
this period, the clay will creep under undrained conditions The model tests included:
and a delayed failure may occur. This is simulated in the
DSS tests and illustrated by an example in Fig. 43.  Five 1g laboratory model tests of an offshore gravity-
The test in Fig. 43 was consolidated with an average base structure with monotonic and cyclic loading on a
shear stress of 62% of the undrained static shear strength, soft clay (Andersen et al. 1989).
corresponding to a slope of about 138. The specimen was  Twelve 1g laboratory model tests with monotonic and
cycled to a permanent shear strain of 5% and left with this cyclic loading on an offshore tripod gravity platform
average shear stress under undrained conditions. It can be (Aas and Andersen 1992).
seen that the specimen develops shear strains that accelerate  One monotonic and one cyclic centrifuge test of a gravity-
and failure occurs after 136 min. base structure similar to the Ekofisk oil storage tank (An-
A reference test is included in Fig. 43 to verify that the dersen et al. 1994)
test was not unstable after consolidation and that the creep  Two series of large-scale field tests with monotonic and
failure was induced by the cyclic loading. The reference cyclic loading of offshore suction anchors in clays. One
test was run by just closing the drainage valves after the series of tests was run with a load inclination of 108
same consolidation period as in the cyclic test. The refer- with the vertical to simulate anchors for a TLP (Andersen
ence test did not develop noticeable shear strains. et al. 1993; Dyvik et al. 1993). The other series was run
The failure mode for a slope subjected to earthquake with a load inclination of 108 with the horizontal to simu-
loading is thus expected to be delayed undrained creep. late anchors with more horizontal loading (Keaveny et al.
Based on the results produced so far, the slope may be ana- 1994).
lysed by first running a dynamic analysis to determine the The comparison of calculated and measured capacities are
permanent shear strain due to the design earthquake. The presented in the references for the various model tests
post-cyclic shear strength may then be determined as the above. The calculated bearing capacity, type of failure
shear stress on the monotonic stress–strain curve at a shear surface, failure surface location, and failure mode (large per-
strain equal to the calculated permanent shear strain. This manent displacements, large cyclic displacements or a com-
shear strength should be reduced by 15%–25% to account bination of the two) agreed well with the measurements for
for (i) the post-cyclic stress–strain curve reaching the virgin all the model tests.
curve at a somewhat larger strain than the permanent strain As an example, some details are given for the last of the
developed during the earthquake and (ii) the time to failure model test series referenced above. The model geometry and
being significantly longer than in the monotonic laboratory the soil are summarized in Fig. 44. The model after it was
test (see Fig. 20). brought to failure is shown in Fig. 45.
In a quick clay slope there will be progressive failure Predicted and measured capacities of the monotonic and
mechanisms (e.g., Andresen and Jostad 2004) and this must the three cyclic TLP tests are compared in Table 1. The

Published by NRC Research Press


532 Can. Geotech. J. Vol. 46, 2009

Fig. 44. Geometry and key soil data for the large-scale field model Fig. 46. Predicted and observed failure surfaces in TLP model tests.
tests of offshore suction anchors. P, applied load. (a) Tests 1, 2, and 3; (b) test 4.

Fig. 45. TLP model after failure (photo by Rune Dyvik, NGI).

important to determine the cyclic shear strength under repre-


sentative stress conditions. In practice, this can be done by
DSS and triaxial tests consolidated to the appropriate in situ
stresses prior to cyclic loading and run under representative
combinations of average and cyclic shear stresses.
A convenient way to present the results from the cyclic
tests is to prepare diagrams where the number of cycles to
Table 1. Predicted and measured failure loads for the TLP failure is plotted as a function of average and cyclic shear
field-model tests. stresses. These diagrams should also contain the failure
mode, i.e., the combination of permanent and cyclic shear
Test No. Test type Predicted/measured failure load strains at failure. The diagrams will thus contain the infor-
1 Monotonic 1.00 mation required to calculate the capacity of a structure under
2 Cyclic 1.05 cyclic loading by a limiting equilibrium method that ac-
3 Cyclic 1.06 counts for (i) strain compatibility and (ii) equilibrium be-
4 Cyclic 1.01 tween the weight of the structure and the average shear
stresses in the soil.
agreement is very good. The predicted and observed failure The paper presents diagrams that contain the cyclic shear
surfaces are shown in Fig. 46. The agreement is also gener- strength of clay, sand, and silt that can be used for practical
ally good, but there are some differences. However, back- design. These diagrams are useful for feasibility studies be-
calculations with the observed failure surfaces gave insignif- fore site-specific cyclic data become available and as a
icant differences in the calculated capacities. The reason for guide when specifying and interpreting site-specific cyclic
the difference in failure surface between test 4 and the other laboratory tests.
tests is that test 4 was subjected to a greater moment. Some cases, such as slopes under earthquake loading,
may require special considerations and the paper presents
some data on cyclic testing of a quick clay under such con-
Summary and conclusions
ditions. One possible failure mode for such cases is un-
Cyclic loading can be very significant for the foundation drained creep initiated by an earthquake. The post-cyclic
design of structures — offshore, along the coast, and on monotonic shear strength may also be reduced by the earth-
land — and for the stability of slopes. quake loading, making the slope vulnerable to increased
The soil elements beneath a structure subjected to cyclic loads or erosion in the period before the earthquake-induced
loading follow different stress paths. The behaviour under excess pore pressures have dissipated.
cyclic loading depends strongly on the stress path and it is Comparisons between calculations and model tests indi-

Published by NRC Research Press


Andersen 533

cate that the foundation capacity under cyclic loading can be Andersen, K.H., and Lauritzsen, R. 1988. Bearing capacity for
reliably determined on the basis of cyclic shear strength de- foundation with cyclic loads. Journal of Geotechnical Engineer-
termined in laboratory tests. ing, 114(5): 540–555. doi:10.1061/(ASCE)0733-9410(1988)
114:5(540).
Acknowledgements Andersen, K.H., Kleven, A., and Heien, D. 1988. Cyclic soil data
for design of gravity structures. Journal of Geotechnical Engi-
The results presented in this paper are based on informa- neering, 114(5): 517–539. doi:10.1061/(ASCE)0733-9410(1988)
tion from joint industry-sponsored research projects, re- 114:5(517).
search projects funded by the Research Council of Norway, Andersen, K.H., Dyvik, R., Lauritzsen, R., Heien, D., Hårvik, L.,
and consulting projects. Numerous colleagues at NGI have and Amundsen, T. 1989. Model tests of offshore platforms. II.
contributed to the results, both in performing and interpret- Interpretation. Journal of Geotechnical Engineering, 115(11):
ing laboratory tests and model tests and by developing theo- 1550–1568. doi:10.1061/(ASCE)0733-9410(1989)115:11(1550).
retical models. This co-operation is greatly appreciated. Andersen, K.H., Hansteen, O.E., and Gutierrez, M. 1991. Bearing
Special thanks are extended to Dr. Fritz Nowacki, NGI, for capacity, displacements, stiffness and hysteretic damping of Stor-
fruitful discussions through many years and projects; ebælt bridge piers under ice loading. In Proceedings of the Semi-
Mr. Jan Lampe, NGI, for his high-quality and careful DSS nar on Design of Exposed Bridge Piers, Copenhagen, Denmark,
testing; Dr. Amir Kaynia, NGI, for discussions on slope be- 22 January 1991. Danish Society of Hydraulic Engineering, Co-
haviour under earthquake loading; and Dr. Suzanne Lacasse, penhagen, Denmark. (Also published in Publikasjon - Norges
NGI, for reviewing and commenting on the paper. Geotekniske Institutt, Vol. 199.)
Andersen, K.H., Dyvik, R., Kikuchi, Y., and Skomedal, E. 1992.
Clay behaviour under irregular cyclic loading. In Proceedings
References of the International Conference on the Behaviour of Offshore
Structures, London, 7–10 July 1992. BPP Technical Services
Aas, P.M., and Andersen, K.H. 1992. Skirted foundations for off- Ltd., London. Vol. 2. pp. 937–950.
shore structures. In Proceedings of the 9th Offshore South East Andersen, K.H., Dyvik, R., Schroder, K., Hansteen, O.E., and By-
Asia Conference and Exhibition, Singapore, 1–4 December sveen, S. 1993. Field tests of anchors in clay. II: Predictions and
1992. Singapore Exhibition Services Pte Ltd., Singapore. interpretation. Journal of Geotechnical Engineering, 119(10):
pp. 305–312. (Also published in Publikasjon - Norges Geote- 1532–1549. doi:10.1061/(ASCE)0733-9410(1993)119:10(1532).
kniske Institutt, Vol. 190, pp. 1–8.) Andersen, K.H., Allard, M.A., and Hermstad, J. 1994. Centrifuge
Andersen, K.H. 1976. Behaviour of clay subjected to undrained model tests of a gravity platform on very dense sand. II: Interpre-
cyclic loading. In Proceedings of the International Conference tation. In Proceedings of the 7th International Conference on Be-
havior of Offshore Structures, BOSS’94, Cambridge, Mass., 12–
on the Behaviour of Offshore Structures, BOSS’76, Trondheim,
15 July 1994. Pergamon Press, New York. Vol. 1. pp. 255–282.
Norway, 2–5 August 1976. Norwegian Institute of Technology,
Andersen, K.H., Murff, J.D., Randolph, M.F., Clukey, E., Erbrich,
Trondheim, Norway. Vol. 1. pp. 392–403. C., and Jostad, H.P. Hansen, B. Aubeny, C., Sharma, P., and Su-
Andersen, K.H. 1988. Properties of soft clay under static and cyclic pachawarote, C. 2005. Suction anchors for deepwater applica-
loading. In Proceedings of the International Conference on Engi- tions. In Proceedings of the International Symposium on
neering Problems of Regional Soils, Beijing, China, 11–15 Au- Frontiers in Offshore Geotechniques (ISFOG), Perth, Australia,
gust 1988. Edited by Chinese Institution of Soil Mechanics and 19–21 September 2005. Edited by S. Gourvenec and M. Cas-
Foundation Engineering. International Academic Publishers, sidy. Taylor & Francis Group, London. pp. 1–30.
Beijing, China. pp. 7–26. (Also published in Publikasjon - Andresen, L., and Jostad, H.P. 2004. Analyses of progressive fail-
Norges Geotekniske Institutt, Vol. 176, pp. 1–20.) ure in long natural slopes. In Numerical Models in Geomecha-
Andersen, K.H. 1991. Foundation design of offshore gravity struc- nics: Proceedings of NUMOG IX, Ottawa, Ont., 25–27 August
tures. In Cyclic Loading of Soils. From Theory to Design. Edi- 2004. Edited by G.N. Pande and S. Pietruszczak. Taylor & Fran-
ted by M.P. O’Reilly and S.F. Brown. Blackie and Son Ltd., cis Group, London. pp. 603–608.
Glasgow and London. pp. 122–173. (Also published in Publikas- Bjerrum, L. 1973. Geotechnical problems involved in foundations
jon - Norges Geotekniske Institutt, Vol. 185.) of structures in the North Sea. Géotechnique, 23(3): 319–358.
Andersen, K.H. 2004. Cyclic clay data for foundation design of Christophersen, H.P., Bysveen, S., and Støve, O.J. 1992. Innovative
structures subjected to wave loading. In Proceedings of the foundation design systems selected for the Snorre field develop-
International Conference on Cyclic Behaviour of Soils and Li- ment. In Proceedings of the 6th International Conference on
quefaction Phenomena, CBS04, Bochum, Germany, 31 March – Behaviour of Offshore Structures, Imperial College, London, 7–
2 April 2004. Edited by T. Triantafyllidis. CRC Press, Taylor & 10 July 1992. Bentham Press, London. pp. 82–94.
Francis Group, London. pp. 371–387. Clausen, C.J.F., DiBiagio, E., Duncan, J.M., and Andersen, K.H.
Andersen, K.H., and Berre, T. 1999. Behaviour of a dense sand un- 1975. Observed behavior of the Ekofisk oil storage tank founda-
der monotonic and cyclic loading. In Proceedings of the 12th tion. In Proceedings of the 7th Offshore Technology Conference,
ECSMGE, Geotechnical Engineering for Transportation Infra- Houston, Tex., 5–8 May 1975. Society of Petroleum Engineers
structure, Amsterdam, the Netherlands, 7–10 June 1999. A.A. (SPE), Richardson, Tex. Paper OTC-2373, Vol. 3. pp. 399–413.
Balkema, Rotterdam, the Netherlands. Vol. 2. pp. 667–676. Dyvik, R., Andersen, K.H., Madshus, C., and Amundsen, T. 1989.
Andersen, K.H., and Høeg, K. 1991. Deformations of soils and dis- Model tests of gravity platforms. I. Description. Journal of Geo-
placements of structures subjected to combined static and cyclic technical Engineering, 115(11): 1532–1549.
loads. In Proceedings of the 10th European Conference on Soil Dyvik, R., Andersen, K.H., Hansen, S.B., and Christophersen, H.P.
Mechanics and Foundation Engineering, Florence, Italy, 26– 1993. Field tests of anchors in clay. I. Description. Journal of
30 May 1991. Associazione Geotecnica Italiana, Rome. Vol 4. Geotechnical Engineering, 119(10): 1515–1531. doi:10.1061/
pp. 1147–1158. (ASCE)0733-9410(1993)119:10(1515).

Published by NRC Research Press


534 Can. Geotech. J. Vol. 46, 2009

Evans, M.D., and Zhou, S. 1995. Liquefaction behaviour of sand- Lunne, T., and Andersen, K.H. 2007. Soft clay shear strength para-
gravel composites. Journal of Geotechnical Engineering, 121(3): meters for deepwater geotechnical design. In Proceedings of the
287–299. doi:10.1061/(ASCE)0733-9410(1995)121:3(287). 6th International Conference: Society for Underwater Technol-
Finn, W.D.L. 1981. Liquefaction potential: Developments since ogy, Offshore Site Investigation and Geotechnics (SUT-OSIG),
1976. In Proceedings of the International Conference on Recent London, 11–13 September 2007. Society for Underwater Tech-
Advances in Geotechnical Earthquake Engineering and Soil Dy- nology, London. pp. 151–176.
namics, St. Louis, Mo., 26 April – 3 May 1981. University of Mulilis, J.P., Chan, C.K., and Seed, H.B. 1975. The effects of
Missouri–Rolla, Rolla, Mo. Vol. 2. pp. 655–681. method of sample preparation on the cyclic stress-strain beha-
Hatanaka, M., Suzuki, Y., Kawasaki, T., and Endo, M. 1988. Cyc- vior of sands. College of Engineering., University of California,
lic undrained shear properties of high quality undisturbed Tokyo Berkeley, Calif. Report EERC 75-18.
gravel. Soils and Foundations, 28(4): 57–68. Mulilis, J.P., Seed, H.B., Chan, C.K., Mitchell, J.K., and Arulanan-
Høeg, K., Dyvik, R., and Sandbækken, G. 2000. Strength of undis- dan, K. 1977a. Effects of sample preparation on sand liquefaction.
turbed versus reconstituted silt and silty sand specimens. Journal Journal of Geotechnical Engineering, ASCE, 103(2): 91–108.
of Geotechnical and Geoenvironmental Engineering, 126(7): Mulilis, J.P., Mori, K., Seed, H.B., and Chan, C.K. 1977b. Resis-
606–617. doi:10.1061/(ASCE)1090-0241(2000)126:7(606). tance to liquefaction due to sustained pressure. Journal of Geo-
Hosono, Y., and Yoshimine, M. 2004. Liquefaction of sand in sim- technical Engineering, 103(7): 793–797.
ple shear condition. In Proceedings of the International Confer- Oda, M., Kawamoto, K., Suzuki, K., Fujimori, H., and Sato, M.
ence on Cyclic Behaviour of Soils and Liquefaction Phenomena, 2001. Micro-structural interpretation in saturated granular soils
CBS04, Bochum, Germany, 31 March – 2 April 2004. Edited by under cyclic loading. Journal of Geotechnical and Geoenviron-
T. Triantafyllidis. CRC Press, Taylor & Francis Group, London. mental Engineering, 127(5): 416–423. doi:10.1061/(ASCE)
pp. 129–136. 1090-0241(2001)127:5(416).
Hyde, A.F.L., Higuchi, T., and Yasuhara, K. 2006. Liquefaction, Park, S.S., and Byrne, P.M. 2004. Numerical modeling of soil lique-
cyclic mobility and failure of silt. Journal of Geotechnical Engi- faction at slope site. In Proceedings of the International Confer-
neering, 132(6): 716–735. ence on Cyclic Behaviour of Soils and Liquefaction Phenomena,
Hyodo, M., Murata, H., Yasufuku, N., and Fujii, T. 1991. Un- CBS04, Bochum, Germany, 31 March – 2 April 2004. Edited by
drained cyclic shear strength and residual strain of saturated T. Triantafyllidis. CRC Press, Taylor & Francis Group, London.
sand by cyclic triaxial tests. Soils and Foundations, 31(3): 60–76. pp. 571–580.
Hyodo, M., Tanimizu, H., Yasufuku, N., and Murata, H. 1994. Un- Porcini, D., Cicciu, G., and Ghionna, V.N. 2004. Laboratory inves-
drained cyclic and monotonic triaxial behaviour of saturated tigation of the undrained cyclic behaviour of a natural coarse
loose sand. Soils and Foundations, 34(1): 19–34. sand from unditurbed and reconstituted samples. In Proceedings
Hyodo, M., Aramaki, N., Itoh, M., and Hyde, A.F.L. 1996. Cyclic of the International Conference on Cyclic Behaviour of Soils
strength and deformation of crushable carbonate sand. Journal of and Liquefaction Phenomena, CBS04, Bochum, Germany,
Soil Dynamics and Earthquake Engineering, 15(5): 331–336. 31 March – 2 April 2004. Edited by T. Triantafyllidis. CRC
doi:10.1016/0267-7261(96)00003-6. Press, Taylor & Francis Group, London. pp. 187–192.
Ishihara, K., and Takatsu, H. 1979. Effects of overconsolidation Sakai, A., and Ochiai, H. 1986. Effects of initial static shear stres-
and K0 conditions on the liquefaction characteristics of sands. ses on the liquefaction of sand. Memoirs of the Faculty of Engi-
Soils and Foundations, 19(4): 59–68. neering, Kyushu University, 46(1): 49–62.
Karlsrud, K., and Nadim, F. 1990. Axial capacity of offshore piles Seed, H.B., Mori, K., and Chan, K.C. 1977. Influence of seismic
in clay. In Proceedings of the 22nd Offshore Technology Con- history on liquefaction of sands.. Journal of the Geotechnical
ference, Houston, Tex., 7–10 May 1990. Offshore Technology Engineering Division, ASCE, 103(4): 257–270.
Conference, Richardson, Tex. Paper 6245, pp. 404–416. Siddiqi, F.H. 1984. Strength evaluation of cohesionless soils with
Keaveny, J.M., Hansen, S.B., Madshus, C., and Dyvik, R. 1994. oversized particles. Ph.D. dissertation, University of California,
Horizontal capacity of large scale model anchors. In Proceedings Davis, Calif.
of the 13th International Conference on Soil Mechanics and Silver, M.L., Chan, C.K., Ladd, R.S., Lee, K.L., Tiedemann, D.A.,
Foundation Engineering, New Delhi, India, 5–10 January 1994. Townsend, F.C., Valera, J.E., and Wilson, J.H. 1976. Cyclic
Oxford & IBH Publishing Co. Pvt. Ltd., New Delhi, India. triaxial strength of standard test sand. Journal of the Geotechni-
Vol. 2. pp. 677–680. cal Engineering Division, 102(5): 511–523.
Kokusho, T., Hara, T., and Hiraoka, R. 2004. Undrained shear Sivathalayan, S., and Ha, D. 2004. Effect of initial stress state on the
strength of granular soils with different particle gradation. Jour- cyclic simple shear behaviour of sands. In Proceedings of the
nal of Geotechnical and Geoenvironmental Engineering, 130(6): International Conference on Cyclic Behaviour of Soils and Li-
621–629. doi:10.1061/(ASCE)1090-0241(2004)130:6(621). quefaction Phenomena, CBS04, Bochum, Germany, 31 March –
Koseki, J., and Ohta, A. 2001. Effects of different consolidation 2 April 2004. Edited by T. Triantafyllidis. CRC Press, Taylor &
conditions on liquefaction resistance and small strain quasi- Francis Group, London. pp. 207–214.
elastic deformation properties of sands containing fines. Soils Tanaka, Y., Kudo, K., Yoshida, Y., and Ikemi, M. 1987. A study
and Foundations, 41(6): 53–62. on the mechanical properties of gravel – Dynamic properties of
Koseki, J., Yoshida, T., and Takeshi, S. 2005. Liquefaction proper- reconstituted sample. Central Research Institute of the Electric
ties of Toyoura sand in cyclic torsional shear tests under low Power Industry, Akibo, Japan. Research Report U87019. [In Ja-
confining stress. Soils and Foundations, 45(5): 103–113. panese.]
Lee, K., and Seed, H.B. 1967. Dynamic strength of anisotropically Tatsuoka, F., Toki, S., Miura, S., Kato, H., Okamoto, M., Yamada,
consolidated sand. Journal of Soil Mechanics and Foundation S., Yasuda, S., and Tanizawa, F. 1986a. Some factors affecting
Division, ASCE, 93(SM5): 169–190. cyclic undrained triaxial strength of sand. Soils and Foundations,
Lee, K., and Vernese, F.J. 1978. End restraint effects on cyclic 26(3): 99–116.
triaxial strength of sand. Journal of Geotechnical Engineering, Tatsuoka, F., Ochi, K., Fujii, S., and Okamoto, M. 1986b. Cyclic
104(6): 705–719. undrained triaxial and torsional shear strength of sands for dif-

Published by NRC Research Press


Andersen 535

ferent sample preparation methods. Soils and Foundations, specimen reconstitution method on the undrained response of
26(3): 23–41. sand. Geotechnical Testing Journal, 22(3): 187–195.
Tatsuoka, F., Kato, H., Kimura, M., and Pradhan, T.B.S. 1988. Li- Wijewickreme, D., and Sanin, M. 2005. Some observations on the
quefaction strength of sands subjected to sustained pressure. cyclic loading response of a natural silt. In Proceedings of the
Soils and Foundations, 28(1): 119–131. 16th International Conference on Soil Mechanics and Geotechni-
Toki, S., Tatsuoka, F., Miura, S., Yoshimi, Y., Yasuda, S., and Ma- cal Engineering, Osaka, Japan, 12–16 September 2005. IOS
kihara, Y. 1986. Cyclic undrained triaxial strength of sand by a Press, Amsterdam, the Netherlands. pp. 627–631.
cooperative test program. Soils and Foundations, 26(3): 117–128. Wijewickreme, D., Sriskandakumar, S., and Byrne, P.M. 2004.
Triantafyllidis, Th., Wichtmann, T., and Niemunis, A. 2004. On the Characterization of cyclic loading of air-pluviated Fraser River
determination of cyclic strain history. In Proceedings of the sand for validation of numerical models. The University of Brit-
International Conference on Cyclic Behaviour of Soils and Li- ish Columbia, Vancouver, B.C. Available from www.civil.ubc.
quefaction Phenomena, CBS04, Bochum, Germany, 31 March – ca/liquefaction/ (3rd meeting - P2).
2 April 2004. Edited by T. Triantafyllidis. CRC Press, Taylor & Wong, R.T., Seed, H.B., and Chan, D.K. 1975. Cyclic loading li-
Francis Group, London. pp. 371–387. quefaction of gravelly soils. Journal of Geotechnical Engineer-
Vaid, Y.P., and Chern, J.C. 1983. Effect of static shear on resis- ing, 101(6): 571–583.
tance to liquefaction. Soils and Foundations, 23(1): 47–60. Yoshimi, Y., Tokimatsu, K., and Ohara, J. 1994. In situ liquefac-
Vaid, Y.P., and Finn, W.D.L. 1979. Static shear and liquefaction po- tion resistance of clean sands over a wide density range. Géo-
tential. Journal of Geotechnical Engineering, 105(10): 1233–1246. technique, 44(3): 479–494.
Vaid, Y.P., Sivathayalan, S., and Stedman, D. 1999. Influence of

Published by NRC Research Press

Das könnte Ihnen auch gefallen