Вы находитесь на странице: 1из 18

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/257523455

Energy approach to unsaturated cyclic strength of sand

Article  in  Bulletin of Earthquake Engineering · April 2012


DOI: 10.1007/s10518-012-9396-1

CITATIONS READS

0 48

2 authors, including:

Seyfettin Umut Umu


Anadolu University
6 PUBLICATIONS   1 CITATION   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Stifness and Damping Response of Soils Under Various Dynamic Loading Amplitudes View project

All content following this page was uploaded by Seyfettin Umut Umu on 21 January 2018.

The user has requested enhancement of the downloaded file.


Bull Earthquake Eng
DOI 10.1007/s10518-012-9396-1

ORIGINAL RESEARCH PAPER

Energy approach to unsaturated cyclic strength of sand

Volkan Okur · Seyfettin Umut Umu

Received: 30 January 2012 / Accepted: 22 October 2012


© Springer Science+Business Media Dordrecht 2012

Abstract The mechanical response to cyclic loading of saturated cohesionless soils is usu-
ally investigated by means of effective stress method considering pore water pressure changes
that lead to reduced strength and stiffness. On the other hand, the behavior of partially sat-
urated sands is different from the behavior of saturated sand deposits. The development of
negative pore water pressures in particular makes it difficult to estimate the behavior of par-
tially saturated sands. The response of partially saturated sands, however, can be examined
in a physically understandable manner by investigating their energy characteristics indepen-
dently of pore pressure behavior. To establish a general framework for understanding the
behavior of partially saturated sand, a total of 52 resonant column and dynamic torsional
shear tests were conducted under undrained conditions. The effects of factors such as the
amplitude of shear strain, relative density, saturation ratio and confining pressure on the
dynamic characteristics of the sand and on energy dissipation were studied. The use of the
energy concept in the evaluation of partially saturated soils is shown to be a promising method
for the evaluation of the cyclic behavior of partially saturated sands.

Keywords Unsaturated sand · Energy dissipation · Resonant column ·


Dynamic torsional shear

1 Introduction

Compared to unsaturated and dry conditions, the response of saturated loose, cohesionless
soil layers under seismic loadings (liquefaction phenomena) have been intensively searched
in almost every aspect of geotechnical engineering. There have been an enormous number of
laboratory and field studies focused on identifying the characteristics and mechanisms that
cause liquefaction. Over the past forty years, these advances have led to the development of
procedures for evaluating liquefaction potential in the field.

V. Okur (B) · S. U. Umu


Department of Civil Engineering, Eskisehir Osmangazi University, 26480 Eskisehir, Turkey
e-mail: vokur@ogu.edu.tr

123
Bull Earthquake Eng

The liquefaction resistance of sand is usually evaluated by means of stress- and strain-based
approaches (Green 2001; Richart et al. 1970; Yang and Sato 2000). Liquefaction evaluation
can also be performed using the amount of energy dissipated in soil during the propagation
of seismic waves through soil deposits, rather than the induced cyclic shear stress or strain.
The assumption is that part of the energy is dissipated throughout the soil during deformation
under cyclic loads and part of the energy is dissipated into the soil. The dissipated energy
per unit volume is represented by the area of the hysteretic stress-strain loop and can be
determined experimentally. There are two main reasons for the development of the energy
approach in liquefaction evaluations. First, the energy released during earthquakes has long
been quantified, and simple relationships have been established with general seismological
parameters. Second, laboratory work by Nemat-Nasser and Shokooh (1979) has shown that
there is a relationship between dissipated energy and generated pore pressures. Further stud-
ies by many researchers has shown that the energy approach could be more practical than
cyclic stress or strain in liquefaction evaluation because both cyclic stress and strain histories
are taken into account (Berrill and Davis 1985; Davis and Berrill 2001, 1982; Figueroa 1990;
Law et al. 1990; Liang et al. 1995; Ostadan et al. 1998; Trifunac 1995).
It is simple to determine the response of saturated soils, especially under undrained con-
ditions, if pore pressures can be measured correctly. However, because of the fluctuation
of the water table due to either natural phenomena (precipitation, evaporation) or manmade
processes (irrigation, pumping), partial saturation conditions may develop. Yang et al. (2003)
demonstrated that partial saturation could be encountered below the phreatic surface due to
the fluctuation or flooding of groundwater. Yegian et al. (2007) showed that entrapped air
bubbles would remain for a long time in a liquefaction-susceptible sand layer under appropri-
ate hydrostatic conditions. Even in gravelly soils, the soil stratum could contain entrapped air
up to 10 m below the phreatic surface (Ishihara et al. 2004; Kokusho 2000; Nagao et al. 2007;
Tsukamoto et al. 2007). All of these studies showed that soils could be partially saturated
even below the phreatic surface and that entrapped air may be held in the soil for a long time.
In loose-medium saturated sandy soils, an increase in the pore water pressure decreases
the effective stress, which causes degradation of the stiffness and shear strength. However,
because of matric suction, it appears that some strength gain occurs in partially saturated
soils due to an increase in effective stress (Lu and Likos 2004). This affects soil dynamic
parameters such as shear wave velocity, initial shear modulus, damping ratio, and unit weight.
Laboratory tests have shown that the liquefaction resistance of sandy soils depends strongly
on the degree of saturation. For a particular value of the cyclic stress ratio, the number of
cycles causing liquefaction was observed to increase significantly with decreasing values of
the pore pressure parameter B. Yoshimi et al. (1989), Tsukamoto et al. (2002) have shown
that a minor decrease in the degree of saturation to 90 % increases the liquefaction resistance
to twice that of fully saturated samples. Martin et al. (1978), Xia and Hu (1991) showed that
a small amount of entrapped air can considerably increase the liquefaction resistance of sand
under a given state of packing. Their test results showed that liquefaction resistance increased
more than 30 % when the degree of saturation decreased by only 2 %. The outcome of these
studies confirmed that a small reduction in the degree of saturation from full saturation can
result in a significant increase in the shear resistance of sand against liquefaction (Bouferra
et al. 2007; Delia 2010; Ishihara et al. 2001; Ishihara and Tsukamoto 2004; Jafari-Mehrabadi
et al. 2007; Okamura et al. 2006; Sherif et al. 1977; Yang 2002; Yang et al. 2004). Partially
saturated soils are also considered to be safe against cyclic shear failure because of the high
compressibility of water (Okamura and Soga 2006).
Partial saturation changes the shear wave velocity of soils, which also affects the propaga-
tion of seismic waves and resulting accelerations through soil deposits. The effect of partial

123
Bull Earthquake Eng

saturation to the site response and seismic shaking of near-surface soils have been studied by
several researchers (Kokusho and Matsumoto 1999; Mucciarelli et al. 2003; Mylonakis and
Gazetas 2002; Yang 2006; Wang and Hao 2002). The findings showed that a slight decrease
in the saturation ratio of soil may have a considerable influence on surface amplitudes, espe-
cially in vertical directions, and also produce larger amplitude ratios between the vertical and
horizontal components at the surface. The studies confirm the potential importance of the
saturation condition of near-surface soils in site response analysis (Ghayoomi and McCartney
2012; Yang and Sato 2002; Yang et al. 2002). Despite the increase in site amplification due
to the suction effect and increase in effective stress, failure rarely occurs in partial saturated
soil deposits.
The 1995 Kobe earthquake in Japan and the 2001 Bhuj earthquake in India are two sig-
nificant field examples of the effect of saturation on liquefaction potential. Lin et al. (2005)
reported over 300 examples of entrapped air in foundation soil preventing serious damage by
liquefaction in the 1995 Kobe earthquake. Singh et al. (2005) reported on the damage of earth
dams in the 2001 Bhuj earthquake. Because the reservoirs of the dams were almost empty at
the time of the earthquake, the soil stratum was only partially saturated. These observations
are in good agreement with the laboratory studies that show that partial saturation increases
the strength of soil against cyclic failure.
The purpose of this paper is to examine the physical implications of simplified energy
approaches in the use of the energy concept for evaluating the response of partially saturated
clean sands to cyclic loading.

2 Overview and basic considerations

Studies on soil liquefaction have revealed that the damage potential of the soil under cyclic
loading can be described as a function of the energy dissipated into the soil during loading.
Nemat-Nasser and Shokooh (1979) found that there is a relationship between the energy
required to rearrange sand grains and the dissipated energy calculated from hysteresis loops.
The relationship was found to be correlated with shear stress, number of cycles, initial void
ratio and pore pressure. Further studies were performed on sandy soils by Figueroa et al.
(1994), Kim and Park (2008), Law et al. (1990), in order to establish a direct relationship
between pore pressure and dissipated energy per unit volume under different loading condi-
tions.
The structural collapse of the soil skeleton due to cyclic shear stresses, with a parallel loss
of energy due to frictional mechanisms, “the cumulative energy loss up to liquefaction” (WL )
has been identified as an index for the assessment of liquefaction potential (Berrill and Davis
1985; Figueroa et al. 1994; Trifunac 1995). Because the cumulative energy method is a prom-
ising technique for the assessment of the liquefaction resistance of saturated sands, further
development of the technique would extend this type of approach to the case of unsaturated
clean sands. Developing an energy approach requires the relationship between the amount of
dissipated energy and other variables such as the degree of saturation and the shear modulus.
Studies have shown that there is a unique relationship between the dissipation energy and the
dynamic soil response in terms of dynamic strength, shear modulus and permanent strain.
The relationship between the intensity of the dissipated energy per unit volume required for
cyclic failure and other important parameters and their effects on the energy dissipation must
also be assessed.
For a perfectly elastic material, if the external forces remain unchanged, then the level of
input energy remains constant. On the other hand, when external forces are applied to a fluid,

123
Bull Earthquake Eng

the energy is directly spent through viscous deformation, and when the forces are removed,
no strain energy is left to be returned. The mechanical characteristics of soils may deviate
from those of an ideal visco-elastic material obeying Hooke’s law. At either small or large
deformation levels, a material may exhibit both recoverable and irreversible strains due to
its visco-elastic nature.
The most common method of characterizing energy dissipation is by measuring the actual
energy dissipated into a sample through a deformation cycle. The area bounded by the hyster-
esis loop expresses the energy loss in the system particles in a two-dimensional stress space,
whereas in terms of stress and strain, the bounded area quantifies the accumulated energy
dissipated per unit volume (W). For torsional shear, W can be evaluated by the following
expression (Figueroa et al. 1994):


n−1
W = 0.5(τi+1 + τi )(γi+1 − γi ) (1)
i=1

where W is energy loss per unit volume, λ is the axial strain, τ is the shear stress and n is
the number of points recorded up to a specified instant in the loading history. This critical
energy can be determined in any conventional cyclic laboratory test if the rate of data acqui-
sition is sufficiently rapid and the hysteretic loops are defined by closely spaced points. The
main advantages of the energy approach over other methods are the following: (a) energy is
directly related to the intensity of dynamic loading; (b) it is not necessary to decompose the
shear stress history and the equivalent number of uniform cycles for selected stress or strain
levels; (c) it is practically independent of the load waveform and, because there is no need
to model a complicated random stress history in laboratory experiments, a simple sinusoidal
pattern of loading can be used in all cases; and (d) it is largely independent of the type of
testing device, which facilitates the comparison of results obtained in different laboratories.
Dissipated energy can be divided into two components. One component is the rearrange-
ment of particles and distortion of their bonding, which triggers plastic strains. The second
component is the accumulation of energy in the form of heat, kinetic energy or surface energy
of the particles, which increases the mobility of soil particles. Accumulation of this energy
with consecutive cycles is the primary reason of dynamic instabilities in soils.

3 Tested material, equipment and test procedures

Standard Toyoura sand, with a mean grain diameter D50 = 0.21 mm, density of solid par-
ticles ρs = 2, 656 Mg/m3 , maximum void ratio emax = 0.992, and minimum void ratio
emin = 0.632, was used to prepare test samples with relative densities varying from 40
to 70 %. To examine the effect of saturation ratio from a broad perspective, it was highly
desirable to prepare all of the test samples according to the same procedure. Samples were
prepared by the moist tamping method, in which the sand is mixed with water to different
moisture contents, spread in molds and compacted in layers with a tamping rod to the desired
density. The mold is 5 cm in diameter and 10 cm in height. The samples were isotropically
consolidated to different confining pressures between 30 and 150 kPa. The properties of the
tested samples and test parameters are summarized in Table 1. “ω” is the initial water content
and Nf is the total number of cycles at the time of failure.
An AC servo motor was used in the loading system for the dynamic torsional shear appara-
tus used in this study, so that dynamic torsional loading under stress control could be applied
accurately to samples in the horizontal direction. The system is also capable of conducting

123
Bull Earthquake Eng

Table 1 Properties of tested samples and test parameters

Sample no. eo σc (kPA) Sr (%) ω (%) Dr (%) Nf

S1ST1 0.775 100 100 29.3 57 131


S1ST2 0.768 50 100 29.1 59 36
S1PS1 0.771 100 72 21.0 58 417
S1PS2 0.782 100 65 19.3 55 406
S1PS3 0.768 100 85 24.7 59 209
S1PS4 0.775 100 44 12.9 57 507
S1PS5 0.775 100 45 13.2 57 512
S2ST1 0.721 30 100 27.3 72 47
S2ST2 0.728 50 100 27.6 70 55
S2PS1 0.735 50 83 23.1 68 253
S2PS2 0.735 50 55 15.3 68 470
2PS3 0.728 100 65 12.7 52 497
S2PS4 0.735 50 56 15.6 68 466
S2PS5 0.735 50 46 22.8 68 265
S3ST1 0.818 50 100 31.0 45 38
S3ST2 0.814 80 100 31.0 46 72
S3ST3 0.822 80 90 28.0 44 175
S3PS1 0.825 80 90 25.6 43 274
S3PS2 0.815 80 82 25.4 45 265
S3PS3 0.814 80 74 22.9 45 356
S3PS4 0.818 80 48 14.9 45 501
S3PS5 0.814 80 77 23.8 46 320
S3PS6 0.814 80 65 20.1 46 421
S3PS7 0.807 80 77 23.5 48 324
S3PS8 0.811 80 55 16.9 47 472
S3PS9 0.814 80 77 23.8 46 325
S3PS10 0.829 80 40 12.6 42 512
S4ST1 0.768 50 100 29.1 59 31
S4ST2 0.771 80 100 29.2 58 47
S4ST3 0.768 80 100 29.1 59 41
S4PS1 0.764 80 92 26.6 60 141
S4PS2 0.768 100 63 26.5 40 153
S4PS3 0.760 80 52 15.0 61 485
S4PS4 0.764 80 47 13.6 60 491
S4PS5 0.771 80 47 13.7 58 487
S4PS6 0.775 80 71 20.8 57 380
S5ST1 0.796 50 100 30.2 51 30
S5ST2 0.793 100 100 30.0 52 59
S5ST3 0.796 100 100 30.2 51 54
S5PS1 0.793 100 75 22.5 52 336
S5PS2 0.804 100 45 13.7 49 512
S5PS3 0.793 100 65 22.2 61 365

123
Bull Earthquake Eng

Table 1 continued
Sample no eo σc (kPA) Sr (%) ω (%) Dr (%) Nf

S5PS4 0.800 100 70 21.2 50 398


S5PS5 0.789 100 81 24.2 53 285
S6ST1 0.757 100 100 28.7 62 56
S6ST2 0.753 100 100 28.5 63 77
S6ST3 0.757 100 100 28.7 62 85
S6PS1 0.750 100 77 21.9 64 357
S6PS2 0.742 100 42 11.8 66 612
S6PS3 0.750 100 64 18.2 64 533
S6PS4 0.742 100 55 15.5 66 580
S6PS5 0.746 100 63 22.0 71 417
eo initial void ratio, σc ’ effective pressure
Sr saturation ratio, Dr relative density
ω initial water content, Nf number of cycles

resonant column tests, so the initial shear modulus values, Gmax were calculated from the
resonance frequency.
The type of cyclic loading used in the laboratory tests is a “multi-stage” type of load-
ing, in which the samples were consolidated and subjected to ten cycles of cyclic torsional
stresses, starting from quite low stress amplitudes under undrained conditions. The samples
were cyclically loaded step by step with gradually increasing stress amplitudes covering a
wide strain range of 10−3 –101 % under a constant confining pressure.

4 Test results

Typical hysteresis loops generated during the tests for different saturation conditions are
shown in Fig. 1. A fresh sample was prepared by way of an identical procedure for each
test. The samples with Dr = 53 − 55 % were consolidated to σc = 100 kPa and subjected
to the same prescribed amplitude of torsional stress cycles, starting from low stress ampli-
tudes under undrained conditions. Figure 1a shows the results for a sample that is fully
saturated, while Fig. 1b, c show the results for samples at degrees of saturation of 85 and
57 %, respectively.
For the saturated sample, the hysteresis loops tend to become gradually leveled as the
sample begins to fail, showing rapid degradation. The build-up of pore water pressure, which
increases rapidly, causes the saturated sample to fail quickly. However, for the samples with
Sr = 85 % and Sr = 57 %, the material response is different and can be explained using the
model proposed by Fisher (1926). According to the model, a water–air meniscus affects the
“stress state” of two solid spheres. The meniscus water at the contact points of two spheres
induces a force normal to the plane that passes through the contact point and orthogonal to
the line that connects the center of the spheres. Mancuso et al. (2002) stated that this force is a
result of menisci water, and amplifies as suction increases. As a result, a greater normal force
holds the spheres together and their contact performs frictionally greater, limiting slippage
strength. It is clear in Fig. 1a that stress-strain response tends to change appreciably at the
time of cyclic failure with shear strain and with the progression of cycles as well. However,

123
Bull Earthquake Eng

Fig. 1 Typical stress strain 150

Cyclic shear stress amplitude, kPa


Sr=100%
Dr=55%
behavior for identical samples σ'c =100 kPa (a)
having different saturation ratios 100 Nf=214

50

-50

-100

-150
-1 0 1 2 3 4 5 6
Cyclic shear strain, γ %
150
Cyclic shear stress amplitude, kPa
Sr=85%
Dr=53%
σ'c =100 kPa
(b)
100 Nf=601

50

-50

-100

-150
-1 0 1 2 3 4 5 6
Cyclic shear strain, γ %
150
Cyclic shear stress amplitude, kPa

Sr=55%
Dr=54%
σ'c =100 kPa
Nf=631
(c)
100

50

-50

-100

-150
-1 0 1 2 3 4 5 6
Cyclic shear strain, γ %
50
(d)
40
Cyclic stress ratio, CSR

30

20
Dr=54%, Sr=55%

10 Dr=53%, Sr=85%

Dr=55%, Sr=100%

0
0 1 2 3 4 5 6
Cyclic shear strain, γ %

123
Bull Earthquake Eng

S1ST2, Dr=59%, Sr=100%


100
S1ST1, Dr=57%, Sr=100%

S1PS3, Dr=59%, Sr=85%


Dissipated unit energy per cycle kJ/m 3

80 S1PS2, Dr=55%, Sr=65%

S1PS1, Dr=58%, Sr=72%

S1PS4, Dr=57%, Sr=44%


60

40

20

0
0 100 200 300 400 500 600
Number of cycles, N

Fig. 2 Buildup of dissipated energy with number of cycles for different saturation ratios

the characteristic hysteresis loops shown in Fig. 1b, c are different from that observed for the
fully saturated sample. First, both of the partially saturated samples take significantly longer
to fail due to the suction effect (Fig. 1b, c), and the progression of cycles causes very small
changes in the stress-strain response. For both of the partially saturated samples, the hyster-
esis loops after a large number of cycles remain wide and nearly steady during considerable
shear stress development. The behavior observed can be explained by the work of Mancuso
et al. (2002), which showed that partially saturated aggregates of spheres show a stiffer and
more resistant load response than that of submerged spheres. However, this effect does not
increase indefinitely, because the normal force induced by the meniscus water approaches a
limiting value due to the progressive reduction in the meniscus radius as suction increases.
This behavior for the same samples is more evident in Fig. 1d when given in terms of cyclic
stress ratio and cyclic shear strain. The number of cycles, N was chosen as 5 of every 10th
cycle and analyzed to obtain a representative cyclic stress-strain behavior.
Figure 2 shows typical results for a test group in which the value of the dissipated unit
energy at the end of each cycle of loading is plotted against the number of cycles. The
amount of dissipated energy per unit volume per cycle into the sample is equal to the area of
the hysteresis loop and is calculated from Eq. (1).
Each of the test groups had two types of samples, fully saturated and partially satu-
rated. The results shown in Fig. 2 are from samples with relative densities of approximately
Dr = 60 %. Sample S1ST2 had an effective confining stress of 50 kPa, whereas other sam-
ples had an effective confining stress of 100 kPa. Sample S1ST2 failed at cycle 36, whereas
sample S1ST1 failed at cycle 131, which means that the energy required for cyclic failure
is substantially smaller for low confining pressures at the same relative density. It is very
well known that confining pressure is a parameter affecting cyclic failure of both cohe-
sive and cohesionless soils. It can also be seen that the dissipated unit energy per cycle at
the time of cyclic failure for S1ST1 is almost twice as great for sample S1ST2 (17 kJ/m3
for S1ST2 and 40 kJ/m3 for S1ST1). A sudden increase in unit dissipated energy at the
time of failure is exhibited for both saturated samples (S1ST1 and S1ST2) and partially

123
Bull Earthquake Eng

S2ST1, Dr=72%, Sr=100%


100
S2ST2, Dr=70%, Sr=100%
Dissipated unit energy per cycle kJ/m 3
S2PS1, Dr=68%, Sr=83%
80
S2PS2, Dr=68%, Sr=55%

S2PS5, Dr=68%, Sr=46%


60

40

20

0
0 100 200 300 400 500 600
Number of cycles, N

Fig. 3 Variation of energy response with respect to number of cycles

saturated samples (S1PS1 and S1PS3, which had saturation ratios of 72 and 85 %, respec-
tively). In addition, the dissipated unit energy per cycle was almost same at the time of failure
for samples S1ST1 and S1PS3, even though they had different saturation ratios. It is also
very interesting to note that there is a significant increase in unit energy during the last cycles
for fully saturated samples and even for sample S1PS3, which had a saturation ratio of 85 %.
However, this effect was not observed for other partially saturated samples with saturation
ratios between 44 and 72 %. Cyclic resistance increases with a decrease in saturation ratio,
and this trend accelerates when saturation ratios are lower than 50 %. Samples S1PS1, S1PS2
and S1PS4 exhibited greater cyclic strength than samples S1ST1 and S1PS3. Similar results
for another test group can be seen in Fig. 3. From the results for this and other test groups,
taken together with the partial build-up of energy, it can be concluded that complete softening
cannot occur in samples with saturation ratios below 60 %. For all of the samples, during the
initial cycles, despite their different saturation ratios, dissipated unit energy values were very
close.
If the degree of saturation is sufficiently low (e.g., Sr < 50 %), the sand tends to exhibit
strain-hardening characteristics with an increasing number of cycles. It can be seen that, in
the case of a partially saturated sample, the dissipation of unit energy per cycle occurs very
slowly (which is why cyclic failure takes so long to occur), to the degree that the energy
values are almost same as those of the saturated samples at the same confining stress (S1ST2
and S1PS4 in Fig. 2, S2ST2 and S2PS3 in Fig. 3). One of the key points illustrated by Figs. 2
and 3 is that when the sample is approaching a failure state, the dissipated energy increases
with every cycle. However, for partially saturated conditions (except for values of Sr > 80),
there is almost no difference between the cycles in terms of dissipated energy.
It should also be noted that for the same density of sand, the dissipated energy curve
has a unique shape unless the sand is saturated. It is apparent from Figs. 2 and 3 that as the
cyclic loading progresses, dissipated unit energy paths move along a smooth upward concave
line for partially saturated samples, which suggests that the matric suction increases cyclic
strength. For saturated samples and samples with saturation ratios Sr >80 %, however, there

123
Bull Earthquake Eng

100
(a)
Dissipated unit energy per cycle kJ/m 3
80

60
S3PS9, Dr=46%, Sr=77%

S3PS3, Dr=45%, Sr=74%


40

S3PS5, Dr=46%, Sr=77%

20 S3PS2, Dr=45%, Sr=82%

S3PS1, Dr=43%, Sr=90%

0
0 1 2 3 4
Cyclic shear strain, γ %

100 S3PS10, Dr=42%, Sr=40%

(b)
Dissipated unit energy per cycle kJ/m 3

S3PS4, Dr=45%, Sr=48%

80 S3PS8, Dr=47%, Sr=55%

S3PS6, Dr=46%, Sr=65%

60

40

20

0
0 1 2 3
Cyclic shear strain, γ %

Fig. 4 Typical relationship between dissipated unit energy per cycle and cyclic shear strain

is a sudden build-up of dissipated energy, indicating sudden softening of the sand, as seen in
Figs. 2 and 3.
Changes in unit energy with changes in shear strain for relatively loose samples are illus-
trated in Fig. 4. For saturation ratios in excess of 75 %, dissipated unit energy tends to increase
linearly (Fig. 4a), The response of samples are similar to saturated ones. However for satu-
ration ratios lower than 75 %, dissipated unit energy shows nonlinear trend (Fig. 4b). This
trend is more evident in terms of accumulated dissipated energy as shown in Fig. 5.
Figure 5 illustrates the relationship between the accumulated energy dissipated per unit
volume and cyclic shear strain at different relative densities for Sr ≈ 65 % and a confining
stress of σc = 100 kPa. For relative densities of 40 and 52 %, the strain paths are close to
each other. However, there is a significant difference for relative densities of 61 and 71 %. It
can be concluded that the denser the sample is, the higher the amount of accumulated energy
dissipated per unit volume for the same saturation ratio.

123
Bull Earthquake Eng

5000

Ac c u m u l a t e d e n e rg y d i s s i p a t e d p e r u n i t v o l u m e ,
4000

3000
kj/m3

2000

1000

0
0 1 2 3 4 5
Cyclic shear strain (%)

Fig. 5 Buildup of dissipated energy per unit volume under a constant cell pressure of 100 kPa

1.00

0.75
G/Gmax

0.50
S4ST2, Dr=58%, Sr=100%

S4PS1, Dr=60%, Sr=92%

S4PS3, Dr=61%, Sr=52%


0.25
S4PS5, Dr=58%, Sr=47%

S4PS6, Dr=57%, Sr=71%

0.00
1 10 100
Dissipated Unit Energy Per Cycle kJ/m 3

Fig. 6 Effects of saturation ratio on shear modulus ratio and dissipated unit energy per cycle

4.1 Shear modulus and dissipated energy

The shear modulus is an important parameter for characterizing the dynamic response of
soils. To investigate the relationship between the shear modulus and the accumulated energy
dissipated per unit volume, additional tests were conducted. The results of another set of tests
on somewhat denser samples at relative densities of approximately 50 and 60 % are presented
in Figs. 6 and 7. Initial shear modulus values were calculated from the results of resonant
column tests, and shear modulus variation with respect to shear strain is calculated from the
results of torsional shear tests. The same trend is detected in the overall response with respect
to the influence of the saturation ratio. The results of the tests in terms of unit dissipated
energy are similar, as shown in Figs. 3 and 4. Figure 6 compares the relationship between

123
Bull Earthquake Eng

1.00

0.75
G/Gmax

0.50

S5ST2, Dr=52%, Sr=100%

S5PS1, Dr=52%, Sr=75%

0.25 S5PS2, Dr=49%, Sr=45%

S5PS4, Dr=50%, Sr=70%

S5PS5, Dr=53%, Sr=81%

0.00
10 100 1000 10000
Dissipated Unit Energy kJ/m 3

Fig. 7 Effects of saturation ratio on shear modulus ratio and dissipated unit energy per cycle

1.00

0.75
G/Gmax

0.50
S4ST2, Dr=58%, Sr=100%

S4PS1, Dr=60%, Sr=92%

S4PS3, Dr=61%, Sr=52%


0.25
S4PS5, Dr=58%, Sr=47%

S4PS6, Dr=57%, Sr=71%

0.00
10 100 1000 10000
Dissipated Unit Energy kJ/m 3

Fig. 8 Effects of saturation ratio on shear modulus ratio and dissipated unit energy

the shear modulus ratio and the dissipated unit energy per cycle per unit volume for different
saturation ratios. It can be seen that the degradation paths for partially saturated samples
are almost coincident with those for saturated samples. However, samples with saturation
ratios lower that 50 % have different paths. It may be seen that, although there is some scatter
among the individual sets of data, the decrease in the shear modulus ratio becomes more
pronounced as Sr increases. These two graphs show how the shear modulus ratio curve rises
as the saturation ratio decreases.
The results for the same test group are presented in terms of shear modulus ratio versus
dissipated unit energy (cumulative energy) in Figs. 8 and 9. It can be seen that at a value
of G/Gmax = 1, the dissipated unit energy value increases 15 times from Sr = 100 % to
Sr = 47 %. On the other hand, at G/Gmax = 0.5, the dissipated unit energy increases by

123
Bull Earthquake Eng

1.00

0.75
G/Gmax

0.50 S5ST2, Dr=52%, Sr=100%

S5PS1, Dr=52%, Sr=75%

S5PS2, Dr=49%, Sr=45%


0.25
S5PS4, Dr=50%, Sr=70%

S5PS5, Dr=53%, Sr=81%

0.00
1 10 100
Dissipated Unit Energy Per Cycle kJ/m 3

Fig. 9 Effects of saturation ratio on shear modulus ratio and dissipated unit energy

approximately one order of magnitude from fully saturated to Sr = 47 %. Figure 9 shows


that at a value of G/Gmax = 1, the dissipated unit energy increases 10 times from Sr = 100 %
to Sr = 45 %. On the other hand, at G/Gmax of 0.5, the dissipated unit energy increases by
approximately 17 times from fully saturated to Sr = 47 %.
It should also be noted that the samples tended to show an almost identical response
despite large differences in the degree of saturation. The shear modulus ratio decreases with
the progression of cycles. However, the rate of this decrease is much slower in samples S2PS3
and S4PS5, as shown in Fig. 8, and sample S2PS2, as shown in Fig. 9, which had saturation
ratios below 50 %.

4.2 The amount of energy required for cyclic failure

Based on the experimental results, it is essential to find out how much energy will be needed
for the material to fail completely. Although it is very empirical, one conservative method
that can be used is to evaluate the energy values corresponding to a predefined strain of cyclic
failure. For the entire group of tests, failure was defined as corresponding to a 2 % single
strain amplitude. The level of energy per unit volume accumulated at 2 % cyclic strain was
calculated for each test. The results are summarized in Fig. 10.
The accumulated energy dissipated per unit volume at the time of cyclic failure tends to
increase with increasing relative density, for degrees of saturation between 42 and 74 %. This
relation does not exist for Sr > 74 %. The less saturated and the denser the sample, the higher
the energy required to fail the sample.
The dissipated unit energy at the last cycle at the time of failure is shown in Fig. 11.
The same ranges of saturation ratio values are plotted against relative density for the same
effective confining pressures. The same tendency is observed for partially and fully saturated
samples, as shown in Fig. 10. As the saturation ratio decreases significantly, more energy is
needed to cause complete failure, due to the suction effect.
The amount of energy per unit volume for a strain level of 2 % is also calculated for dif-
ferent confining pressures (Fig. 12). There is a noticeable increase in the energy dissipation,

123
Bull Earthquake Eng

8000 Sr=100%

Sr=75-85%

Sr=64-74%
Accumulated energy, ΔW, kj/m 3

6000 Sr=42-55%

4000

2000

σ c= 100 kPa

0
40 50 60 70 80
Relatif Density, Dr (%)

Fig. 10 Variation of accumulated energy with relative density for different saturation ratio ranges

Sr=100%
100
Dissipated unit energy per cycle, kj/cycle/m 3

Sr=75-85%

80 Sr=64-74%

Sr=42-55%
60

40

20
σ 'c=100 kPa

0
40 50 60 70 80
Relative Density, Dr (%)

Fig. 11 Effect of saturation ratio on dissipated unit energy with relative density

especially for lower saturation ratios. It is also apparent that for the same effective confining
pressure, the dissipated energy per unit volume increases with relative density. One of the
highlights of Fig. 12 is the fact that the accumulated energy W increases in proportion
to relative density. However, when the saturation ratio decreases, the quantity W tends
to increase with increasing relative density, which is consistent with the real behavior of
sandy soils. Another characteristic feature observed in Fig. 12 is that the effect of confining
pressure is more significant for samples with saturation ratios Sr = 100 − 75 %. The effect
of confining pressure tends to decrease as the saturation ratio decreases and relative density
increases.

123
Bull Earthquake Eng

1000
Accumulated energy, ΔW, kj/m 3

Accumulated energy, ΔW, kj/m 3


2500
50 kPa 50 kPa

800 80 kPa 80 kPa


2000
100 kPa 100 kPa

600 1500

400 1000

200
500
S=100%
S=75-90%
0
40 50 60 70 80 0
40 50 60 70 80
Relatif Density, Dr (%) Relatif Density, Dr (%)
10000
Accumulated energy, ΔW, kj/m 3

Accumulated energy, ΔW, kj/m 3


12000
80 kPa 80 kPa
100 kPa
10000 100 kPa
8000

8000
6000
6000
4000
4000

2000
2000
S=60-75% S=42-60%
0 0
40 50 60 70 80 40 50 60 70 80
Relatif Density, Dr (%) Relatif Density, Dr (%)

Fig. 12 Effect of confining pressure on accumulated energy considering relative density and saturation ratio

5 Conclusions

A series of resonant column and cyclic torsional tests were conducted on sand samples. Each
test series consisted of 5–10 samples with similar relative densities varying from 40–70 %
and saturation ratios of Sr = 40 − 100 %. After determining the resonant frequencies from
resonant column tests, the samples were subjected to cyclic torsional stress with the ampli-
tude increased every five cycles. Based on the results of these tests, the following conclusions
can be drawn.
The cyclic strength of sand depends on its relative density, its saturation ratio, and its
confining pressure.
Partially saturated soils with saturation ratios Sr < 60 % can be considered to be safe
against cyclic shear because of matric suction.
The level of confining stress has a profound influence on the dynamic behavior of partially
saturated sand.
Cyclic failure of the soil structure occurs when the amount of accumulated internal energy
exceeds a specific energy level.
For partially saturated samples, the unit energy required for the same deformed state is up
to 3–4 times greater than for saturated samples.
The accumulated energy dissipated per unit volume and the ratio of accumulated energy to
relative density and confining pressure increases with increasing relative density, particularly
for Sr < 60.

Acknowledgments The writers would like to acknowledge the Eskisehir Osmangazi University, Scientific
Research Project Unit for the support of this work through project No: 200915006. The writers would also
like to thank the reviewers for their time and their thoughtful comments and suggestions.

123
Bull Earthquake Eng

References

Berrill JB, Davis RO (1985) Energy dissipation and seismic liquefaction of sands: revised Model. Soils Found
25(2):106–118
Bouferra R, Benseddiq N, Shahrour I (2007) Saturation and preloading effects on the cyclic behavior of sand.
Int J Geomech ASCE 7(5):396–401
Davis RO, Berrill JB (2001) Pore pressure and dissipated energy in earthquakes-field verification. J Geotech
Geoenviron Eng ASCE 127(3):269–274
Davis RO, Berrill JB (1982) Energy dissipation and seismic liquefaction in sands. Earthq Eng Struct Dyn
10:56–68
Delia N (2010) Laboratory testing of the monotonic behavior of partially saturated sandy soil. Earth Sci Res
J 14(2):181–186
Figueroa JL (1990) A method for evaluating soil liquefaction by energy principles. In: Fourth US national
conference on earthquake engineering, Palm Springs, California, 20–24; pp 695–704
Figueroa JL, Saada AS, Liang L, Dahisaria NM (1994) Evaluation of soil liquefaction by energy principles.
J Geotech Eng ASCE 120(9):1554–1569
Fisher RA (1926) On the capillary forces in an ideal soil. J Agric Sci 16:492–505
Ghayoomi M, McCartney JS (2012) Centrifuge evaluation of the impact of partial saturation on the ampli-
fication of peak ground acceleration in soil layers. In: GeoCongress: state of the art and practice in
geotechnical engineering roles and influences of physical modeling on state of the art and practice of
geotechnical earthquake engineering, pp 1968–1977
Green RA (2001) Energy based evaluation and remediation of liquefiable soils. Ph.D. thesis, Department of
Civil Engineering, Virginia Polythecnic Institute and State University, Blacksburg
Ishihara K, Tsuchiya H, Huang Y, Kamada K (2001) Recent studies on liquefaction resistance of sand effect of
saturation. In: Proceedings of 4th conference recent advances in geotechnical earthquake. Engineering,
2001 (Keynote Lecture)
Ishihara K, Tsukamoto Y, Kamada K (2004) Undrained behaviour of near-saturated sand in cyclic and mono-
tonic loading, In: Conference cyclic behavior of soils and liquefaction phenomena, Bochum, pp 27–39
Ishihara K, Tsukamoto Y (2004) Cyclic strength of imperfectly saturated sands and analysis of liquefaction.
In: Proceedings of Japan Academy, 80, pp 372–391
Jafari-Mehrabadi A, Abdi MA, Popescu R (2007) Analysis of liquefaction susceptibility of nearly saturated
sands. Int J Numer Anal Methods Geomech 31(5):691–714
Kim S, Park K (2008) Proposal of liquefaction potential assessment procedure using real earthquake loading.
J Civil Eng ASCE 12(1):15–24
Kokusho T (2000) Correlation of pore-pressure B-value with P-wave velocity and Poisson’s ratio for imper-
fectly saturated sand or gravel. Soils Found 40(4):95–102
Kokusho T, Matsumoto M (1999) Nonlinear site amplification in vertical array records during Hyogo-ken
Nanbu earthquake. Soils Found 2:1–9
Law KT, Cao YL, He GN (1990) An energy approach for assessing seismic liquefaction potential. Can Geotech
J 27:320–329
Liang L, Figueroa JL, Saada AS (1995) Liquefaction under random loading: unit energy approach. J Geotech
Eng ASCE 121(11):776–781
Lin CH, Lee VW, Trifunac MD (2005) The reflection of plane waves in a poroelastic half-space saturated
with inviscid fluid. Soil Dyn Earthq Eng 25:205–223
Lu N, Likos W (2004) Unsaturated soil mechanics. Wiley, NJ 556
Mancuso C, Vassallo R, D’Onofrio A (2002) Small strain behaviour of a silty sand in controlled suction
resonant column–torsional shear tests. Can Geotech J 39(1):22–31
Martin GR, Finn WDL, Seed HB (1978) Effects of system compliance on liquefaction test. J Geotech Eng
Div ASCE 104(GT4):463–479
Mucciarelli M, Gallipoli MR, Arcieri M (2003) The stability of the horizontal-to-vertical spectral ratio of
triggered noise and earthquake recordings. Bull Seismol Soc Am 93(3):1407–1412
Mylonakis G, Gazetas G (2002) Kinematic pile response to vertical P-wave seismic excitation. J Geotech
Geoenviron Eng 128(10):860–867
Nagao K, Azegami Y, Yamada S, Suemasa N, Katada T (2007) A micro-bubble injection method for a counter-
measure against liquefaction. In: 4th International conference on earthquake geotechnical engineering,
Thessaloniki, pp 25–28
Nemat-Nasser S, Shokooh A (1979) A unified approach to densification and liquefaction of cohesionless sand
in cyclic shearing. Can Geotech J 16:659–678
Okamura M, Ishihara M, Tamura K (2006) Degree of saturation and liquefaction resistances of sand improved
with sand compaction pile. J Geotech Geoenviron Eng ASCE 132(2):258–264

123
Bull Earthquake Eng

Okamura M, Soga Y (2006) Effects of pore fluid compressibility on liquefaction resistance of partially satu-
rated sand. Soils Found 46(5):695–700
Ostadan F, Deng N, Arango I (1998) Energy-based method for liquefaction potential evaluation. In: Eleventh
European conference on earthquake engineering, Paris, pp 6–11
Richart FE, Hall JR, Woods RD (1970) Vibrations of soils and foundations. Prentice-Hall, Inc, Englewood
Cliffs 414
Sherif MA, Tsuchiya C, Ishibashi I (1977) Saturation ef-fect on initial soil liquefaction. J Geotech Eng Div
ASCE 103(8):914–917
Singh R, Roy D, Jain SK (2005) Analysis of earth dams affected by the 2001 Bhuj earthquake. Eng Geol
80:282–291
Trifunac MD (1995) Empirical criteria for liquefaction in sands via standard penetration tests and seismic
wave energy. Soil Dyn Earthq Eng 14:419–426
Tsukamoto Y, Kamata T, Tatsuoka F, Ishihara K (2007) Undrained flow characteristics of partially saturated
sands soils in triaxial tests. In: 4th International conference on earthquake geotechnical engineering,
Thessaloniki, pp 25–28
Tsukamoto Y, Ishihara K, Nakazawa H, Kamada K, Huang Y (2002) Resistance of partly saturated sand to
liquefaction with reference to longitudinal and shear wave velocities. Soils Found 42(6):93–104
Wang S, Hao H (2002) Effects of random variations of soil properties on site amplification of seismic ground
motions. Soil Dyn Earthq Eng 22:551–564
Xia H, Hu T (1991) Effects of saturation and back pressure on sand liquefaction. J Geotech Eng ASCE
117(9):1347–1362
Yang J, Avidis S, Sato T, Li XS (2003) Influence of vertical acceleration on soil liquefaction: new findings
and implications. In: Proceeding soil and rock America, vol 1, Cambridge, Mass
Yang J (2002) Liquefaction resistance of sand in relation to P-wave velocity. Geotechnique 52(4):295–298
Yang J, Savidis S, Roemer M (2004) Evaluating liquefaction strength of partially saturated sand. J Geotech
Geoenviron Eng ASCE 130(9):975–979
Yang J, Sato T (2000) Interpretation of seismic vertical amplification observed at an array site. Bull Seismol
Soc Am 90:275–285
Yang J, Sato T (2002) Analytical study of saturation effects on seismic vertical amplification of a soil layer.
Geotechnique 51(2):161–165
Yang J, Sato T, Savidis S, Li XS (2002) Horizontal and vertical components of ground motions at liquefiable
sites. Soil Dyn Earthq Eng 22:229–240
Yang J (2006) Frequency dependent amplification of unsaturated surface soil layer. J Geotech Geoenviron
Eng ASCE 132(4):526–531
Yegian MK, Eseller BE, Alshawabkeh A, Ali S (2007) Induced-partial saturation for liquefaction mitigation:
experimental investigation. J Geotech Geoenviron Eng ASCE 133(4):372–380
Yoshimi Y, Tanaka K, Tokimatsu K (1989) Liquefaction resistance of partially saturated sand. Soils Found
29(3):157–162

123
View publication stats

Вам также может понравиться