SlideShare a Scribd company logo
1 of 58
Download to read offline
465 
CHAPTER 12 
Copyrighted Material 
Time Effects on Strength and 
Deformation 
12.1 INTRODUCTION 
Virtually every soil ‘‘lives’’ in that all of its properties 
undergo changes with time–some insignificant, but 
others very important. Time-dependent chemical, 
geomicrobiological, and mechanical processes may 
result in compositional and structural changes that 
lead to softening, stiffening, strength loss, strength 
gain, or altered conductivity properties. The need to 
predict what the properties and behavior will be 
months to hundreds or thousands of years from now 
based on what we know today is a major challenge in 
geoengineering. Some time-dependent changes and 
their effects as they relate to soil formation, composi-tion, 
weathering, postdepositional changes in sedi-ments, 
the evolution of soil structure, and the like are 
considered in earlier chapters of this book. Emphasis 
in this chapter is on how time under stress changes the 
structural, deformation, and strength properties of 
soils, what can be learned from knowledge of these 
changes, and their quantification for predictive pur-poses. 
When soil is subjected to a constant load, it deforms 
over time, and this is usually called creep. The inverse 
phenomenon, usually termed stress relaxation, is a 
drop in stress over time after a soil is subjected to a 
particular constant strain level. Creep and relaxation 
are two consequences of the same phenomenon, that 
is, time-dependent changes in structure. The rate and 
magnitude of these time-dependent deformations are 
determined by these changes. 
Time-dependent deformations and stress relaxation 
are important in geotechnical problems wherein long-term 
behavior is of interest. These include long-term 
settlement of structures on compressible ground, de-formations 
of earth structures, movements of natural 
and excavated slopes, squeezing of soft ground around 
tunnels, and time- and stress-dependent changes in soil 
properties. The time-dependent deformation response 
of a soil may assume a variety of forms owing to the 
complex interplays among soil structure, stress history, 
drainage conditions, and changes in temperature, pres-sure, 
and biochemical environment with time. Time-dependent 
deformations and stress relaxations usually 
follow logical and often predictable patterns, at least 
for simple stress and deformation states such as uni-axial 
and triaxial compression, and they are described 
in this chapter. Incorporation of the observed behav-ior 
into simple constitutive models for analytical de-scription 
of time-dependent deformations and stress 
changes is also considered. 
Time-dependent deformation and stress phenomena 
in soils are important not only because of the imme-diate 
direct application of the results to analyses of 
practical problems, but also because the results can be 
used to obtain fundamental information about soil 
structure, interparticle bonding, and the mechanisms 
controlling the strength and deformation behavior. 
Both microscale and macroscale phenomena are dis-cussed 
because understanding of microscale processes 
can provide a rational basis for prediction of macro-scale 
responses. 
Copyright © 2005 John Wiley & Sons Retrieved from: www.knovel.com
466 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
12.2 GENERAL CHARACTERISTICS 
1. As noted in the previous section soils exhibit 
both creep1 and stress relaxation (Fig. 12.1). 
Creep is the development of time-dependent 
shear and/or volumetric strains that proceed 
at a rate controlled by the viscouslike resistance 
of soil structure. Stress relaxation is a time-dependent 
decrease in stress at constant defor-mation. 
The relationship between creep strain 
Material 
Copyrighted and the logarithm of time may be linear, con-cave 
upward, or concave downward as shown 
by the examples in Fig. 12.2. 
2. The magnitude of these effects increases with 
increasing plasticity, activity, and water content 
Figure 12.1 Creep and stress relaxation: (a) Creep under 
constant stress and (b) stress relaxation under constant strain. 
1The term creep is used herein to refer to time-dependent shear 
strains and/ or volumetric strains that develop at a rate controlled by 
the viscous resistance of the soil structure. Secondary compression 
refers to the special case of volumetric strain that follows primary 
consolidation. The rate of secondary compression is controlled by 
the viscous resistance of the soil structure, whereas, the rate of pri-mary 
consolidation is controlled by hydrodynamic lag, that is, how 
fast water can escape from the soil. 
of the soil. The most active clays usually exhibit 
the greatest time-dependent responses (i.e., 
smectite  illite  kaolinite). This is because 
the smaller the particle size, the greater is the 
specific surface, and the greater the water ad-sorption. 
Thus, under a given consolidation 
stress or deviatoric stress, the more active and 
plastic clays (smectites) will be at higher water 
content and lower density than the inactive clays 
(kaolinites). Normally consolidated soils exhibit 
larger magnitude of creep than overconsolidated 
soils. However, the basic form of behavior is 
essentially the same for all soils, that is, undis-turbed 
and remolded clay, wet clay, dry clay, 
normally and overconsolidated soil, and wet and 
dry sand. 
3. An increase in deviatoric stress level results in 
an increased rate of creep as shown in Fig. 12.1. 
Some soils may fail under a sustained creep 
stress significantly less (as little as 50 percent) 
than the peak stress measured in a shear test, 
wherein a sample is loaded to failure in a few 
minutes or hours. This is termed creep rupture, 
and an early illustration of its importance was 
the development of slope failures in the Cucar-acha 
clay shale, which began some years after 
the excavation of the Panama Canal (Casa-grande 
and Wilson, 1951). 
4. The creep response shown by the upper curve 
in Fig. 12.1 is often divided into three stages. 
Following application of a stress, there is first a 
period of transient creep during which the strain 
rate decreases with time, followed by creep at 
nearly a constant rate for some period. For ma-terials 
susceptible to creep rupture, the creep 
rate then accelerates leading to failure. These 
three stages are termed primary, secondary, and 
tertiary creep. 
5. An example of strain rates as a function of stress 
for undrained creep of remolded illite is shown 
in Fig. 12.3. At low deviator stress, creep rates 
are very small and of little practical importance. 
The curve shapes for deviator stresses up to 
about 1.0 kg/cm2 are compatible with the pre-dictions 
of rate process theory, discussed in Sec-tion 
12.4. At deviator stress approaching the 
strength of the material, the strain rates become 
very large and signal the onset of failure. 
6. A characteristic relationship between strain rate 
and time exists for most soils, as shown, for 
example, in Fig. 12.4 for drained triaxial com-pression 
creep of London clay (Bishop, 1966) 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
GENERAL CHARACTERISTICS 467 
Material 
Copyrighted Figure 12.2 Sustained stress creep curves illustrating different forms of strain vs. logarithm 
of time behavior. 
and Fig. 12.5 for undrained triaxial compression 
creep of soft Osaka clay (Murayama and Shi-bata, 
1958). At any stress level (shown as a per-centage 
of the strength before creep in Fig. 12.4 
and in kg/cm2 in Fig. 12.5), the logarithm of the 
strain rate decreases linearly with increase in the 
logarithm of time. The slope of this relationship 
is essentially independent of the creep stress; 
increases in stress level shift the line vertically 
upward. The slope of the log strain rate versus 
log time line for drained creep is approximately 
1. Undrained creep often results in a slope be- 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
468 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
Material 
Copyrighted Figure 12.3 Variation of creep strain rate with deviator stress for undrained creep of re-molded 
illite. 
tween 0.8 and 1 for this relationship. The 
onset of failure under higher stresses is signaled 
by a reversal in slope, as shown by the topmost 
curve in Fig. 12.5. 
7. Pore pressure may increase, decrease, or remain 
constant during creep, depending on the volume 
change tendencies of the soil structure and 
whether or not drainage occurs during the de-formation 
process. In general, saturated soft 
sensitive clays under undrained conditions are 
most susceptible to strength loss during creep 
due to reduction in effective stress caused by 
increase in pore water pressure with time. Heav-ily 
overconsolidated clays under drained con-ditions 
are also susceptible to creep rupture due 
to softening associated with the increase in wa-ter 
content by dilation and swelling. 
8. Although stress relaxation has been less studied 
than creep, it appears that equally regular pat-terns 
of deformation behavior are observed, for 
example, Larcerda and Houston (1973). 
9. Deformation under sustained stress ordinarily 
produces an increase in stiffness under the ac-tion 
of subsequent stress increase, as shown 
schematically in Fig. 12.6. This reflects the 
time-dependent structural readjustment or ‘‘ag-ing’’ 
that follows changes in stress state. It is 
analogous to the quasi-preconsolidation effect 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
GENERAL CHARACTERISTICS 469 
Material 
Copyrighted Figure 12.4 Strain rate vs. time relationships during drained creep of London clay (data 
from Bishop, 1966). 
due to secondary compression discussed in Sec-tion 
8.11; however, it may develop under un-drained 
as well as drained conditions. 
10. As shown in Fig. 12.7, the locations of both the 
virgin compression line and the value of the pre-consolidation 
pressure, , determined in the p 
laboratory are influenced by the rate of loading 
during one-dimensional consolidation (Graham 
et al., 1983a; Leroueil et al., 1985). Thus, esti-mations 
of the overconsolidation ratio of clay 
deposits in the field are dependent on the load-ing 
rates and paths used in laboratory tests for 
determination of the preconsolidation pressure. 
If it is assumed that the relationship between 
strain and logarithm of time during compression 
is linear over the time ranges of interest and that 
the secondary compression index Ce is constant 
regardless of load, the rate-dependent precon-solidation 
pressure p at ˙1 can be related to the 
axial strain rate as follows (Silvestri et al, 1986; 
Soga and Mitchell, 1996; Leroueil and Marques, 
1996): 
Ce / (CcCr)  p ˙1 ˙1       (12.1) p(ref) ˙1(ref) ˙1(ref) 
where Cc is the virgin compression index, Cr is 
the recompression index and p(ref) is the pre-consolidation 
pressure at a reference strain rate 
˙ . In this equation, the rate effect increases 1(ref) 
with the value of   Ce/(Cc  Cr). The var-iation 
of preconsolidation pressure with strain 
rate is shown in Fig. 12.8 (Soga and Mitchell, 
1996). The data define straight lines, and the 
slope of the lines gives the parameter . In gen-eral, 
the value of  ranges between 0.011 and 
0.094. Leroueil and Marques (1996) report val-ues 
between 0.029 and 0.059 for inorganic 
clays. 
11. The undrained strength of saturated clay in-creases 
with increase in rate of strain, as shown 
in Figs. 12.9 and 12.10. The magnitude of the 
effect is about 10 percent for each order of mag-nitude 
increase in the strain rate. The strain rate 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
470 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
Material 
Figure 12.5 Strain rate vs. time relationships during undrained creep of Osaka alluvial clay 
Copyrighted (Murayama and Shibata, 1958). 
Figure 12.6 Effect of sustained loading on (a) stress–strain 
and strength behavior and (b) one-dimensional compression 
behavior. 
effect is considerably smaller for sands. In a 
manner similar to Eq. (12.1), a rate parameter 
 can be defined as the slope of a log–log plot 
of deviator stress at failure qƒ at a particular 
strain rate ˙1 relative to qƒ(ref), the strength at a 
reference strain rate ˙ , versus strain rate. 1(ref) 
This gives the following equation: 
 qƒ ˙1    (12.2) qƒ(ref) ˙1(ref) 
The value of  ranges between 0.018 and 0.087, 
similar to the  rate parameter values used to 
define the rate effect on consolidation pressure 
in Eq. (12.1). Higher values of  are associated 
with more metastable soil structures (Soga and 
Mitchell, 1996). Rate dependency decreases 
with increasing sample disturbance, which is 
consistent with this finding. 
12.3 TIME-DEPENDENT 
DEFORMATION–STRUCTURE INTERACTION 
In reality, completely smooth curves of the type shown 
in the preceding figures for strain and strain rate as a 
function of time may not exist at all. Rather, as dis- 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
TIME-DEPENDENT DEFORMATION–STRUCTURE INTERACTION 471 
Copyrighted Material 
Figure 12.7 Rate dependency on one-dimensional compression characteristics of Batiscan 
clay: (a) compression curves and (b) preconsolidation pressure (Leroueil et al., 1985). 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
472 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
Material 
Figure 12.8 Strain rate dependence on preconsolidation pressure determined from one-dimensional 
constant strain rate tests (Soga and Mitchell, 1996). 
Copyrighted Figure 12.9 Effect of strain rate on undrained strength (Kulhawy and Mayne 1990). Re-printed 
with permission from EPRI. 
cussed by Ter-Stepanian (1992), a ‘‘jump-like structure 
reorganization’’ may occur, reflecting a stochastic char-acter 
for the deformation, as shown in Fig. 12.11 for 
creep of an undisturbed diatomaceous, lacustrine, ov-erconsolidated 
clay. Ter-Stepanian (1992) suggests that 
there are four levels of deformation: (1) the molecular 
level, which consists of displacement of flow units by 
surmounting energy barriers, (2) mutual displacement 
of particles as a result of bond failures, but without 
rearrangement, (3) the structural level of soil defor-mation 
involving mutual rearrangements of particles, 
and (4) deformation at the aggregate level. Behavior at 
levels 3 and 4 is discussed below; that at levels 1 and 
2 is treated in more detail in Section 12.4. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
TIME-DEPENDENT DEFORMATION–STRUCTURE INTERACTION 473 
Material 
Figure 12.10 Strain rate dependence on undrained shear strength determined using constant 
strain rate CU tests (Soga and Mitchell, 1996). 
Copyrighted Figure 12.11 Nonuniformity of creep in an undisturbed, di-atomaceous, 
lacustrine, overconsolidated clay (from Ter- 
Stepanian, 1992). 
Time-Dependent Process of Particle Rearrangement 
Creep can lead to rearrangement of particles into more 
stable configurations. Forces at interparticle contacts 
have both normal and tangential components, even if 
the macroscopic applied stress is isotropic. If, during 
the creep process, there is an increase in the proportion 
of applied deviator stress that is carried by interparticle 
normal forces relative to interparticle tangential forces, 
then the creep rate will decrease. Hence, the rate at 
which deformation level 3 occurs need not be uniform 
owing to the particulate nature of soils. Instead it will 
reflect a series of structural readjustments as particles 
move up, over, and around each other, thus leading to 
the somewhat irregular sequence of data points shown 
in Fig. 12.11. 
Microscopically, creep is likely to occur in the weak 
clusters discussed in Section 11.6 because the contacts 
in them are at limiting frictional equilibrium. Any 
small perturbation in applied load at the contacts or 
time-dependent loss in material strength can lead to 
sliding, breakage or yield at asperities. As particles 
slip, propped strong-force network columns are dis-turbed, 
and these buckle via particle rolling as dis-cussed 
in Section 11.6. 
To examine the effects of particle rearrangement, 
Kuhn (1987) developed a discrete element model that 
considers sliding at interparticle contacts to be visco-frictional. 
The rate at which sliding of two particles 
relative to each other occurs depends on the ratio of 
shear to normal force at their contact. The relationship 
between rate and force is formulated in terms of rate 
process theory (see Section 12.4), and the mechanistic 
representations of the contact normal and shear forces 
are shown in Fig. 12.12. The time-dependent compo-nent 
in the tangential forces model is given as a ‘‘sinh-dashpot’’. 
2 The average magnitudes of both normal and 
2Kuhn (1987) used the following equation for rate of sliding at a 
contact: 
2kT 
F  ƒt ˙X 
  exp   sinh   h RT 2kTn ƒn 1 
where n1 is the number of bonds per unit of normal force, ƒt is the 
tangential force and ƒn is the normal force. The others are parameters 
related to rate process theory as described in Section 12.4. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
474 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
Copyrighted Material 
Figure 12.12 Normal and tangential interparticle force mod-els 
according to Kuhn (1987). 
Figure 12.13 Creep curves developed by numerical analysis 
of an irregular packing of circular disks (from Kuhn and 
Mitchell, 1993). 
tangential forces at individual contacts can change dur-ing 
deformation even though the applied boundary 
stresses are constant. Small changes in the tangential 
and normal force ratio at a contact can have a very 
large influence on the sliding rate at that contact. These 
changes, when summed over all contacts in the shear 
zone, result in a decrease or increase in the overall 
creep rate. 
A numerical analysis of an irregular packing of cir-cular 
disks using the sinh-dashpot representation gives 
creep behavior comparable to that of many soils as 
shown in Fig. 12.13 (Kuhn and Mitchell, 1993). The 
creep rate slows if the average ratio of tangential to 
normal force decreases, whereas it accelerates and may 
ultimately lead to failure if the ratio increases. In some 
cases, the structural changes that are responsible for 
the decreasing strain rate and increased stiffness may 
cause the overall soil structure to become more meta-stable. 
Then, after the strain reaches some limiting 
value, the process of contact force transfer from de-creasing 
tangential to increasing normal force reverses. 
This marks the onset of creep rupture as the structure 
begins to collapse. A similar result was obtained by 
Rothenburg (1992) who performed discrete particle 
simulations in which smooth elliptical particles were 
cemented with a model exhibiting viscous character-istics 
in both normal and tangential directions. 
Particle Breakage During Creep 
Particle breakage can contribute to time-dependent de-formation 
of sands (Leung et al., 1996; Takei et al., 
2001; McDowell, 2003). Leung et al. (1996) performed 
one-dimensional compression tests on sands, and Fig. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
TIME-DEPENDENT DEFORMATION–STRUCTURE INTERACTION 475 
Dry sand 
RD = 75% 
Pressure = 15.4 MPa 

 
After 
5 days After 
Before 
290 s 

 
Test 
Material 
Copyrighted   
Sieve Size (μm) 
100 
80 
60 
40 
20 
0 
0 

 
 
 
100 200 300 400 

 
Percentage Passing (%) 
Figure 12.14 Changes in particle size distribution of sand 
before loading and after two different load durations (from 
Leung, et al., 1996). 
12.14 shows the particle size distribution curves for 
samples before loading and after two different load du-rations. 
The amount of particle breakage increased 
with load duration. Microscopic observations revealed 
that angular protrusions of the grains were ground off, 
producing fines. The fines fill the voids between larger 
particles and crushed particles progressively rear-ranged 
themselves with time. 
Aging—Time-Dependent Strengthening of Soil 
Structure 
The structural changes that occur during creep that is 
continuing at a decreasing rate cause an increase in 
soil stiffness when the soil is subjected to further stress 
increase as shown in Fig. 12.6. Leonards and Alt-schaeffl 
(1964) showed that this increase in preconso-lidation 
pressure cannot be accounted for in terms of 
the void ratio decrease during the sustained compres-sion 
period. Time-dependent changes of these types are 
a consequence of ‘‘aging’’ effects, which alter the 
structural state of the soil. The fabric obtained by creep 
may be different from that caused by increase in stress, 
even though both samples arrive at the same void ratio. 
Leroueil et al. (1996) report a similar result for an ar-tificially 
sedimented clay from Quebec, as shown in 
Fig. 12.15a. They also measured the shear wave ve-locities 
after different times during the tests using 
bender elements and computed the small strain elastic 
shear modulus. Figure 12.15b shows the change in 
shear modulus with void ratio. 
Additional insight into the structural changes ac-companying 
the aging of clays is provided by the re-sults 
of studies by Anderson and Stokoe (1978) and 
Nakagawa et al. (1995). Figure 12.16 shows changes 
in shear modulus with time under a constant confining 
pressure for kaolinite clay during consolidation (An-derson 
and Stokoe, 1978). Two distinct phases of shear 
modulus–time response are evident. During primary 
consolidation, values of the shear modulus increase 
rapidly at the beginning and begin to level off as the 
excess pore pressure dissipates. After the end of pri-mary 
consolidation, the modulus increases linearly 
with the logarithm of time during secondary compres-sion. 
The expected change in shear modulus due to void 
ratio change during secondary compression can be es-timated 
using the following empirical formula for 
shear modulus as a function of void ratio and confining 
pressure (Hardin and Black, 1968): 
(2.97  e)2 G  A p0.5 (12.3) 
1  e 
where A is a unit dependant material constant, e is the 
void ratio, and p is the mean effective stress. The 
dashed line in Fig. 12.16 shows the calculated in-creases 
in the shear modulus due to void ratio decrease 
using Eq. (12.3). It is evident that the change in void 
ratio alone does not provide an explanation for the sec-ondary 
time-dependent increase in shear modulus. This 
aging effect has been recorded for a variety of mate-rials, 
ranging from clean sands to natural clays (Afifi 
and Richart, 1973; Kokusho, 1987; Mesri et al., 1990, 
and many others). Further discussion of aging phenom-ena 
is given in Section 12.11. 
Time-Dependent Changes in Soil Fabric 
Changes in soil fabric with time under stress influence 
the stability of soil structure. Changes in sand fabric 
with time after load application in one-dimensional 
compression were measured by Bowman and Soga 
(2003). Resin was used to fix sand particles after var-ious 
loading times. Pluviation of the sand produced a 
horizontal preferred particle orientation of soil grains, 
and increased vertical loading resulted in a greater ori-entation 
of particle long axes parallel to the horizontal, 
which is in agreement with the findings of Oda (1972a, 
b, c), Mitchell et al. (1976), and Jang and Frost (1998). 
Over time, however, the loading of sand caused parti-cle 
long axes to rotate toward the vertical direction 
(i.e., more isotropic fabric). 
Experimental evidence (Bowman and Soga, 2003) 
showed that large voids became larger, whereas small 
voids became smaller, and particles group or cluster 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
476 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
2.6 
2.4 
2.2 
2.0 
Material 
1.8 
Copyrighted A A 
B B 
C C 
D D 
4 6 8 10 20 
Vertical Effective Stress σv (kPa) 
0.5 1 
2 
E 
Small-strain Stiffness G0 (MPa) 
F 
E 
F 
Stiffness Change 
During the Primary 
Consolidation 
Between B and C 
Normal 
Consolidation Line 
Creep for 
120 days 
Increase in Stiffness 
during Creep (C-D) 
Quasi-preconsolidation 
Pressure 
Destructuring 
State 
(a) (b ) 
5 
Void Ratio e 
Figure 12.15 (a) Compression curve and (b) variation of the maximum shear modulus G0 
with void ratio for artificially sedimented Jonquiere clay (from Leroueil et al., 1996). 
together with time. Based on these particulate level 
findings, it appears that the movements of particles 
lead to interlocking zones of greater local density. The 
interlocked state may be regarded as the final state of 
any one particle under a particular applied load, due to 
kinematic restraint. The result, with time, is a stiffer, 
more efficient, load-bearing structure, with areas of rel-atively 
large voids and neighboring areas of tightly 
packed particles. The increase in stiffness is achieved 
by shear connections obtained by the clustering. Then, 
when load is applied, the increased stiffness and 
strength of the granular structure provides greater re-sistance 
to the load and the observed aging effect is 
seen. The numerical analysis in Kuhn and Mitchell 
(1993) led to a similar hypothesis for how a more 
‘‘braced’’ structure develops with time. For load appli-cation 
in a direction different to that during the aging 
period, however, the strengthening effect of aging may 
be less, as the load-bearing particle column direction 
differs from the load direction. 
Time-Dependent Changes in Physicochemical 
Interaction of Clay and Pore Fluid 
A portion of the shear modulus increase during sec-ondary 
compression of clays is believed to result from 
a strengthening of physicochemical bonds between 
particles. To illustrate this, Nakagawa et al. (1995) ex-amined 
the physicochemical interactions between clays 
and pore fluid using a special consolidometer in which 
the sample resistivity and pore fluid conductivity could 
be measured. Shear wave velocities were obtained us-ing 
bender elements to determine changes in the stiff-ness 
characteristics of the clay during consolidation. 
Kaolinite clay mixed with saltwater was used for the 
experiment, and changes in shear wave velocities and 
electrical properties were monitored during the tests. 
The test results showed that the pore fluid compo-sition 
and ion mobility changed with time. At each 
load increment, as the effective stress increased with 
pore pressure dissipation, the shear wave velocities, 
and therefore the shear modulus, generally increased 
with time as shown in Fig. 12.17. It may be seen, how-ever, 
that in some cases, the shear wave velocities at 
the beginning of primary consolidation decreased 
slightly from the velocities obtained immediately be-fore 
application of the incremental load, probably as a 
result of soil structure breakdown. During the subse-quent 
secondary compression stage, the shear wave ve-locity 
again increased. As was the case for the results 
in Fig.12.16, the increases in shear wave velocity dur- 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
TIME-DEPENDENT DEFORMATION–STRUCTURE INTERACTION 477 
50 
40 
30 
20 
10 
0 
0.5 
1.0 
1.5 
2.0 
IG = ΔG per log time 
= 6.2 MPa 
Copyrighted Material 
Primary Consolidation Secondary Compression 
Sample Height Change (mm) 
101 102 103 104 
Time (min) 
100 101 102 103 104 
Time (min) 
Ball Kaolinite 
Possible Change in G 
by Void Ratio Decrease 
Only Estimated Using 
Eq. (12.3) 
Initial Consolidation Pressure = 70 kPa 
Initial Void Ratio e0 = 1.1 
Shear Modulus of less than γ = 10-3% (MPa) 
Figure 12.16 Modulus and height changes as a function of 
time under constant confining pressure for kaolinite: (a) shear 
modulus and (b) height change (from Anderson and Stokoe, 
1978). 
Figure 12.17 Changes in shear wave velocity during pri-mary 
consolidation and secondary compression of kaolinite. 
Consolidation pressures: (a) 11.8 kPa and (b) 190 kPa (from 
Nakagawa et al., 1995). 
ing secondary compression are greater than can be ac-counted 
for by increase in density. 
The electrical conductivity of the sample measured 
by filter electrodes increased during the early stages of 
consolidation, but then decreased continuously there-after 
as shown in Fig. 12.18. The electrical conductiv-ity 
is dominated by flow through the electrolyte 
solution in the pores. During the initial compression, a 
breakdown of structure releases ions into the pore wa-ter, 
increasing the electrical conductivity. With time, 
the conductivity decreased, suggesting that the released 
ions are accumulating near particle surfaces. Some of 
these released ions are expelled from the specimen as 
consolidation progressed as shown in Fig. 12.18b. A 
slow equilibrium under a new state of effective stress 
is hypothesized to develop that involves both small 
particle rearrangements, associated with decrease in 
void ratio during secondary compression, and devel-opment 
of increased contact strength as a result of pre-cipitation 
of salts from the pore water and/or other 
processes. 
Primary consolidation can be considered a result of 
drainage of pore water fluid from the macropores, 
whereas secondary compression is related to the de-layed 
deformation of micropores in the clay aggregates 
(Berry and Poskitt, 1972; Matsuo and Kamon, 1977; 
Sills, 1995). The mobility of water in the micropores 
is restricted due to small pore size and physicochem-ical 
interactions close to the clay particle surfaces. Ak-agi 
(1994) did compression tests on specially prepared 
clay containing primarily Ca in the micropores and Na 
in the macropores. Concentrations of the two ions in 
the expelled water at different times after the start of 
consolidation were consistent with this hypothesis. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
478 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
Activation Frequency 
The energy to enable a flow unit to cross a barrier may 
be provided by thermal energy and by various applied 
potentials. For a material at rest, the potential energy– 
displacement relationship is represented by curve A in 
Fig. 12.21. From statistical mechanics it is known that 
Copyrighted Material 
Figure 12.18 Changes in electrical conductivity of the pore 
water during primary consolidation and secondary compres-sion 
of kaolinite. Consolidation pressures: (a) 95 kPa and (b) 
190 kPa (from Nakagawa et al., 1995). 
Figure 12.19 Energy barriers and activation energy. 
12.4 SOIL DEFORMATION AS A RATE 
PROCESS 
Deformation and shear failure of soil involve time-dependent 
rearrangement of matter. As such, these 
phenomena are amenable for study as rate processes 
through application of the theory of absolute reaction 
rates (Glasstone et al., 1941). This theory provides 
both insights into the fundamental nature of soil 
strength and functional forms for the influences of sev-eral 
factors on soil behavior. 
Detailed development of the theory, which is based 
on statistical mechanics, may be found in Eyring 
(1936), Glasstone et al. (1941), and elsewhere in the 
physical chemistry literature. Adaptations to the study 
of soil behavior include those by Abdel-Hady and Her-rin 
(1966), Andersland and Douglas (1970), Christen-sen 
and Wu (1964), Mitchell (1964), Mitchell et al. 
(1968, 1969), Murayama and Shibata (1958, 1961, 
1964), Noble and Demirel (1969), Wu et al. (1966), 
Keedwell (1984), Feda (1989, 1992), and Kuhn and 
Mitchell (1993). 
Concept of Activation 
The basis of rate process theory is that atoms, mole-cules, 
and/or particles participating in a time-dependent 
flow or deformation process, termed flow 
units, are constrained from movement relative to each 
other by energy barriers separating adjacent equilib-rium 
positions, as shown schematically by Fig. 12.19. 
The displacement of flow units to new positions re-quires 
the acquisition of an activation energy 
F of 
sufficient magnitude to surmount the barrier. The po-tential 
energy of a flow unit may be the same following 
the activation process, or higher or lower than it was 
initially. These conditions are shown by analogy with 
the rotation of three blocks in Fig. 12.20. In each case, 
an energy barrier must be crossed. The assumption of 
a steady-state condition is implicit in most applications 
to soils concerning the at-rest barrier height between 
successive equilibrium positions. 
The magnitude of the activation energy depends on 
the material and the type of process. For example, val-ues 
of 
F for viscous flow of water, chemical reac-tions, 
and solid-state diffusion of atoms in silicates are 
about 12 to 17, 40 to 400, and 100 to 150 kJ/mol of 
flow units, respectively. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
SOIL DEFORMATION AS A RATE PROCESS 479 
Material 
Figure 12.20 Examples of activated processes: (a) steady-state, (b) increased stability, and 
(c) decreased stability. 
Copyrighted Figure 12.21 Effect of a shear force on energy barriers. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
480 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
the average thermal energy per flow unit is kT, where 
k is Boltzmann’s constant (1.38  1023 J K1) and T 
is the absolute temperature (K). Even in a material at 
rest, thermal vibrations occur at a frequency given by 
kT/h, where h is Planck’s constant (6.624  1034 J 
s1). The actual thermal energies are divided among 
the flow units according to a Boltzmann distribution. 
It may be shown that the probability of a given unit 
becoming activated, or the proportion of flow units that 
are activated during any one oscillation is given by 
 
F 
p( 
F)  exp (12.4) NkT 
Material 
Copyrighted where N is Avogadro’s number (6.02  1023), and Nk 
is equal to R, the universal gas constant (8.3144 J K1 
mol1). The frequency of activation  then is 
kT  
F 
   exp (12.5) h NkT 
In the absence of directional potentials, energy bar-riers 
are crossed with equal frequency in all directions, 
and no consequences of thermal activations are ob-served 
unless the temperature is sufficiently high that 
softening, melting, or evaporation occurs. If, however, 
a directed potential, such as a shear stress, is applied, 
then the barrier heights become distorted as shown by 
curve B in Fig. 12.21. If ƒ represents the force acting 
on a flow unit, then the barrier height is reduced by an 
amount (ƒ/2) in the direction of the force and in-creased 
by a like amount in the opposite direction, 
where  represents the distance between successive 
equilibrium positions.3 Minimums in the energy curve 
are displaced a distance  from their original positions, 
representing an elastic distortion of the material struc-ture. 
The reduced barrier height in the direction of force 
ƒ increases the activation frequency in that direction to 
kT  
F/N  ƒ/2 
  →  exp (12.6) h kT 
and in the opposite direction, the increased barrier 
height decreases the activation frequency to 
kT  
F/N  ƒ/2 
  ←  exp (12.7) h kT 
3Work (ƒ / 2) done by the force ƒ as the flow unit drops from the 
peak of the energy barrier to a new equilibrium position is assumed 
to be given up as heat. 
The net frequency of activation in the direction of 
the force then becomes 
kT 
F (→) (←) 2 exp    ƒ 
     sinh h RT 2kT 
(12.8) 
Strain Rate Equation 
At any instant, some of the activated flow units may 
successfully cross the barrier; others may fall back into 
their original positions. For each unit that is successful 
in crossing the barrier, there will be a displacement . 
The component of  in a given direction times the 
number of successful jumps per unit time gives the rate 
of movement per unit time. If this rate of movement 
is expressed on a per unit length basis, then the strain 
rate ˙ is obtained. 
Let X  F (proportion of successful barrier cross-ings 
and ) such that 
˙  X[( →)  ( ←)] (12.9) 
Then from Eq. (12.8) 
kT  
F   ƒ 
 ˙  2X exp sinh (12.10) h RT 2kT 
The parameter X may be both time and structure de-pendent. 
If (ƒ/2kT)  1, then sinh(ƒ/2kT)  (ƒ/2kT), and 
the rate is directly proportional to ƒ. This is the case 
for ordinary Newtonian fluid flow and diffusion where 
1 
˙ 	   (12.11) 
 
where 	˙ is the shear strain rate,  is dynamic viscosity, 
and  is shear stress. 
For most solid deformation problems, however, 
(ƒ/2kT)  1 (Mitchell et al., 1968), so 
ƒ 1 ƒ 
sinh   exp  (12.12) 2kT 2 2kT 
and 
kT  
F   ƒN 
 ˙  X expexp(12.13) h RT 2RT 
Equation (12.13) applies except for very small stress 
intensities, where the exponential approximation of the 
hyperbolic sine is not justified. Equations (12.10) and 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
BONDING, EFFECTIVE STRESSES, AND STRENGTH 481 
(12.13) or comparable forms have been used to obtain 
dashpot coefficients for rheological models, to obtain 
functional forms for the influences of different factors 
on strength and deformation rate, and to study defor-mation 
rates in soils. For example, Kuhn and Mitchell 
(1993) used this form as part of the particle contact 
law in discrete element modeling as described in the 
previous section. Puzrin and Houlsby (2003) used it as 
an internal function of a thermomechanical-based 
model and derived a rate-dependent constitutive model 
for soils. 
Copyrighted Material 
Soil Deformation as a Rate Process 
Although there does not yet appear to be a rigorous 
proof of the correctness of the detailed statistical me-chanics 
formulation of rate process theory, even for 
simple chemical reactions, the real behavior of many 
systems has been substantially in accord with it. Dif-ferent 
parts of Eq. (12.13) have been tested separately 
(Mitchell et al., 1968). It was found that the tempera-ture 
dependence of creep rate and the stress depend-ence 
of the experimental activation energy [Eq. 
(12.14)] were in accord with predictions. These results 
do not prove the correctness of the theory; they do, 
however, support the concept that soil deformation is 
a thermally activated process. 
Arrhenius Equation 
Equation (12.13) may be written 
kT E 
˙ X exp   (12.14) h RT 
where 
ƒN 
E  
F  (12.15) 
2 
is termed the experimental activation energy. For all 
conditions constant except T, and assuming that 
X(kT/h)  constant  A, 
 E 
 ˙  A exp (12.16) RT 
Equation (12.16) is the same as the well-known em-pirical 
equation proposed by Arrhenius around 1900 
to describe the temperature dependence of chemical 
reaction rates. It has been found suitable also for 
characterization of the temperature dependence of 
processes such as creep, stress relaxation, secondary 
compression, thixotropic strength gain, diffusion, and 
fluid flow. 
12.5 BONDING, EFFECTIVE STRESSES, AND 
STRENGTH 
Using rate process theory, the results of time-dependent 
stress–deformation measurements in soils 
can be used to obtain fundamental information about 
soil structure, interparticle bonding, and the mecha-nisms 
controlling strength and deformation behavior. 
Deformation Parameters from Creep Test Data 
If the shear stress on a material is  and it is distributed 
uniformly among S flow units per unit area, then 
 
ƒ  (12.17) 
S 
Displacement of a flow unit requires that interatomic 
or intermolecular forces be overcome so that it can be 
moved. Let it be assumed that the number of flow units 
and the number of interparticle bonds are equal. 
If D represents the deviator stress under triaxial 
stress conditions, the value of ƒ on the plane of max-imum 
shear stress is 
D 
ƒ  (12.18) 
2S 
so Eq. (12.13) becomes 
kT  
F   D 
 ˙  X expexp(12.19) h RT 4SkT 
This equation describes creep as a steady-state 
process. Soils do not creep at constant rate, however, 
because of continued structural changes during de-formation 
as described in Section 12.3, except for 
the special case of large deformations after mobiliza-tion 
of full strength. Thus, care must be taken in ap-plication 
of Eq. (12.19) to ensure that comparisons of 
creep rates and evaluations of the influences of differ-ent 
factors are made under conditions of equal struc-ture. 
The time dependency of creep rate and the 
possible time dependencies of the parameters in Eq. 
(12.19) are considered in Section 12.8. 
Determination of Activation Energy From Eq. 
(12.14) 
 ln(˙/T) E 
  (12.20) 
(1/T) R 
provided strain rates are considered under conditions 
of unchanged soil structure. Thus, the value of E can 
be determined from the slope of a plot of ln(˙ /T) ver-sus 
(1/T). Procedures for evaluation of strain rates for 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
482 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
soils at different temperatures but at the same structure 
are given by Mitchell et al. (1968, 1969). 
Determination of Number of Bonds For stresses 
large enough to justify approximating the hyperbolic 
sine function by a simple exponential in the creep rate 
equation and small enough to avoid tertiary creep, the 
logarithm of strain rate varies directly with the deviator 
stress. For this case, Eq. (12.19) can be written 
 ˙ K(t) exp(D) (12.21) 
Material 
4A procedure for evaluation of  from the results of a test at a 
succession of stress levels on a single sample is given by Mitchell 
Copyrighted et al. (1969). where 
kT 
F 
K(t)  X exp  (12.22) h RT 
 
  (12.23) 
4SkT 
Parameter  is a constant for a given value of ef-fective 
consolidation pressure and is given by the slope 
of the relationship between log strain rate and stress. 
It is evaluated using strain rates at the same time after 
the start of creep tests at several stress intensities. With 
 known, /S is calculated as a measure of the number 
of interparticle bonds.4 
Activation Energies for Soil Creep 
Activation energies for the creep of several soils and 
other materials are given in Table 12.1. The free energy 
of activation for creep of soils is in the range of about 
80 to 180 kJ/mol. Four features of the values for soils 
in Table 12.1 are significant: 
1. The activation energies are relatively large, much 
2. Variations in water content (including complete 
3. The values for sand and clay are about the same. 
4. Clays in suspension with insufficient solids to 
Table 12.1 Activation Energies for Creep of Several Materials 
Material 
higher than for viscous flow of water. 
drying), adsorbed cation type, consolidation pres-sure, 
void ratio, and pore fluid have no significant 
effect on the required activation energy. 
form a continuous structure deform with an ac-tivation 
Activation Energy 
energy equal to that of water. 
(kJ/ mol)a Reference 
1. Remolded illite, saturated, water contents of 
30 to 43% 
105–165 Mitchell, et al. (1969) 
2. Dried illite: samples air-dried from 
saturation, then evacuated 
155 Mitchell, et al. (1969) 
3. San Francisco Bay mud, undisturbed 105–135 Mitchell, et al. (1969) 
4. Dry Sacramento River sand 105 Mitchell, et al. (1969) 
5. Water 16–21 Glasstone, et al. (1941) 
6. Plastics 30–60 Ree and Eyring (1958) 
7. Montmorillonite–water paste, dilute 84–109 Ripple and Day (1966) 
8. Soil asphalt 113 Abdel-Hady and Herrin (1966) 
9. Lake clay, undisturbed and remolded 96–113 Christensen and Wu (1964) 
10. Osaka clay, overconsolidated 120–134 Murayama and Shibata (1961) 
11. Concrete 226 Polivka and Best (1960) 
12. Metals 210 Finnie and Heller (1959) 
13. Frozen soils 393 Andersland and Akili (1967) 
14. Sault Ste. Marie clay, suspensions, 
discontinuous structures 
Same as 
water 
Andersland and Douglas (1970) 
15. Sault Ste. Marie clay, Li, Na, K forms, 
in H2O and CCl4, consolidated 
117 Andersland and Douglas (1970) 
aThe first four values are experimental activation energies, E. Whether the remainder are values of 
F or E is not 
always clear in the references cited. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
BONDING, EFFECTIVE STRESSES, AND STRENGTH 483 
Number of Interparticle Bonds 
Evaluation of S requires knowledge of , the separation 
distance between successive equilibrium positions in 
the interparticle contact structure. A value of 0.28 nm 
(2.8 A˚ 
) has been assumed because it is the same as the 
distance separating atomic valleys in the surface of a 
silicate mineral. It is hypothesized that deformation in-volves 
the displacement of oxygen atoms along con-tacting 
particle surfaces, as well as periodic rupture of 
Copyrighted Material 
bonds at interparticle contacts. Figure 12.22 shows this 
interpretation for  schematically. If the above as-sumption 
for  is incorrect, calculated values of S will 
still be in the same correct relative proportion as long 
as  remains constant during deformation. 
Normally Consolidated Clay Results of creep tests 
at different stress intensities for different consolidation 
pressures enable computation of S as a function of con-solidation 
pressure. Values obtained for undisturbed 
San Francisco Bay mud are shown in Fig. 12.23. The 
open point is for remolded bay mud. An undisturbed 
specimen was consolidated to 400 kPa, rebounded to 
Figure 12.22 Interpretation of  in terms of silicate mineral 
surface structure. 
50 kPa, and then remolded at constant water content. 
The effective consolidation pressure dropped to 25 kPa 
as a result of the remolding. The drop in effective 
stress was accompanied by a corresponding decrease 
in the number of interparticle bonds. Tests on remolded 
illite gave comparable results. A continuous inverse re-lationship 
between the number of bonds and water 
content over a range of water contents from more than 
40 percent to air-dried and vacuum-desiccated clay is 
shown in Fig. 12.24. The dried material had a water 
content of 1 percent on the usual oven-dried basis. The 
very large number of bonds developed by drying is 
responsible for the high dry strength of clay. 
Overconsolidated Clay Samples of undisturbed 
San Francisco Bay mud were prepared to overcon-solidation 
ratios of 1, 2, 4, and 8 following the stress 
paths shown in the upper part of Fig. 12.25. The sam-ple 
represented by the triangular data point was re-molded 
after consolidation and unloading to point d, 
where it had a water content of 52.3 percent. The un-drained 
compressive strength as a function of consol-idation 
pressure is shown in the middle section of Fig. 
12.25, and the number of bonds, deduced from the 
creep tests, is shown in the lower part of the figure. 
The effect of overconsolidation is to increase the num-ber 
of interparticle bonds over the values for normally 
consolidated clay. Some of the bonds formed during 
consolidation are retained after removal of much of the 
consolidation pressure. 
Values of compressive strength and numbers of 
bonds from Fig. 12.25 are replotted versus each other 
in Fig. 12.26. The resulting relationship suggests that 
strength depends only on the number of bonds and is 
independent of whether the clay is undisturbed, re-molded, 
normally consolidated, or overconsolidated. 
Dry Sand Creep tests on oven-dried sand yielded 
results of the same type as obtained for clay, as shown 
in Fig. 12.27, suggesting that the strength-generating 
and creep-controlling mechanisms may be similar for 
both types of material. 
Composite Strength-Bonding Relationship Values 
of S and strength for many soils are combined in Fig. 
12.28. The same proportionality exists for all the ma-terials, 
which may seem surprising, but which in reality 
should be expected, as discussed further later. 
Significance of Activation Energy and Bond Number 
Values 
The following aspects of activation energies and num-bers 
of interparticle bonds are important in the under-standing 
of the deformation and strength behavior of 
uncemented soils. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
484 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
Material 
Figure 12.23 Number of interparticle bonds as a function of consolidation pressure for 
normally consolidated San Francisco Bay mud. 
Copyrighted Figure 12.24 Number of bonds as a function of water con-tent 
for illite. 
1. The values of activation energy for deformation 
of soils are high in comparison with other ma-terials 
and indicate breaking of strong bonds. 
2. Similar creep behavior for wet and dry clay and 
for wet and dry sand indicates that deformation 
is not controlled by viscous flow of water. 
3. Comparable values of activation energy for wet 
and dry soil indicate that water is not respon-sible 
for bonding. 
4. Comparable values of activation energy for clay 
and sand support the concept that interparticle 
bond strengths are the same for both types of 
material. This is supported also by the unique-ness 
of the strength versus number of bonds re-lationship 
for all soils. 
5. The activation energy and presumably, there-fore, 
the bonding type are independent of con-solidation 
pressure, void ratio, and water 
content. 
6. The number of bonds is directly proportional to 
effective consolidation pressure for normally 
consolidated clays. 
7. Overconsolidation leads to more bonds than in 
normally consolidated clay at the same effective 
consolidation pressure. 
8. Strength depends only on the number of bonds. 
9. Remolding at constant water content causes a 
decrease in the effective consolidation pressure, 
which means also a decrease in the number of 
bonds. 
10. There are about 100 times as many bonds in dry 
clay as in wet clay. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
BONDING, EFFECTIVE STRESSES, AND STRENGTH 485 
Material 
Copyrighted Figure 12.25 Consolidation pressure, strength, and bond numbers for San Francisco Bay 
mud. 
Although it may be possible to explain these results 
in more than one way, the following interpretation ac-counts 
well for them. The energy 
F activates a mole 
of flow units. The movement of each flow unit may 
involve rupture of single bonds or the simultaneous 
rupture of several bonds. Shear of dilute montmoril-lonite– 
water pastes involves breaking single bonds 
(Ripple and Day, 1966). For viscous flow of water, the 
activation energy is approximately that for a single hy-drogen 
bond rupture per flow unit displacement, even 
though each water molecule may form simultaneously 
up to four hydrogen bonds with its neighbors. If the 
single-bond interpretation is also correct for soils, then 
consistency in Eq. (12.10) requires that shear force ƒ 
pertain to the force per bond. On this basis, parameter 
S indicates the number of single bonds per unit area. 
In the event activation of a flow unit requires simul-taneous 
rupture of n bonds, then S represents 1/nth of 
the total bonds in the system. 
That the activation energy for deformation of soil is 
well into the chemical reaction range (40 to 400 kJ/ 
mol) does not prove that bonding is of the primary 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
486 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
(2) it is the same for wet and dry soils, and (3) it is 
essentially the same for different adsorbed cations and Copyrighted Material 
Figure 12.26 Strength as a function of number of bonds for 
San Francisco Bay mud. 
Figure 12.28 Composite relationship between shear strength 
and number of interparticle bonds (from Matsui and Ito, 
1977). Reprinted with permission from The Japanese Society 
of SMFE. 
valence type because simultaneous rupture of several 
weaker bonds could yield values of the magnitude ob-served. 
On the other hand, the facts that (1) the acti-vation 
energy is much greater than for flow of water, 
Figure 12.27 Strength as a function of number of bonds for dry Antioch River sand. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
BONDING, EFFECTIVE STRESSES, AND STRENGTH 487 
pore fluids (Andersland and Douglas, 1970) suggest 
that bonding is through solid interparticle contacts. 
Physical evidence for the existence of solid-to-solid 
contact between clay particles has been obtained in the 
form of photomicrographs of particle surfaces that 
were scratched during shear (Matsui et al., 1977, 1980) 
and acoustic emissions (Koerner et al., 1977). 
Activation energy values of 125 to 190 kJ/mol are 
of the same order as those for solid-state diffusion of 
oxygen in silicate minerals. This supports the concept 
that creep movements of individual particles could re-sult 
Copyrighted Material 
from slow diffusion of oxygen ions in and around 
interparticle contacts. The important minerals in both 
sand and clay are silicates, and their surface layers 
consist of oxygen atoms held together by silicon at-oms. 
Water in some form is adsorbed onto these sur-faces. 
The water structure consists of oxygens held 
together by hydrogen. It is not too different from that 
of the silicate layer in minerals. Thus, a distinct bound-ary 
between particle surface and water may not be dis-cernable. 
Under these conditions, a more or less 
continuous solid structure containing water molecules 
that propagates through interparticle contacts can be 
visualized. 
An individual flow unit could be an atom, a group 
of atoms or molecules, or a particle. The preceding 
arguments are based on the interpretation that individ-ual 
atoms are the flow units. This is consistent with 
both the relative and actual values of S that have been 
determined for different soils. Furthermore, by using a 
formulation of the rate process equation that enabled 
calculation of the flow unit volume from creep test 
data, Andersland and Douglas (1970) obtained a value 
of about 1.7 A˚ 
3, which is of the same order as that of 
individual atoms. On the other hand, Keedwell (1984) 
defined flow units between quartz sand particles as 
consisting of six O2 ions and six Si4 ions and be-tween 
two montmorillonite clay particles as consisting 
of four H2O molecules. 
If particles were the flow units, not only would it be 
difficult to visualize their thermal vibrations, but then 
S would relate to the number of interparticle contacts. 
It is then difficult to conceive how simply drying a clay 
could give a 100-fold increase in the number of inter-particle 
contacts, as would have to be the case accord-ing 
to Fig. 12.27. A more plausible interpretation is 
that drying, while causing some increase in the number 
of interparticle contacts during shrinkage, causes 
mainly an increase in the number of bonds per contact 
because of increased effective stress. 
At any value of effective stress, the value of S is 
about the same for both sand and clay. The number of 
interparticle contacts should be vastly different; how-ever, 
for equal numbers of contacts per particle, the 
number per unit volume should vary inversely with the 
cube of particle size. Thus, the number of clay particles 
of 1-m particle size should be some nine orders of 
magnitude greater than for a sand of 1-mm average 
particle size. Each contact between sand particles 
would involve many bonds; in clay, the much greater 
number of contacts would mean fewer bonds per par-ticle. 
The contact area required to develop bonds in the 
numbers indicted in Figs. 12.23 to 12.27 is very small. 
For example, for a compressive strength of 3 kg/cm2 
( 300 kPa) there are 8  1010 bonds/cm2 of shear 
surface. Oxygen atoms on the surface of a silicate min-eral 
have a diameter of 0.28 nm. Allowing an area 0.30 
nm on a side for each oxygen gives 0.09 nm2, or 9  
1016 cm2, per bonded oxygen for a total area of 9  
1016  8  1010  7.2  105 cm2/cm2 of soil cross 
section. 
Hypothesis for Bonding, Effective Stress, and 
Strength 
Normal effective stresses and shear stresses can be 
transmitted only at interparticle contacts in most soils.5 
The predominant effects of the long-range physico-chemical 
forces of interaction are to control the initial 
soil fabric and to alter the forces transmitted at contact 
points from what they would be due to applied stresses 
alone. 
Interparticle contacts are effectively solid, and it is 
likely that both adsorbed water and cations in the con-tact 
zone participate in the structure. An interparticle 
contact may contain many bonds that may be strong, 
approaching the primary valence type. The number of 
bonds at any contact depends on the compressive force 
transmitted at the contact, and the Terzaghi–Bowden 
and Tabor adhesion theory of friction presented in Sec-tion 
11.4, can account for strength. The macroscopic 
strength is directly proportional to the number of 
bonds. 
For normally consolidated soils the number of bonds 
is directly proportional to the effective stress. As a re-sult 
of particle rearrangements and contacts formed 
during virgin compression, an overconsolidated soil at 
a given effective stress has a greater number of bonds 
and higher strength than a normally consolidated soil. 
This effect is more pronounced in clays than in sands 
because the larger and bulky sand grains tend to re- 
5 Pure sodium montmorillonite may be an exception since a part of 
the normal stress can be carried by physicochemical forces of inter-action. 
The true effective stress may be less than the apparent effec-tive 
stress by R  A as discussed in Chapter 7. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
488 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
cover their original shapes when unloaded, thus rup-turing 
most of the bonds in excess of those needed to 
resist the lower stress. The strength of the interparticle 
contacts can vary over a wide range, depending on the 
number of bonds per contact. 
The unique relationship between strength and num-ber 
of bonds for all soils, as indicted by Fig. 12.28, 
reflects the fact that the minerals comprising most soils 
are silicates, and they all have similar surface struc-tures. 
Copyrighted Material 
In the absence of chemical cementation, interparticle 
bonds may form in response to interparticle contact 
forces generated by either applied stresses, physico-chemical 
forces of interaction, or both. Any bonds ex-isting 
in the absence of applied effective stress, that is, 
when   0, are responsible for true cohesion. There 
should be no difference between friction and cohesion 
in terms of the shearing process. Complete failure in 
shear involves simultaneous rupture or slipping of all 
bonds along the shear plane. 
12.6 SHEARING RESISTANCE AS A RATE 
PROCESS 
Deformation at large strain can approach a steady-state 
condition where there is little further structural change 
with time (such as at critical state). In this case, Eq. 
(12.19) can be used to describe the shearing resistance 
as a function of strain rate and temperature. If the max-imum 
shear stress  is substituted for the deviator stress 
D, then 
1  3   (12.24) 
2 
and 
kT ˙ X exp 
F  exp  
  (12.25) h RT 2SkT 
Taking logarithms of both sides of Eq. (12.25) gives 
kT 
F  
ln  ˙  lnX   (12.26) h RT 2SkT 
By assuming X(kT/h) is a constant equal to B 
(Mitchell, 1964), Eq. (12.26) can be rearranged to give 
2S 2SkT   
˙   
F  ln(12.27) N  B 
From the relationships in Section 12.5, the following 
relationship between bonds per unit area and effective 
stress is suggested. 
S  a  b (12.28) ƒ 
where a and b are constants and ƒ is the effective 
normal stress on the shear plane. Thus, Eq. (12.27) 
becomes 
2a 
F 2akT  ˙ 2b 
F 2bkT  
˙  ln   ln  N  B N  B 
ƒ (12.29) 
Equation (12.29) is of the same form as the Cou-lomb 
equation for strength: 
  c   tan  (12.30) ƒ 
By analogy, 
2a 
F 2akT ˙ 
c  ln (12.31) 
N  B 
2b 
F 2bkT ˙ 
tan   ln (12.32) 
N  B 
These equations state that both cohesion and friction 
depend on the number of bonds times the bond 
strength, as reflected by the activation energy, and that 
the values of c and  should depend on the rate of 
deformation and the temperature. 
Strain Rate Effects 
All other factors being equal, the shearing resistance 
should increase linearly with the logarithm of the rate 
of strain. This is shown to be the case in Fig. 12.9, 
which contains data for 26 clays. Additional data for 
several clays are shown in Fig. 12.29, where shearing 
resistance as a function of the speed of vane rotation 
in a vane shear test is plotted. Analysis of the relation-ship 
between shearing stress and angular rate of vane 
rotation  shows that 
 /
 log  decreases with an 
increase in water content. This follows directly from 
Eq. (12.29) because 
d 2akT 2bkT 2kT 
    (a  b) d ln(˙ /B)   ƒ  ƒ 
(12.33) 
that is, d /d ln (˙ /B) is proportional to the number of 
bonds, which decreases with increasing water content. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
CREEP AND STRESS RELAXATION 489 
Material 
Copyrighted Figure 12.29 Effect of rate of shear on shearing resistance of remolded clays as determined 
by the laboratory vane apparatus (prepared from the data of Karlsson, 1963). 
This interpretation of the data in Figs. 12.9 and 12.29 
assumes that the effective stress was unaffected by 
changes in the strain rate, which may not necessarily 
be true in all cases. 
Effect of Temperature 
Assumptions of reasonable values for parameters show 
that the term (˙ /B) is less than one (Mitchell, 1964). 
Thus the quantity ln(˙ /B) in Eq. (12.29) is negative, 
and an increase in temperature should give a decrease 
in strength, all other factors being constant. That this 
is the case is demonstrated by Fig. 12.30, which shows 
deviator stress as a function of temperature for samples 
of San Francisco Bay mud compared under conditions 
of equal mean effective stress and structure. Other ex-amples 
of the influence of temperature on strength are 
shown in Figs. 11.6 and 11.133. 
12.7 CREEP AND STRESS RELAXATION 
Although the designation of a part of the strain versus 
time relationship as steady state or secondary creep 
may be convenient for some analysis purposes, a true 
steady state can exist only for conditions of constant 
structure and stress. Such a set of conditions is likely 
only for a fully destructured soil, and a fully destruc-tured 
state is likely to persist only during deformation 
at a constant rate, that is, at failure. This state is often 
called ‘‘steady state,’’ in which the soil is deforming 
continuously at constant volume under constant shear 
and confining stresses (Castro, 1975; Castro and 
Poulos, 1977). 
Otherwise, bond making and bond breaking occur 
at different rates as a result of different internal time-and 
strain-dependent phenomena, which might include 
thixotropic hardening, viscous flows of water and ad- 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
490 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
Copyrighted Material 
Figure 12.30 Influence of temperature on the shearing re-sistance 
of San Francisco Bay mud. Comparison is for sam-ples 
at equal mean effective stress and at the same structure. 
Figure 12.31 Variation of creep strain rate with deviator 
stress for drained creep of London clay (data from Bishop, 
1966). 
sorbed films, chemical, and biological transformations, 
and the like. Furthermore, distortions of the soil struc-ture 
and relative movements between particles cause 
changes in the ratio of tangential to normal forces at 
interparticle contacts that may be responsible for large 
changes in creep rate. Because of these time depend-encies 
some of the parameters in Eq. (12.19) may be 
time dependent. For example, Feda (1989) accounted 
for the time dependency of creep rate by taking 
changes in the number of structural bonds into account. 
Therefore, application of Eq. (12.19) for the determi-nation 
of the bonding and effective stress relationships 
discussed in Section 12.5 required comparison of creep 
rates under conditions of comparable time and struc-ture. 
The influence of creep stress magnitude on the creep 
rate at a given time after the application of the stress 
to identical samples of a soil was shown in Fig. 12.3. 
At low stresses the creep rates are small and of little 
practical importance. The curve shape is compatible 
with the hyperbolic sine function predicted by rate 
process theory, as given by Eq. (12.10). In the mid-range 
of stresses, a nearly linear relationship is found 
between logarithm of strain rate and stress, also as pre-dicted 
by Eq. (12.10) for the case where the argument 
of the hyperbolic sine is greater than 1. At stresses 
approaching the strength of the material, the strain rate 
becomes very large and signals the onset of failure. 
Other examples of the relationships between logarithm 
of strain rate and creep stress corresponding to differ-ent 
times after the application of the creep stress are 
given in Fig. 12.31 for drained tests on London clay 
and Fig. 12.32 for undrained tests on undisturbed San 
Francisco Bay mud. Only values for the midrange of 
stresses are shown in Figs. 12.31 and 12.32. 
Effect of Composition 
In general, the higher the clay content and the more 
active the clay, the more important are stress relaxation 
and creep, as illustrated by Figs. 4.22 and 4.23, where 
creep rates, approximated by steady-state values, are 
related to clay type, clay content, and plasticity. Time-dependent 
deformations are more important at high 
water contents than at low. Deviatoric creep and sec-ondary 
compression are greater in normally consoli-dated 
than overconsolidated soils. 
Although the magnitude of creep strains and strain 
rates may be small in sand or dry soil, the form of the 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
CREEP AND STRESS RELAXATION 491 
is a special case of volumetric creep. 
Deviatoric creep is often accompanied by volumetric 
creep. The ratio of volumetric to deviatoric creep fol- 
Copyrighted Material 
Figure 12.32 Variation of creep strain rate with deviator 
stress for undrained creep of normally consolidated San Fran-cisco 
Bay mud. 
1 
0.8 
0.6 
0.4 
0.2 
Drained Triaxial Test 
Confining Pressure σ3 = 414 kPa 
Deviator Stress from 344 to 377 kPa 
A B 
C 
D 
E 
2820 min 
1450 min 
90 min 
20 min 
2 min 
0 
0.5 1.0 1.5 2.0 2.5 3.0 3.5 
Deviator strain (%) 
1.0 
0.8 
0.6 
0.4 
0.2 
0 
(a) 
0 1 2 3 4 
Strain increment ratio dεv/dεs during creep 
(b) 
Volumetric Strain (%) 
Stress ratio q/p 
Figure 12.33 Dilatancy relationship obtained from drained 
creep tests on kaolinite: (a) development of volumetric and 
deviatoric strains with time and (b) effect of stress ratio on 
strain increment ratio d/ds (from Walker, 1969). 
behavior conforms with the patterns described and il-lustrated 
above. This is to be expected, as the basic 
creep mechanism is the same in all inorganic soils.6 
Water may ‘‘lubricate’’ the particles and possibly in-crease 
the creep rate even though the basic mechanism 
of creep is the same for dry and wet materials (Losert 
et al., 2000). Takei et al. (2001) showed that the de-velopment 
of creep strains due to time-dependent 
breakage of talc specimens increased more for satu-rated 
specimens than dry ones. However, a negligible 
effect of water on creep rate was reported by Ahn-Dan 
et al. (2001) who performed creep tests on unsaturated 
and saturated crushed gravel and by Leung et al. 
(1996) who performed one-dimensional compression 
6Volumetric creep and secondary compression of organic soils, peat, 
and municipal waste fills can develop also as a result of decompo-sition 
of organic matter. 
creep tests on sands. The conflicting evidence may be 
due to the presence or absence of impurities that may 
lubricate or cement the soil in the presence of water 
(Human, 1992; Bowman, 2003). 
Volume Change and Pore Pressures 
Due to the known coupling effects between shearing 
and volumetric plastic deformations in soils, an in-crease 
in either mean pressure or deviator stress can 
generate both types of deformations. Creep behavior is 
no exception. Time-dependent shear deformations are 
usually referred to as deviatoric creep or shear creep. 
Time-dependent deformations under constant stress re-ferred 
to as volumetric creep. Secondary compression 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
492 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
lows a plastic dilatancy rule. Walker (1969) inves-tigated 
the time-dependent change of these two com-ponents 
from incremental drained triaxial creep tests 
on normally consolidated kaolinite. The increase in 
shear strains with increase in volumetric strains at dif-ferent 
times is shown in Fig. 12.33a. At the beginning 
of the triaxial test, the deviator stress was instantane-ously 
increased from 344 to 377 kPa and kept constant. 
pressure increase with time for consolidated undrained 
creep tests on illite at several stress intensities is shown 
in Fig. 12.34. Figure 12.35 shows a slow decrease in 
Copyrighted Material 
After an immediate increase in shear strains at constant 
volume (AB in Fig. 12.33a), section BD corresponds 
to primary consolidation that is controlled by the dis-sipation 
of pore pressures. After point D, creep oc-curred, 
and the ratio of volumetric to deviatoric strains 
was independent of time. This ratio decreased with in-creasing 
stress ratio as shown in Fig. 12.33b. This ob-servation 
led to the time-dependent flow rule, which is 
similar to the dilatancy rule described in Section 11.20. 
Sand deforms with time in a similar manner. Under 
progressive deviatoric creep, the volumetric creep re-sponse 
is highly dependent on density, the stress level, 
and the stress path before creep. The rate of both vol-umetric 
and deviatoric creep increases with confining 
pressure, particularly after particle crushing becomes 
important at high stresses (Yamamuro and Lade, 1993). 
For dense sand under high deviator stress, dilative 
creep is observed (Murayama et al., 1984; Mejia et al., 
1988). The volumetric response of dense sand and 
gravel with time is a highly complex function of stress 
history, with some samples contracting or dilating 
(Lade and Liu, 1998; Ahn-Dan et al., 2001). Some 
dense sand samples contract initially but then dilate 
with time (Bowman and Soga, 2003). Further discus-sion 
of the creep behavior of sands in relation to me-chanical 
aging phenomena is given in Section 12.11. 
The fundamental process of creep strain develop-ment 
is therefore similar to that of time-independent 
plastic strains, and the same framework of soil plastic-ity 
can possibly be used. It can be argued whether 
it is necessary to separate the deformation into 
time-dependent and independent components. Rate-independent 
behavior can be considered as the limiting 
case of rate-dependent behavior at a very slow rate of 
loading. 
Volumetric-deviatoric creep coupling implies that 
rapid application of a stress or a strain invariably re-sults 
in rapid change of pore water pressures in a sat-urated 
soil under undrained conditions. For a constant 
total minor principal stress, the magnitude of the pore 
pressure change depends on the volume change ten-dencies 
of the soil when subjected to shear distortions. 
These tendencies are, in turn, controlled by the void 
ratio, structure, and effective stress, and can be quan-tified 
in terms of the pore pressure parameter A as dis-cussed 
in Chapters 8 and 10. An example showing pore 
Figure 12.34 Pore pressure development with time during undrained creep of illite. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
CREEP AND STRESS RELAXATION 493 
Material 
Figure 12.35 Normalized pore pressure vs. time relationships during creep of kaolinite. 
Copyrighted pore pressure during the sustained loading of kaolinite. 
Similar behavior was demonstrated in the measured 
stress paths of undrained creep test on San Francisco 
Bay mud (Arulanandan et al., 1971). As shown in Fig. 
12.36, the effective stress states shifted toward the fail-ure 
line. At higher stress levels, the specimens even-tually 
underwent creep rupture. However, soil strength 
in terms of effective stresses does not change unless 
there are chemical, biological, or mineralogical 
changes during the creep period. This is illustrated by 
the stress paths shown schematically in Fig. 12.37, 
where the pre- and postcreep strengths fall on the same 
failure envelope. 
Effects of Temperature 
An increase in temperature decreases effective stress, 
increases pore pressure, and weakens the soil structure. 
Creep rates ordinarily increase and the relaxation 
stresses corresponding to specific values of strain de-crease 
at higher temperature. These effects are illus-trated 
by the data shown in Figs. 12.38 and 12.39. 
Effects of Test Type, Stress System, and Stress Path 
Most measurements of time-dependent deformation 
and stress relaxation in soils have been done on sam-ples 
consolidated isotropically and tested in triaxial 
compression or by measurement of secondary com-pression 
in oedometer tests. However, most soils in 
nature have been subjected to an anisotropic stress his-tory, 
and deformation conditions conform more to 
plane strain than triaxial compression in many cases. 
Some investigations of these factors have been made. 
Although the general form of the stress–strain–time 
and stress–strain rate–time relationships are similar to 
those shown above for triaxial loading conditions, the 
actual values may differ considerably. 
For example, undisturbed Haney clay, a gray silty 
clay from British Columbia, with a sensitivity in the 
range of 6 to 10, was tested both in triaxial compres-sion 
and plane strain (Campanella and Vaid, 1974). 
Samples were normally consolidated both isotropically 
and under K0 conditions to the same vertical effective 
stress. Samples consolidated isotropically were tested 
in triaxial compression. Coefficient K0 consolidation 
was used for both K0 triaxial and plane strain tests. 
The results shown in Fig. 12.40 indicate that the pre-creep 
stress history had a significant effect on the 
deformations. The plane strain and K0 consolidated 
triaxial samples gave about the same creep behavior 
under the same deviatoric stress, which suggests that 
preventing strain in one horizontal direction and/or the 
intermediate principal stress were not factors of major 
importance for this soil under the test conditions used. 
Interaction Between Consolidation and Creep 
Experimental evidence suggests that creep occurs dur-ing 
primary consolidation (Leroueil et al., 1985; Imai 
and Tang, 1992). Following the initial large change 
following load application, the pore pressure may ei-ther 
dissipate, with accompanying volume change if 
drainage is allowed, or change slowly during creep or 
stress relaxation, if drainage is prevented. The devel-opment 
of complete effective stress and void ratio 
equilibrium may take a long time. One illustration of 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
494 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
Material 
0 10 20 30 40 50 60 70 80 
Mean Pressure p (kPa) 
Copyrighted 0 80 160 240 320 400 480 560 640 
Figure 12.36 Measured stress paths of undrained creep tests of San Francisco Bay mud. 
Initial confining pressure: (a) 49 kPa and (b) 392 kPa (from Arulanandan et al., 1971). 
50 
40 
30 
20 
10 
400 
320 
160 
80 
0 
Normal Undrained Triaxial 
Compression Test: Effective 
Stress Path 
Total Stress Path of 
Triaxial Compression Test 
1 
3 
Possible Critical 
State Line 
Effective Stress State 
After 1,000 min Creep 
After 20,000 min Creep 
Effective Stress Path 
of Undrained Creep 
0 
Normal Undrained Triaxial 
Compression Test: Effective 
Stress Path 
Total Stress Path of 
Triaxial Compression Test 
1 
3 
Possible Critical 
State Line 
Effective Stress State 
After 1,000 min Creep 
After 20,000 min Creep 
Effective Stress Path 
of Undrained Creep 
240 
(a) 
(b) 
Deviator Stress q (kPa) Deviator Stress q (kPa) 
Mean Pressure p (kPa) 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
CREEP AND STRESS RELAXATION 495 
Figure 12.37 Effects of undrained creep on the Material 
strength of normally consolidated clay. 
Copyrighted Figure 12.38 Creep curves for Osaka clay tested at different temperatures—undrained tri-axial 
compression (Murayama, 1969). 
this is given by Fig. 10.5, where it is shown that the 
relationship between void ratio and effective stress is 
dependent on the time for compression under any 
given stress. Another is given by Fig. 12.41, which 
shows pore pressures during undrained creep of San 
Francisco Bay mud. In each sample, consolidation un-der 
an effective confining pressure of 100 kPa was al-lowed 
for 1800 min prior to the cessation of drainage 
and the start of a creep test. The consolidation period 
was greater than that required for 100 percent primary 
consolidation. The curve marked 0 percent stress level 
refers to a specimen maintained undrained but not sub-ject 
to a deviator stress. This curve indicates that each 
of the other tests was influenced by a pore pressure 
that contained a contribution from the prior consoli-dation 
history. 
The magnitude and rate of pore pressure develop-ment 
if drainage is prevented following primary con-solidation 
depend on the time allowed for secondary 
compression prior to the prevention of further drain-age. 
This is illustrated by the data in Fig. 12.42, which 
show pore pressure as a function of time for samples 
that have undergone different amounts of secondary 
compression. 
In summary, creep deformation depends on the ef-fective 
stress path followed and any changes in stress 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
496 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
Figure 12.39 Influence of temperature on the Material 
initial and final stresses in stress relaxation 
tests on Osaka clay—undrained triaxial compression (Murayama, 1969). 
Copyrighted Figure 12.40 Creep curves for isotropically and K0-consolidated samples of undisturbed 
Haney clay tested in triaxial and plane strain compression (from Campanella and Vaid, 1974). 
Reproduced with permission from the National Research Council of Canada. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
RATE EFFECTS ON STRESS–STRAIN RELATIONSHIPS 497 
Figure 12.41 Pore pressure development during undrained creep of San Francisco Bay mud 
after consolidation at 100 kPa for 1800 min Material 
(from Holzer et al., 1973). Reproduced with 
permission from the National Research Council of Canada. 
Copyrighted Figure 12.42 Pore pressure development under undrained conditions following different 
periods of secondary compression (from Holzer et al., 1973). Reproduced with permission 
from the National Research Council of Canada. 
with time. Furthermore, time-dependent volumetric re-sponse 
is governed both by the rate of volumetric creep 
and by the rate of consolidation. The latter is a com-plex 
function of drainage conditions and material prop-erties, 
especially the permeability and compressibility. 
Because the effective stress path is controlled by the 
rate of loading and drainage conditions, the separation 
of consolidation and creep deformations can be diffi-cult 
in the early stage of time-dependent deformation 
as given by section BD in Fig. 12.33a. In some cases, 
a fully coupled analysis of soil–pore fluid interaction 
with an appropriate time-dependent constitutive model 
is necessary to reconcile the time-dependent deforma-tions 
observed in the field and laboratory. 
12.8 RATE EFFECTS ON STRESS–STRAIN 
RELATIONSHIPS 
An increase in strain rate during soil compression is 
manifested by increased stiffness, as was noted in Sec-tion 
12.3. In essence, the state of the soil jumps to the 
stress–strain curve that corresponds to the new strain 
rate. Commonly, this rate-dependent stress–strain 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
498 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
5% /h 
0.5% /h 
Copyrighted Material 
0.05% /h 
Belfast Clay 4 m 
σ1c = σv0 
16%/h 
1%/h 
0.25%/h 
Winnipeg Clay 11.5 m 
σ1c  σv0 
CAU Triaxial Compression Tests 
Relaxation Tests (R) 
0 
0.6 
0.5 
0.4 
0.3 
0.2 
0.1 
0 
4 8 12 16 20 
Axial Strain 
(σ1 – σ3)/2σ1c 
R 
R 
R 
Effective Stress σv (kPa) 
εv2 
. 
SP1 test 
SP2 test 
εv1 
. 
εv1 . = 2.70  10–6 s–1 
εv2 = 1.05  10–7 s–1 
εv2 . 
εv1 
. 
. 
. εv1 
εv1 
εv1 
. 
εv1 
. 
. 
. 
εv3 
. 
εv2 
. 
εv2 
ε.v3 = 1.34  10–5 s–1 
0 
5 
10 
15 
20 
25 
30 
0 50 100 150 200 250 
Strain εv (%) 
Figure 12.43 Rate-dependent stress–strain relations of 
clays: (a) undrained triaxial compression tests of Belfast 
and Winnipeg clay (Graham et al., 1983a) and (b) one-dimensional 
compression tests of Batiscan clay (Leroueil et 
al., 1985). 
curve, noted by S˘ 
uklje (1957), is the same as if the 
soil had been loaded from the beginning at the new 
strain rate. This phenomenon is often observed in 
clays. Examples are given in Fig. 12.43a for undrained 
triaxial compression tests of Belfast and Winnipeg 
clays (Graham et al., 1983a) and Fig. 12.43b for one-dimensional 
compression tests of Batiscan clay (Ler-oueil 
et al., 1985). 
Yield and Strength Envelopes of Clays 
The undrained shear strength and apparent precon-solidation 
pressure of soils decrease with decreasing 
strain rate or increasing duration of testing. Preconsol-idation 
pressures obtained from one-dimensional con-solidation 
tests and undrained shear strengths obtained 
from triaxial tests are just two points on a soil’s yield 
envelope in stress space. For a given metastable soil 
structure, the degree of rate dependency of preconsol-idation 
pressure is similar to that of undrained shear 
strength (Soga and Mitchell, 1996). If the apparent pre-consolidation 
pressure depends on the strain rate at 
which the soil is deforming, then the same analogy can 
be expanded to the assumption that the size of the en-tire 
yield envelope is also strain rate dependent (Tav-enas 
and Leroueil, 1977). Figure 12.44 shows a family 
of strength envelopes corresponding to constant strain 
rates7 obtained from drained and undrained creep tests 
on stiff plastic Mascouche clay from Quebec (Leroueil 
and Marques, 1996). 
The effective stress failure line of soil is uniquely 
defined regardless of the magnitude of the strain rate 
applied in undrained compression. Figure 12.45a 
shows the failure line of Haney clay (Vaid and Cam-panella, 
1977). The line represents the stress conditions 
at the maximum ratio of  /. The data were obtained 1 3 
by various undrained tests, and a unique failure line 
can be observed. Figure 12.45b shows the undrained 
stress paths and the critical state line of reconstituted 
mixtures of sand and clay with plasticity indices rang-ing 
from 10 to 30 (Nakase and Kamei, 1986). A unique 
critical state line can be observed although the rates of 
shearing are different. The change in undrained shear 
strength with strain rate results from a difference in 
generation of excess pore pressures. A decrease in 
strain rate leads to larger excess pore pressures at fail-ure 
due to creep deformation. 
7The strain rate is defined as ˙  	˙2  ˙2, where ˙ is the volu- vs v s v 
metric strain rate and ˙s is the deviator strain rate (Leroueil and Mar-ques, 
1996). Whether the use of this strain rate measure is appropriate 
or not remains to be investigated. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
RATE EFFECTS ON STRESS–STRAIN RELATIONSHIPS 499 
Material 
Copyrighted Figure 12.44 Influence of strain rate on the yield surface of Masouche clay (from Leroueil 
and Marques, 1996). 
Excess Pore Pressure Generation in Normally 
Consolidated Clays 
Excess pore pressure development depends primarily 
on the collapse of soil structure. Accordingly, strain is 
the primary factor controlling pore pressure generation. 
This is shown in Fig. 12.46 by the undrained stress– 
strain–pore pressure response of normally consolidated 
natural Olga clay (Lefebvre and LeBouef, 1987). The 
natural clay specimens were normally consolidated un-der 
consolidation pressures larger than the field over-burden 
pressure and then sheared at different strain 
rates. Although the deviator stress at any strain in-creases 
with increasing strain rate, the pore pressure 
versus strain curves are about the same at all strain 
rates. At a given deviator stress, the pore pressure gen-eration 
was larger at slower strain rates as a result of 
more creep under slow loading. This is consistent with 
the observation made in connection with undrained 
creep of clays as discussed in Section 12.7. Strain-driven 
pore pressure generation was also suggested by 
Larcerda and Houston (1973) who showed that pore 
pressure does not change significantly during triaxial 
stress relaxation tests in which the axial strain is kept 
constant. 
Overconsolidated Clays 
Rate dependency of undrained shear strength decreases 
with increasing overconsolidation, since there is no 
contraction or collapse tendency observed during creep 
of heavily overconsolidated clays. Sheahan et al. 
(1996) prepared reconstituted specimens of Boston 
blue clay at different overconsolidation ratios and 
sheared them at different strain rates in undrained con-ditions. 
Figure 12.47 shows that the undrained stress 
path and the strength were much more strain rate de-pendent 
for lightly overconsolidated clay (OCR  1 
and 2) than for more heavily overconsolidated clay 
(OCR  4 and 8). The results also show that the 
strength failure envelope is independent of strain rate 
as discussed earlier. 
The strain rate effects on stress–strain–pore pressure 
response of overconsolidated structured Olga clays are 
shown in Fig. 12.48 (Lefebvre and LeBouef, 1987). 
The natural samples were reconsolidated to the field 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
500 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
Const. Stress Creep 
Const. Rate of Strain Shear 
Const. Rate of Loading Shear 
Aged Samples 
Const. Load Creep 
Step Creep 
Thixotropic Hardened 
0.1 
0 
0 
0.1 
0.2 
Material 
Line 
Copyrighted State Critical-Line 
K0-Critical-State Line 
0 
0.2 
0.2 
0.3 
0.3 
0.4 
0.4 
0.4 
0.5 0.6 
0.6 
0.7 0.8 
M-15 
710-1: ε1 
710-2: ε.2 
710-3: ε3 
0.8 
(σ1 + σ3)/2σ1c 
(σ1 – σ3)/2σ1c 
(a) 
q/σvc 
(b) 
1.0 
0.8 
0.6 
0.4 
0.2 
0 
–0.2 
–0.4 
–0.6 
1.0 
p/σvc 
M-10 
. 
ε(%/min) 
. . 
710-1: ε1 
710-2: ε2 
710-3: ε3 
Critical-State Line 
K0-Line 
Critical-State Line 
0 0.2 0.4 0.6 0.8 1.0 
p/σvc 
q/σvc 
1.0 
0.8 
0.6 
0.4 
0.2 
0 
–0.2 
–0.4 
–0.6 
. 
ε(%/min) 
. 
. 
. 
M15 Soil (Plasticity Index = 15) M10 Soil (Plasticity Index = 10) 
Figure 12.45 Strain rate independent failure line: (a) Haney clay (from Vaid and Campa-nella, 
1977) and (b) reconstituted mixtures of sand and clay (from Nakase and Kamei, 1986). 
overburden pressure. The deformation is brittle, with 
strain softening indicating development of localized 
shear failure planes. Up to the peak stress, the response 
follows what has been described previously, that is the 
stress–strain response is rate dependent and the pore 
pressure generation is strain dependent but indepen-dent 
of rate. However, after the peak, the pore pressure 
generation becomes rate dependent. This is due to local 
drainage within the specimens as the deformation be-comes 
localized. As the time to failure increased, there 
is more opportunity for local drainage toward the di-lating 
shear band and the measured pore pressure may 
not represent the overall behavior of the specimens. 
The difference in softening due to swelling at the fail- 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
RATE EFFECTS ON STRESS–STRAIN RELATIONSHIPS 501 
Normally Consolidated Olga Clay 
Undrained Triaxial Compression Tests 
Initial Isotropic Confining Pressure =137 kPa 
Axial Strain Rate 
Material 
Copyrighted 1 2 3 4 5 6 7 8 
Axial Strain (%) 
Axial Strain Rate 
0.1 %/hr 
0.5 %/hr 
2.6 %/hr 
12.3 %/hr 
1 2 3 4 5 6 7 8 
Axial Strain (%) 
120 
100 
80 
60 
40 
20 
0 
140 
120 
100 
80 
60 
40 
20 
0 
0.1 %/hr 
0.5 %/hr 
2.6 %/hr 
12.3 %/hr 
Excess Pore Pressure Δu (kPa) Deviator Stress q (kPa) 
Figure 12.46 Stress–strain and pore pressure–strain curves 
for normally consolidated Olga clay (from Lefebvre and 
LeBouef, 1987). 
Overconsolidated Olga Clay 
Undrained Triaxial Compression Tests 
Initial Isotropic Confining Pressure = 17.6 kPa 
Axial Strain Rate 
1 2 3 4 5 6 7 8 
Axial Strain (%) 
Axial Strain Rate 
0.1 %/hr 
0.5 %/hr 
2.6 %/hr 
12.3 %/hr 
1 2 3 4 5 6 7 8 
Axial Strain (%) 
70 
60 
50 
40 
20 
10 
0 
10 
0 
0.1 %/hr 
0.5 %/hr 
2.5 %/hr 
12.3 %/hr 
30 
20 
Excess Pore Pressure Δu (kPa) Deviator Stress q (kPa) 
Figure 12.48 Stress–strain and pore pressure–strain curves 
for overconsolidated Olga clay (from Lefebvre and LeBouef, 
1987). 
OCR=1 
OCR=2 
OCR=4 
OCR=8 
OCR=2 
OCR=4 
Effective Stress Path for Axial 
Strain Rate = 0.50 %/hr 
0.2 0.4 0.6 0.8 
0.4 
0.3 
0.2 
0.1 
0.0 
-0.1 
Effective Stress State at Peak for 
Axial Strain Rate = 0.051 %/hr 
Axial Strain Rate = 0.50 %/hr 
Axial Strain Rate = 5.0 %/hr 
Axial Strain Rate = 49 %/hr 
Initial K0 Consolidation State 
(σa + σr)/2σvm 
(σa – σr)/2σvm 
OCR=1 
Large 
Rate 
OCR=8 Effect 
Negligible 
Rate Effect 
Figure 12.47 Rate dependency stress path and strength of 
overconsolidated Boston blue clay (from Sheahan et al., 
1996). 
ure plane results in apparent rate dependency at large 
strains. Similar observations were made by Atkinson 
and Richardson (1987) who examined local drainage 
effects by measuring the angles of intersection of shear 
bands with very different times of failure. 
Rate Effects on Sands 
Similar rate-dependent stress–strain behavior is ob-served 
in sands (Lade et al., 1997), but the effects are 
quite small in many cases (Tatsuoka et al., 1997; Di 
Benedetto et al., 2002). An example of time depend-ency 
observed for drained plane strain compression 
tests of Hostun sand is shown in Fig. 12.49 (Matsushita 
et al., 1999). The stress–strain curves for three differ-ent 
strain rates (1.25  101, 1.25  102, and 1.25  
103 %/min) are very similar, indicating very small 
rate effects when the specimens are sheared at a con-stant 
strain rate. On the other hand, the change from 
one rate to another temporarily increases or decreases 
the resistance to shear. The influence of acceleration 
rather than the rate is reflected by the significant creep 
deformation and stress relaxation of this rate-insensitive 
material as shown the figure. This is differ- 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
502 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
6.0 
5.5 
5.0 
4.5 
4.0 
3.5 
3.0 
Variation of Stress-strain curve by Constant 
Strain Rate Tests at Axial Strain Rates = 0.125, 
0.0125 and 0.00125 %/min. A Very Small Rate 
Effect Is Observed for Continuous Loading. 
CRS at 
0.125%/min 
Creep 
Material 
Copyrighted Creep 
Creep 
CRS at 
0.125%/min 
Stress Relaxation 
CRS at 0.00125%/min 
CRS at 0.125%/min 
Creep 
CRS at 0.125%/min 
CRS at 
0.00125%/min 
Accidental Pressure Drop Followed 
by Relaxation Stage 
0 1 2 3 4 5 6 7 8 
Stress Ratio σa /σr 
Shear Strain γ = εa – εr (%) 
Figure 12.49 Creep and stress relaxation of Hostun sand 
(from Matsushita et al., 1999). 
NSF-Clay 
Isotropically 
Consolidated 
Esec = Δq/εa, Eeq = (Δq)SA / (εa)SA 
Emax = 239 MPa p0 = 300 kPa 
300 
200 
100 
0 
Young's Modulus, Esec or Eeq (MPa) 
10-3 10-2 10-1 100 
Axial Strain, εa or 
Single Amplitude of Cyclic Axial Strain, (εa)SA (%) 
Figure 12.50 Clay stiffness degradation curves at three 
strain rates (from Shibuya et al., 1996). 
Coefficient of Strain Rate, (γ) 
Shear Strain, γ (%) 
Figure 12.51 Strain rate parameter G and strain level for 
several clays (from Lo Presti et al., 1996). 
ent from the observations made for clays as shown in 
Fig. 12.43 in which a unique stress–strain–strain rate 
relationship was observed. Hence, the modeling of 
stress–strain–rate behavior of sands appears to be 
more complicated than that of clays, and further in-vestigation 
is needed, as time-dependent behavior of 
sands can be of significance in geotechnical construc-tion 
as discussed further in Section 12.10. 
Stiffness at Small and Intermediate Strains 
Although the magnitude is small, the strain rate de-pendency 
of the stress–strain relationship is observed 
even at small strain levels for clays. The stiffness in-creases 
less than 6 percent per 10-fold increase in 
strain rate (Leroueil and Marques, 1996). The rate de-pendency 
on stiffness degradation curves measured by 
monotonic loading of a reconstituted clay is shown in 
Fig. 12.50 (Shibuya et al., 1996). At different strain 
levels, the increase in the secant shear modulus with 
shear strain rate 	˙ is often expressed by the following 
equation (Akai et al., 1975; Isenhower and Stokoe, 
1981; Lo Presti et al., 1996; Tatsuoka et al., 1997): 

G 
 (	)  (12.34) G 
 log ˙ 	  G(	, ˙ 	 ) ref 
where 
G is the increase in secant shear modulus with 
increase in log strain rate 
 log 	˙ , and G(	,	˙ ) is the ref 
secant shear modulus at strain 	 and reference strain 
rate 	˙ . The magnitude is large in clays, considerably ref 
less in silty and clayey sands, and small in clean sands 
(Lo Presti et al., 1996; Stokoe et al., 1999). The vari-ation 
of G with strain is shown in Fig. 12.51 for dif- 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
MODELING OF STRESS–STRAIN–TIME BEHAVIOR 503 
ferent plasticity clays (Lo Presti et al., 1996). The 
magnitude of rate dependency increases with strain 
level, especially for strain levels larger than 0.01 per-cent, 
which is within the preplastic region zone 3 de-scribed 
in Section 11.17. 
Rate Effects During Cyclic Loading 
The frequencies of cyclic loading to which a soil is 
subjected can vary widely. For example, the frequency 
of sea and ocean waves is in the range of 102 to 101 
Hz, earthquakes are in the range of 0.1 to a few hertz, 
and machine foundations are in the range of 10 to 100 
Hz. Similarly to monotonic loading, the effect of load-ing 
problems requiring the determination of de- 
Copyrighted Material 
frequency on shear modulus degradation is small 
in clean, coarse-grained soils (Bolton and Wilson, 
1989; Stokoe et al., 1995), but the effect becomes more 
significant in fine-grained soils (Stokoe et al., 1995; 
d’Onofrio et al., 1999; Matesic and Vucetic, 2003; 
Meng and Rix, 2004). An example of frequency effects 
on a shear modulus degradation curve for a clay ob-tained 
from cyclic loading is shown in Fig. 12.50 along 
with the monotonic data. Figure 12.52 shows the effect 
of frequency on shear modulus of several soils at very 
small shear strains (less than 103 percent) measured 
by torsional shear and resonant column apparatuses 
(Meng and Rix, 2004). The effect is 10 percent in-crease 
per log cycle at most. 
At a given frequency of cyclic loading, the strain 
rate applied to a soil increases with applied shear strain 
as shown by the equation below: 
˙ 	  4ƒ	 (12.35) c 
250 
200 
150 
100 
50 
where ƒ is the frequency and 	c is the cyclic shear 
strain amplitude. Using Eq. (12.35), Matesic and Vu-cetic 
(2003) report values of G of 2 to 11 percent for 
clays and 0.2 to 6 percent for sands as the strain rate 
increases 10-fold. The values of G in general de-creased 
when the applied cyclic shear strain increased 
from 5  104 percent to 1  102 percent. It should 
be noted that the strain range examined is within the 
non-linear elastic range (zone 1 to zone 2 in Section 
11.17). The monotonic loading data presented in Fig. 
12.51 show that the rate effect becomes more pro-nounced 
at larger strain, that is, as plastic deformations 
become more significant. Hence, it is possible that the 
fundamental mechanisms of rate dependency are dif-ferent 
at small elastic strain levels than at larger plastic 
strains. 
Small strain damping shows more complex fre-quency 
dependency, as shown in Fig. 12.53 (Shibuya 
et al., 1995; Meng and Rix, 2004). At a frequency of 
more than 10 Hz, the damping ratio increases with 
increased frequency, possibly due to pore fluid viscos-ity 
effects. As the applied frequency decreases, the 
damping ratio decreases. However, at a frequency less 
than 0.1 Hz, the damping ratio starts to increase with 
decreasing frequency. This may result from creep of 
the soil (Shibuya et al., 1995). 
12.9 MODELING OF STRESS–STRAIN–TIME 
BEHAVIOR 
Constitutive models are needed for the solution of geo-technical 
Sandy Elastic Silt 
0 
10-2 10-1 100 101 102 
Frequency (Hz) 
Sandy Silty Clay 
Kaolin 
Vallencca Clay 
Kaolin 
Sandy Lean Clay 
Kaolin Subgrade 
Fat Clay 
Shear Modulus G (MPa) 
Figure 12.52 Rate dependency of cyclic small strain stiffness of a sandy elastic silt (from 
Meng and Rix, 2004). 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
504 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 
Clayey Subgrades 
Sandy Lean Clay 
10-2 10-1 100 101 102 
Frequency Material 
(Hz) 
Copyrighted 6 
5 
4 
3 
2 
1 
Sandy Silty Clay 
Sandy Elastic Silt 
Kaolin 
Vallencca Clay Fat Clay 
Pisa Clay 
Augusta Clay 
Damping (%) 
Figure 12.53 Effect of strain rate of damping ratio of soils (from Shibuya et al., 1995 and 
Meng and Rix, 2004). 
formations, displacements, and strength and stability 
changes that occur over time periods of different 
lengths. Various approaches have been used, including 
empirical curve fitting, extensions of rate process 
theory, rheological models, and advanced theories 
of viscoelasticity and viscoplasticity. Owing to the 
complexity of stress states, the many factors that influ-ence 
the creep and stress relaxation properties of a soil, 
and the difficulty of accounting for concurrent volu-metric 
and deviatoric deformations in systems that are 
many times undergoing consolidation as well as sec-ondary 
compression or creep, it is not surprising that 
development of general models that can be readily im-plemented 
in engineering practice is a challenging un-dertaking. 
Nonetheless, some progress has been made in estab-lishing 
functional forms and relationships that can be 
applied for simple analyses and comparisons, and one 
of these is developed in this section. A complete re-view 
and development of all recent theories and pro-posed 
relationships for creep and stress relaxation is 
beyond the scope of this book. Comprehensive reviews 
of many models for representation of the time-dependent 
plastic response of soils are given in Adachi 
et al. (1996). 
General Stress–Strain–Time Function 
Strain Rate Relationships between axial strain rate 
˙ and time t of the type shown in Figs. 12.4 and 12.5 
can be expressed by 
 ˙t 
ln  m ln  (12.36) ˙(t ,D) t 1 1 
or 
t 
ln ˙  ln ˙(t ,D) m ln  (12.37) 1  t1 
where ˙(t ,D) is the axial strain rate at unit time and 1 
is a function of stress intensity D, m is the absolute 
value of the slope of the straight line on the log strain 
rate versus log time plot, and t1 is a reference time, for 
example, 1 min. Values of m generally fall in the range 
of 0.7 to 1.3 for triaxial creep tests; lower values are 
reported for undrained conditions than for drained con-ditions. 
For the development shown here, the stress 
intensity D is taken as the deviator stress (1  3). A 
shear stress or stress level could also be used. 
The same data plotted in the form of Figs. 12.3, 
12.31, and 12.32 can be expressed by 
 
˙ln
   D (12.38) ˙(t,D ) 0 
or 
ln ˙  ln ˙(t,D )  D (12.39) 0 
in which ˙(t,D ) is a fictitious value of strain rate at 0 
D  0, a function of time after start of creep, and  
is the slope of the linear part of the log strain rate 
versus stress plot. From Eqs. (12.37) and (12.39) 
t 
ln ˙(t ,D)  m ln  1  ln ˙(t,D ) 0  D (12.40) t 1 
For D  0, 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
63027 12
63027 12
63027 12
63027 12
63027 12
63027 12
63027 12
63027 12
63027 12
63027 12
63027 12
63027 12
63027 12
63027 12
63027 12
63027 12
63027 12
63027 12

More Related Content

What's hot

Effect of differential settlement on frame forces a parametric study
Effect of differential settlement on frame forces   a parametric studyEffect of differential settlement on frame forces   a parametric study
Effect of differential settlement on frame forces a parametric studyeSAT Journals
 
Settlement of shallow foundations
Settlement of shallow foundationsSettlement of shallow foundations
Settlement of shallow foundationsAmirmasoud Taghavi
 
Goetech. engg. Ch# 03 settlement analysis signed
Goetech. engg. Ch# 03 settlement analysis signedGoetech. engg. Ch# 03 settlement analysis signed
Goetech. engg. Ch# 03 settlement analysis signedIrfan Malik
 
Consolidation mechanism
Consolidation mechanismConsolidation mechanism
Consolidation mechanismParth Joshi
 
8 compressibility and consolidation
8 compressibility and consolidation8 compressibility and consolidation
8 compressibility and consolidationSaurabh Kumar
 
Deformation activity
Deformation activityDeformation activity
Deformation activityDNHSGRADE8
 
Fold development
Fold developmentFold development
Fold developmentZeomSlim
 
rock mechanics
rock mechanics rock mechanics
rock mechanics 1971995
 
Physico-Mechanical properties of rock materials
Physico-Mechanical properties of rock materialsPhysico-Mechanical properties of rock materials
Physico-Mechanical properties of rock materialsUlimella Siva Sankar
 
Consolidation nicee ppt
Consolidation nicee pptConsolidation nicee ppt
Consolidation nicee pptfutsa
 
case study on terzaghi’s theory
case study on terzaghi’s theorycase study on terzaghi’s theory
case study on terzaghi’s theoryAbhishek Mangukiya
 
7 slopes and slope stability
7 slopes and slope stability7 slopes and slope stability
7 slopes and slope stabilityRaghav Gadgil
 

What's hot (20)

pre consolidation pressure
pre consolidation pressurepre consolidation pressure
pre consolidation pressure
 
Settlement of soil/foundation
Settlement of soil/foundationSettlement of soil/foundation
Settlement of soil/foundation
 
Effect of differential settlement on frame forces a parametric study
Effect of differential settlement on frame forces   a parametric studyEffect of differential settlement on frame forces   a parametric study
Effect of differential settlement on frame forces a parametric study
 
Settlement of shallow foundations
Settlement of shallow foundationsSettlement of shallow foundations
Settlement of shallow foundations
 
Goetech. engg. Ch# 03 settlement analysis signed
Goetech. engg. Ch# 03 settlement analysis signedGoetech. engg. Ch# 03 settlement analysis signed
Goetech. engg. Ch# 03 settlement analysis signed
 
63027 10
63027 1063027 10
63027 10
 
Consolidation mechanism
Consolidation mechanismConsolidation mechanism
Consolidation mechanism
 
Ch9 settlement (1)
Ch9 settlement (1)Ch9 settlement (1)
Ch9 settlement (1)
 
8 compressibility and consolidation
8 compressibility and consolidation8 compressibility and consolidation
8 compressibility and consolidation
 
liquifaction
liquifactionliquifaction
liquifaction
 
Deformation activity
Deformation activityDeformation activity
Deformation activity
 
Fold development
Fold developmentFold development
Fold development
 
Consolidation
ConsolidationConsolidation
Consolidation
 
U01226132138
U01226132138U01226132138
U01226132138
 
Chapter 14
Chapter 14Chapter 14
Chapter 14
 
rock mechanics
rock mechanics rock mechanics
rock mechanics
 
Physico-Mechanical properties of rock materials
Physico-Mechanical properties of rock materialsPhysico-Mechanical properties of rock materials
Physico-Mechanical properties of rock materials
 
Consolidation nicee ppt
Consolidation nicee pptConsolidation nicee ppt
Consolidation nicee ppt
 
case study on terzaghi’s theory
case study on terzaghi’s theorycase study on terzaghi’s theory
case study on terzaghi’s theory
 
7 slopes and slope stability
7 slopes and slope stability7 slopes and slope stability
7 slopes and slope stability
 

Viewers also liked

40047683 aashto-geometric-design-for-highway-and-streets-2001
40047683 aashto-geometric-design-for-highway-and-streets-200140047683 aashto-geometric-design-for-highway-and-streets-2001
40047683 aashto-geometric-design-for-highway-and-streets-2001Defrim Kabashi
 
Question and Answers on Terzaghi’s Bearing Capacity Theory (usefulsearch.org)...
Question and Answers on Terzaghi’s Bearing Capacity Theory (usefulsearch.org)...Question and Answers on Terzaghi’s Bearing Capacity Theory (usefulsearch.org)...
Question and Answers on Terzaghi’s Bearing Capacity Theory (usefulsearch.org)...Make Mannan
 
Du pont guía de diseño
Du pont   guía de diseño Du pont   guía de diseño
Du pont guía de diseño vegus
 
Monografia puentes aashto lrfd 2007. ing. salvador y pedro
Monografia puentes aashto lrfd 2007. ing. salvador y pedroMonografia puentes aashto lrfd 2007. ing. salvador y pedro
Monografia puentes aashto lrfd 2007. ing. salvador y pedroEdwin Cruz
 
Fundamentals of engineering geology
Fundamentals of engineering geologyFundamentals of engineering geology
Fundamentals of engineering geologyWan Zuhairi Yaacob
 

Viewers also liked (8)

63027 08
63027 0863027 08
63027 08
 
63027 11b
63027 11b63027 11b
63027 11b
 
Aashto guide specifications for lrfd seismic
Aashto guide specifications for lrfd seismicAashto guide specifications for lrfd seismic
Aashto guide specifications for lrfd seismic
 
40047683 aashto-geometric-design-for-highway-and-streets-2001
40047683 aashto-geometric-design-for-highway-and-streets-200140047683 aashto-geometric-design-for-highway-and-streets-2001
40047683 aashto-geometric-design-for-highway-and-streets-2001
 
Question and Answers on Terzaghi’s Bearing Capacity Theory (usefulsearch.org)...
Question and Answers on Terzaghi’s Bearing Capacity Theory (usefulsearch.org)...Question and Answers on Terzaghi’s Bearing Capacity Theory (usefulsearch.org)...
Question and Answers on Terzaghi’s Bearing Capacity Theory (usefulsearch.org)...
 
Du pont guía de diseño
Du pont   guía de diseño Du pont   guía de diseño
Du pont guía de diseño
 
Monografia puentes aashto lrfd 2007. ing. salvador y pedro
Monografia puentes aashto lrfd 2007. ing. salvador y pedroMonografia puentes aashto lrfd 2007. ing. salvador y pedro
Monografia puentes aashto lrfd 2007. ing. salvador y pedro
 
Fundamentals of engineering geology
Fundamentals of engineering geologyFundamentals of engineering geology
Fundamentals of engineering geology
 

Similar to 63027 12

Pyke static and cyclic liquefaction - annotated
Pyke   static and cyclic liquefaction - annotatedPyke   static and cyclic liquefaction - annotated
Pyke static and cyclic liquefaction - annotatedRobert Pyke
 
Soil liquefaction seminar report
Soil liquefaction seminar reportSoil liquefaction seminar report
Soil liquefaction seminar reportshaik ahmed jawad
 
liquefaction of soil
liquefaction of soilliquefaction of soil
liquefaction of soilHUSSAINGHAZI1
 
Multistage hydraulic fracturing;fluid flow and geomechanics
Multistage hydraulic fracturing;fluid flow and geomechanicsMultistage hydraulic fracturing;fluid flow and geomechanics
Multistage hydraulic fracturing;fluid flow and geomechanicsHarshal Pende
 
Understanding Permeability of Hydraulic Fracture Networks A Sandbox Analog Mo...
Understanding Permeability of Hydraulic Fracture Networks A Sandbox Analog Mo...Understanding Permeability of Hydraulic Fracture Networks A Sandbox Analog Mo...
Understanding Permeability of Hydraulic Fracture Networks A Sandbox Analog Mo...Renee Heldman
 
Assessment and risk reduction measurement of liquefaction of soil
Assessment and risk reduction measurement of liquefaction of soilAssessment and risk reduction measurement of liquefaction of soil
Assessment and risk reduction measurement of liquefaction of soilvivatechijri
 
waterinfluencesdifferentbehavioursofsoil.pptx
waterinfluencesdifferentbehavioursofsoil.pptxwaterinfluencesdifferentbehavioursofsoil.pptx
waterinfluencesdifferentbehavioursofsoil.pptxINDRANIL BANERJEE
 
Sequence Stratigraphy.pptx
Sequence Stratigraphy.pptxSequence Stratigraphy.pptx
Sequence Stratigraphy.pptxSaadTaman
 
Liquefaction by alisha kaplan
Liquefaction by alisha kaplanLiquefaction by alisha kaplan
Liquefaction by alisha kaplanmk.amiri
 
UNIT-III Consolidation.ppt
UNIT-III Consolidation.pptUNIT-III Consolidation.ppt
UNIT-III Consolidation.pptmythili spd
 
Slope stability question answers
Slope stability question answersSlope stability question answers
Slope stability question answersJyoti Khatiwada
 
Forsyth what is_the_lithosphere
Forsyth what is_the_lithosphereForsyth what is_the_lithosphere
Forsyth what is_the_lithospherequinterofelipe010
 
Consolidation
ConsolidationConsolidation
ConsolidationAmr Assr
 

Similar to 63027 12 (20)

Pyke static and cyclic liquefaction - annotated
Pyke   static and cyclic liquefaction - annotatedPyke   static and cyclic liquefaction - annotated
Pyke static and cyclic liquefaction - annotated
 
Soil liquefaction seminar report
Soil liquefaction seminar reportSoil liquefaction seminar report
Soil liquefaction seminar report
 
liquefaction of soil
liquefaction of soilliquefaction of soil
liquefaction of soil
 
63027 11a
63027 11a63027 11a
63027 11a
 
Multistage hydraulic fracturing;fluid flow and geomechanics
Multistage hydraulic fracturing;fluid flow and geomechanicsMultistage hydraulic fracturing;fluid flow and geomechanics
Multistage hydraulic fracturing;fluid flow and geomechanics
 
Understanding Permeability of Hydraulic Fracture Networks A Sandbox Analog Mo...
Understanding Permeability of Hydraulic Fracture Networks A Sandbox Analog Mo...Understanding Permeability of Hydraulic Fracture Networks A Sandbox Analog Mo...
Understanding Permeability of Hydraulic Fracture Networks A Sandbox Analog Mo...
 
Liquefaction of soil
Liquefaction of soilLiquefaction of soil
Liquefaction of soil
 
Assessment and risk reduction measurement of liquefaction of soil
Assessment and risk reduction measurement of liquefaction of soilAssessment and risk reduction measurement of liquefaction of soil
Assessment and risk reduction measurement of liquefaction of soil
 
waterinfluencesdifferentbehavioursofsoil.pptx
waterinfluencesdifferentbehavioursofsoil.pptxwaterinfluencesdifferentbehavioursofsoil.pptx
waterinfluencesdifferentbehavioursofsoil.pptx
 
UNIT-III.ppt
UNIT-III.pptUNIT-III.ppt
UNIT-III.ppt
 
Sequence Stratigraphy.pptx
Sequence Stratigraphy.pptxSequence Stratigraphy.pptx
Sequence Stratigraphy.pptx
 
Liquefaction by alisha kaplan
Liquefaction by alisha kaplanLiquefaction by alisha kaplan
Liquefaction by alisha kaplan
 
UNIT-III Consolidation.ppt
UNIT-III Consolidation.pptUNIT-III Consolidation.ppt
UNIT-III Consolidation.ppt
 
Sullivan
SullivanSullivan
Sullivan
 
Shear Strength Behavior for Unsoaked and Soaked Multistage Triaxial Drained T...
Shear Strength Behavior for Unsoaked and Soaked Multistage Triaxial Drained T...Shear Strength Behavior for Unsoaked and Soaked Multistage Triaxial Drained T...
Shear Strength Behavior for Unsoaked and Soaked Multistage Triaxial Drained T...
 
Slope stability question answers
Slope stability question answersSlope stability question answers
Slope stability question answers
 
Forsyth what is_the_lithosphere
Forsyth what is_the_lithosphereForsyth what is_the_lithosphere
Forsyth what is_the_lithosphere
 
Consolidation
ConsolidationConsolidation
Consolidation
 
MP.pptx
MP.pptxMP.pptx
MP.pptx
 
Research thesis
Research thesisResearch thesis
Research thesis
 

More from Vcoi Vit

2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri31
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri312eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri31
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri31Vcoi Vit
 
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri14
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri142eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri14
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri14Vcoi Vit
 
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri13
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri132eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri13
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri13Vcoi Vit
 
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri05
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri052eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri05
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri05Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san46
Giao trinh nuoi_trong_thuy_san46Giao trinh nuoi_trong_thuy_san46
Giao trinh nuoi_trong_thuy_san46Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san28
Giao trinh nuoi_trong_thuy_san28Giao trinh nuoi_trong_thuy_san28
Giao trinh nuoi_trong_thuy_san28Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san23
Giao trinh nuoi_trong_thuy_san23Giao trinh nuoi_trong_thuy_san23
Giao trinh nuoi_trong_thuy_san23Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san22
Giao trinh nuoi_trong_thuy_san22Giao trinh nuoi_trong_thuy_san22
Giao trinh nuoi_trong_thuy_san22Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san21
Giao trinh nuoi_trong_thuy_san21Giao trinh nuoi_trong_thuy_san21
Giao trinh nuoi_trong_thuy_san21Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san16
Giao trinh nuoi_trong_thuy_san16Giao trinh nuoi_trong_thuy_san16
Giao trinh nuoi_trong_thuy_san16Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san05
Giao trinh nuoi_trong_thuy_san05Giao trinh nuoi_trong_thuy_san05
Giao trinh nuoi_trong_thuy_san05Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san02
Giao trinh nuoi_trong_thuy_san02Giao trinh nuoi_trong_thuy_san02
Giao trinh nuoi_trong_thuy_san02Vcoi Vit
 
Giao trinh tong hop sv50
Giao trinh tong hop sv50Giao trinh tong hop sv50
Giao trinh tong hop sv50Vcoi Vit
 
Giao trinh tong hop sv49
Giao trinh tong hop sv49Giao trinh tong hop sv49
Giao trinh tong hop sv49Vcoi Vit
 
Giao trinh tong hop sv48
Giao trinh tong hop sv48Giao trinh tong hop sv48
Giao trinh tong hop sv48Vcoi Vit
 
Giao trinh tong hop sv47
Giao trinh tong hop sv47Giao trinh tong hop sv47
Giao trinh tong hop sv47Vcoi Vit
 
Giao trinh tong hop sv46
Giao trinh tong hop sv46Giao trinh tong hop sv46
Giao trinh tong hop sv46Vcoi Vit
 
Giao trinh tong hop sv45
Giao trinh tong hop sv45Giao trinh tong hop sv45
Giao trinh tong hop sv45Vcoi Vit
 
Giao trinh tong hop sv43
Giao trinh tong hop sv43Giao trinh tong hop sv43
Giao trinh tong hop sv43Vcoi Vit
 
Giao trinh tong hop sv41
Giao trinh tong hop sv41Giao trinh tong hop sv41
Giao trinh tong hop sv41Vcoi Vit
 

More from Vcoi Vit (20)

2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri31
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri312eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri31
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri31
 
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri14
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri142eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri14
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri14
 
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri13
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri132eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri13
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri13
 
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri05
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri052eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri05
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri05
 
Giao trinh nuoi_trong_thuy_san46
Giao trinh nuoi_trong_thuy_san46Giao trinh nuoi_trong_thuy_san46
Giao trinh nuoi_trong_thuy_san46
 
Giao trinh nuoi_trong_thuy_san28
Giao trinh nuoi_trong_thuy_san28Giao trinh nuoi_trong_thuy_san28
Giao trinh nuoi_trong_thuy_san28
 
Giao trinh nuoi_trong_thuy_san23
Giao trinh nuoi_trong_thuy_san23Giao trinh nuoi_trong_thuy_san23
Giao trinh nuoi_trong_thuy_san23
 
Giao trinh nuoi_trong_thuy_san22
Giao trinh nuoi_trong_thuy_san22Giao trinh nuoi_trong_thuy_san22
Giao trinh nuoi_trong_thuy_san22
 
Giao trinh nuoi_trong_thuy_san21
Giao trinh nuoi_trong_thuy_san21Giao trinh nuoi_trong_thuy_san21
Giao trinh nuoi_trong_thuy_san21
 
Giao trinh nuoi_trong_thuy_san16
Giao trinh nuoi_trong_thuy_san16Giao trinh nuoi_trong_thuy_san16
Giao trinh nuoi_trong_thuy_san16
 
Giao trinh nuoi_trong_thuy_san05
Giao trinh nuoi_trong_thuy_san05Giao trinh nuoi_trong_thuy_san05
Giao trinh nuoi_trong_thuy_san05
 
Giao trinh nuoi_trong_thuy_san02
Giao trinh nuoi_trong_thuy_san02Giao trinh nuoi_trong_thuy_san02
Giao trinh nuoi_trong_thuy_san02
 
Giao trinh tong hop sv50
Giao trinh tong hop sv50Giao trinh tong hop sv50
Giao trinh tong hop sv50
 
Giao trinh tong hop sv49
Giao trinh tong hop sv49Giao trinh tong hop sv49
Giao trinh tong hop sv49
 
Giao trinh tong hop sv48
Giao trinh tong hop sv48Giao trinh tong hop sv48
Giao trinh tong hop sv48
 
Giao trinh tong hop sv47
Giao trinh tong hop sv47Giao trinh tong hop sv47
Giao trinh tong hop sv47
 
Giao trinh tong hop sv46
Giao trinh tong hop sv46Giao trinh tong hop sv46
Giao trinh tong hop sv46
 
Giao trinh tong hop sv45
Giao trinh tong hop sv45Giao trinh tong hop sv45
Giao trinh tong hop sv45
 
Giao trinh tong hop sv43
Giao trinh tong hop sv43Giao trinh tong hop sv43
Giao trinh tong hop sv43
 
Giao trinh tong hop sv41
Giao trinh tong hop sv41Giao trinh tong hop sv41
Giao trinh tong hop sv41
 

63027 12

  • 1. 465 CHAPTER 12 Copyrighted Material Time Effects on Strength and Deformation 12.1 INTRODUCTION Virtually every soil ‘‘lives’’ in that all of its properties undergo changes with time–some insignificant, but others very important. Time-dependent chemical, geomicrobiological, and mechanical processes may result in compositional and structural changes that lead to softening, stiffening, strength loss, strength gain, or altered conductivity properties. The need to predict what the properties and behavior will be months to hundreds or thousands of years from now based on what we know today is a major challenge in geoengineering. Some time-dependent changes and their effects as they relate to soil formation, composi-tion, weathering, postdepositional changes in sedi-ments, the evolution of soil structure, and the like are considered in earlier chapters of this book. Emphasis in this chapter is on how time under stress changes the structural, deformation, and strength properties of soils, what can be learned from knowledge of these changes, and their quantification for predictive pur-poses. When soil is subjected to a constant load, it deforms over time, and this is usually called creep. The inverse phenomenon, usually termed stress relaxation, is a drop in stress over time after a soil is subjected to a particular constant strain level. Creep and relaxation are two consequences of the same phenomenon, that is, time-dependent changes in structure. The rate and magnitude of these time-dependent deformations are determined by these changes. Time-dependent deformations and stress relaxation are important in geotechnical problems wherein long-term behavior is of interest. These include long-term settlement of structures on compressible ground, de-formations of earth structures, movements of natural and excavated slopes, squeezing of soft ground around tunnels, and time- and stress-dependent changes in soil properties. The time-dependent deformation response of a soil may assume a variety of forms owing to the complex interplays among soil structure, stress history, drainage conditions, and changes in temperature, pres-sure, and biochemical environment with time. Time-dependent deformations and stress relaxations usually follow logical and often predictable patterns, at least for simple stress and deformation states such as uni-axial and triaxial compression, and they are described in this chapter. Incorporation of the observed behav-ior into simple constitutive models for analytical de-scription of time-dependent deformations and stress changes is also considered. Time-dependent deformation and stress phenomena in soils are important not only because of the imme-diate direct application of the results to analyses of practical problems, but also because the results can be used to obtain fundamental information about soil structure, interparticle bonding, and the mechanisms controlling the strength and deformation behavior. Both microscale and macroscale phenomena are dis-cussed because understanding of microscale processes can provide a rational basis for prediction of macro-scale responses. Copyright © 2005 John Wiley & Sons Retrieved from: www.knovel.com
  • 2. 466 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 12.2 GENERAL CHARACTERISTICS 1. As noted in the previous section soils exhibit both creep1 and stress relaxation (Fig. 12.1). Creep is the development of time-dependent shear and/or volumetric strains that proceed at a rate controlled by the viscouslike resistance of soil structure. Stress relaxation is a time-dependent decrease in stress at constant defor-mation. The relationship between creep strain Material Copyrighted and the logarithm of time may be linear, con-cave upward, or concave downward as shown by the examples in Fig. 12.2. 2. The magnitude of these effects increases with increasing plasticity, activity, and water content Figure 12.1 Creep and stress relaxation: (a) Creep under constant stress and (b) stress relaxation under constant strain. 1The term creep is used herein to refer to time-dependent shear strains and/ or volumetric strains that develop at a rate controlled by the viscous resistance of the soil structure. Secondary compression refers to the special case of volumetric strain that follows primary consolidation. The rate of secondary compression is controlled by the viscous resistance of the soil structure, whereas, the rate of pri-mary consolidation is controlled by hydrodynamic lag, that is, how fast water can escape from the soil. of the soil. The most active clays usually exhibit the greatest time-dependent responses (i.e., smectite illite kaolinite). This is because the smaller the particle size, the greater is the specific surface, and the greater the water ad-sorption. Thus, under a given consolidation stress or deviatoric stress, the more active and plastic clays (smectites) will be at higher water content and lower density than the inactive clays (kaolinites). Normally consolidated soils exhibit larger magnitude of creep than overconsolidated soils. However, the basic form of behavior is essentially the same for all soils, that is, undis-turbed and remolded clay, wet clay, dry clay, normally and overconsolidated soil, and wet and dry sand. 3. An increase in deviatoric stress level results in an increased rate of creep as shown in Fig. 12.1. Some soils may fail under a sustained creep stress significantly less (as little as 50 percent) than the peak stress measured in a shear test, wherein a sample is loaded to failure in a few minutes or hours. This is termed creep rupture, and an early illustration of its importance was the development of slope failures in the Cucar-acha clay shale, which began some years after the excavation of the Panama Canal (Casa-grande and Wilson, 1951). 4. The creep response shown by the upper curve in Fig. 12.1 is often divided into three stages. Following application of a stress, there is first a period of transient creep during which the strain rate decreases with time, followed by creep at nearly a constant rate for some period. For ma-terials susceptible to creep rupture, the creep rate then accelerates leading to failure. These three stages are termed primary, secondary, and tertiary creep. 5. An example of strain rates as a function of stress for undrained creep of remolded illite is shown in Fig. 12.3. At low deviator stress, creep rates are very small and of little practical importance. The curve shapes for deviator stresses up to about 1.0 kg/cm2 are compatible with the pre-dictions of rate process theory, discussed in Sec-tion 12.4. At deviator stress approaching the strength of the material, the strain rates become very large and signal the onset of failure. 6. A characteristic relationship between strain rate and time exists for most soils, as shown, for example, in Fig. 12.4 for drained triaxial com-pression creep of London clay (Bishop, 1966) Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 3. GENERAL CHARACTERISTICS 467 Material Copyrighted Figure 12.2 Sustained stress creep curves illustrating different forms of strain vs. logarithm of time behavior. and Fig. 12.5 for undrained triaxial compression creep of soft Osaka clay (Murayama and Shi-bata, 1958). At any stress level (shown as a per-centage of the strength before creep in Fig. 12.4 and in kg/cm2 in Fig. 12.5), the logarithm of the strain rate decreases linearly with increase in the logarithm of time. The slope of this relationship is essentially independent of the creep stress; increases in stress level shift the line vertically upward. The slope of the log strain rate versus log time line for drained creep is approximately 1. Undrained creep often results in a slope be- Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 4. 468 12 TIME EFFECTS ON STRENGTH AND DEFORMATION Material Copyrighted Figure 12.3 Variation of creep strain rate with deviator stress for undrained creep of re-molded illite. tween 0.8 and 1 for this relationship. The onset of failure under higher stresses is signaled by a reversal in slope, as shown by the topmost curve in Fig. 12.5. 7. Pore pressure may increase, decrease, or remain constant during creep, depending on the volume change tendencies of the soil structure and whether or not drainage occurs during the de-formation process. In general, saturated soft sensitive clays under undrained conditions are most susceptible to strength loss during creep due to reduction in effective stress caused by increase in pore water pressure with time. Heav-ily overconsolidated clays under drained con-ditions are also susceptible to creep rupture due to softening associated with the increase in wa-ter content by dilation and swelling. 8. Although stress relaxation has been less studied than creep, it appears that equally regular pat-terns of deformation behavior are observed, for example, Larcerda and Houston (1973). 9. Deformation under sustained stress ordinarily produces an increase in stiffness under the ac-tion of subsequent stress increase, as shown schematically in Fig. 12.6. This reflects the time-dependent structural readjustment or ‘‘ag-ing’’ that follows changes in stress state. It is analogous to the quasi-preconsolidation effect Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 5. GENERAL CHARACTERISTICS 469 Material Copyrighted Figure 12.4 Strain rate vs. time relationships during drained creep of London clay (data from Bishop, 1966). due to secondary compression discussed in Sec-tion 8.11; however, it may develop under un-drained as well as drained conditions. 10. As shown in Fig. 12.7, the locations of both the virgin compression line and the value of the pre-consolidation pressure, , determined in the p laboratory are influenced by the rate of loading during one-dimensional consolidation (Graham et al., 1983a; Leroueil et al., 1985). Thus, esti-mations of the overconsolidation ratio of clay deposits in the field are dependent on the load-ing rates and paths used in laboratory tests for determination of the preconsolidation pressure. If it is assumed that the relationship between strain and logarithm of time during compression is linear over the time ranges of interest and that the secondary compression index Ce is constant regardless of load, the rate-dependent precon-solidation pressure p at ˙1 can be related to the axial strain rate as follows (Silvestri et al, 1986; Soga and Mitchell, 1996; Leroueil and Marques, 1996): Ce / (CcCr) p ˙1 ˙1 (12.1) p(ref) ˙1(ref) ˙1(ref) where Cc is the virgin compression index, Cr is the recompression index and p(ref) is the pre-consolidation pressure at a reference strain rate ˙ . In this equation, the rate effect increases 1(ref) with the value of Ce/(Cc Cr). The var-iation of preconsolidation pressure with strain rate is shown in Fig. 12.8 (Soga and Mitchell, 1996). The data define straight lines, and the slope of the lines gives the parameter . In gen-eral, the value of ranges between 0.011 and 0.094. Leroueil and Marques (1996) report val-ues between 0.029 and 0.059 for inorganic clays. 11. The undrained strength of saturated clay in-creases with increase in rate of strain, as shown in Figs. 12.9 and 12.10. The magnitude of the effect is about 10 percent for each order of mag-nitude increase in the strain rate. The strain rate Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 6. 470 12 TIME EFFECTS ON STRENGTH AND DEFORMATION Material Figure 12.5 Strain rate vs. time relationships during undrained creep of Osaka alluvial clay Copyrighted (Murayama and Shibata, 1958). Figure 12.6 Effect of sustained loading on (a) stress–strain and strength behavior and (b) one-dimensional compression behavior. effect is considerably smaller for sands. In a manner similar to Eq. (12.1), a rate parameter can be defined as the slope of a log–log plot of deviator stress at failure qƒ at a particular strain rate ˙1 relative to qƒ(ref), the strength at a reference strain rate ˙ , versus strain rate. 1(ref) This gives the following equation: qƒ ˙1 (12.2) qƒ(ref) ˙1(ref) The value of ranges between 0.018 and 0.087, similar to the rate parameter values used to define the rate effect on consolidation pressure in Eq. (12.1). Higher values of are associated with more metastable soil structures (Soga and Mitchell, 1996). Rate dependency decreases with increasing sample disturbance, which is consistent with this finding. 12.3 TIME-DEPENDENT DEFORMATION–STRUCTURE INTERACTION In reality, completely smooth curves of the type shown in the preceding figures for strain and strain rate as a function of time may not exist at all. Rather, as dis- Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 7. TIME-DEPENDENT DEFORMATION–STRUCTURE INTERACTION 471 Copyrighted Material Figure 12.7 Rate dependency on one-dimensional compression characteristics of Batiscan clay: (a) compression curves and (b) preconsolidation pressure (Leroueil et al., 1985). Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 8. 472 12 TIME EFFECTS ON STRENGTH AND DEFORMATION Material Figure 12.8 Strain rate dependence on preconsolidation pressure determined from one-dimensional constant strain rate tests (Soga and Mitchell, 1996). Copyrighted Figure 12.9 Effect of strain rate on undrained strength (Kulhawy and Mayne 1990). Re-printed with permission from EPRI. cussed by Ter-Stepanian (1992), a ‘‘jump-like structure reorganization’’ may occur, reflecting a stochastic char-acter for the deformation, as shown in Fig. 12.11 for creep of an undisturbed diatomaceous, lacustrine, ov-erconsolidated clay. Ter-Stepanian (1992) suggests that there are four levels of deformation: (1) the molecular level, which consists of displacement of flow units by surmounting energy barriers, (2) mutual displacement of particles as a result of bond failures, but without rearrangement, (3) the structural level of soil defor-mation involving mutual rearrangements of particles, and (4) deformation at the aggregate level. Behavior at levels 3 and 4 is discussed below; that at levels 1 and 2 is treated in more detail in Section 12.4. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 9. TIME-DEPENDENT DEFORMATION–STRUCTURE INTERACTION 473 Material Figure 12.10 Strain rate dependence on undrained shear strength determined using constant strain rate CU tests (Soga and Mitchell, 1996). Copyrighted Figure 12.11 Nonuniformity of creep in an undisturbed, di-atomaceous, lacustrine, overconsolidated clay (from Ter- Stepanian, 1992). Time-Dependent Process of Particle Rearrangement Creep can lead to rearrangement of particles into more stable configurations. Forces at interparticle contacts have both normal and tangential components, even if the macroscopic applied stress is isotropic. If, during the creep process, there is an increase in the proportion of applied deviator stress that is carried by interparticle normal forces relative to interparticle tangential forces, then the creep rate will decrease. Hence, the rate at which deformation level 3 occurs need not be uniform owing to the particulate nature of soils. Instead it will reflect a series of structural readjustments as particles move up, over, and around each other, thus leading to the somewhat irregular sequence of data points shown in Fig. 12.11. Microscopically, creep is likely to occur in the weak clusters discussed in Section 11.6 because the contacts in them are at limiting frictional equilibrium. Any small perturbation in applied load at the contacts or time-dependent loss in material strength can lead to sliding, breakage or yield at asperities. As particles slip, propped strong-force network columns are dis-turbed, and these buckle via particle rolling as dis-cussed in Section 11.6. To examine the effects of particle rearrangement, Kuhn (1987) developed a discrete element model that considers sliding at interparticle contacts to be visco-frictional. The rate at which sliding of two particles relative to each other occurs depends on the ratio of shear to normal force at their contact. The relationship between rate and force is formulated in terms of rate process theory (see Section 12.4), and the mechanistic representations of the contact normal and shear forces are shown in Fig. 12.12. The time-dependent compo-nent in the tangential forces model is given as a ‘‘sinh-dashpot’’. 2 The average magnitudes of both normal and 2Kuhn (1987) used the following equation for rate of sliding at a contact: 2kT F ƒt ˙X exp sinh h RT 2kTn ƒn 1 where n1 is the number of bonds per unit of normal force, ƒt is the tangential force and ƒn is the normal force. The others are parameters related to rate process theory as described in Section 12.4. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 10. 474 12 TIME EFFECTS ON STRENGTH AND DEFORMATION Copyrighted Material Figure 12.12 Normal and tangential interparticle force mod-els according to Kuhn (1987). Figure 12.13 Creep curves developed by numerical analysis of an irregular packing of circular disks (from Kuhn and Mitchell, 1993). tangential forces at individual contacts can change dur-ing deformation even though the applied boundary stresses are constant. Small changes in the tangential and normal force ratio at a contact can have a very large influence on the sliding rate at that contact. These changes, when summed over all contacts in the shear zone, result in a decrease or increase in the overall creep rate. A numerical analysis of an irregular packing of cir-cular disks using the sinh-dashpot representation gives creep behavior comparable to that of many soils as shown in Fig. 12.13 (Kuhn and Mitchell, 1993). The creep rate slows if the average ratio of tangential to normal force decreases, whereas it accelerates and may ultimately lead to failure if the ratio increases. In some cases, the structural changes that are responsible for the decreasing strain rate and increased stiffness may cause the overall soil structure to become more meta-stable. Then, after the strain reaches some limiting value, the process of contact force transfer from de-creasing tangential to increasing normal force reverses. This marks the onset of creep rupture as the structure begins to collapse. A similar result was obtained by Rothenburg (1992) who performed discrete particle simulations in which smooth elliptical particles were cemented with a model exhibiting viscous character-istics in both normal and tangential directions. Particle Breakage During Creep Particle breakage can contribute to time-dependent de-formation of sands (Leung et al., 1996; Takei et al., 2001; McDowell, 2003). Leung et al. (1996) performed one-dimensional compression tests on sands, and Fig. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 11. TIME-DEPENDENT DEFORMATION–STRUCTURE INTERACTION 475 Dry sand RD = 75% Pressure = 15.4 MPa After 5 days After Before 290 s Test Material Copyrighted Sieve Size (μm) 100 80 60 40 20 0 0 100 200 300 400 Percentage Passing (%) Figure 12.14 Changes in particle size distribution of sand before loading and after two different load durations (from Leung, et al., 1996). 12.14 shows the particle size distribution curves for samples before loading and after two different load du-rations. The amount of particle breakage increased with load duration. Microscopic observations revealed that angular protrusions of the grains were ground off, producing fines. The fines fill the voids between larger particles and crushed particles progressively rear-ranged themselves with time. Aging—Time-Dependent Strengthening of Soil Structure The structural changes that occur during creep that is continuing at a decreasing rate cause an increase in soil stiffness when the soil is subjected to further stress increase as shown in Fig. 12.6. Leonards and Alt-schaeffl (1964) showed that this increase in preconso-lidation pressure cannot be accounted for in terms of the void ratio decrease during the sustained compres-sion period. Time-dependent changes of these types are a consequence of ‘‘aging’’ effects, which alter the structural state of the soil. The fabric obtained by creep may be different from that caused by increase in stress, even though both samples arrive at the same void ratio. Leroueil et al. (1996) report a similar result for an ar-tificially sedimented clay from Quebec, as shown in Fig. 12.15a. They also measured the shear wave ve-locities after different times during the tests using bender elements and computed the small strain elastic shear modulus. Figure 12.15b shows the change in shear modulus with void ratio. Additional insight into the structural changes ac-companying the aging of clays is provided by the re-sults of studies by Anderson and Stokoe (1978) and Nakagawa et al. (1995). Figure 12.16 shows changes in shear modulus with time under a constant confining pressure for kaolinite clay during consolidation (An-derson and Stokoe, 1978). Two distinct phases of shear modulus–time response are evident. During primary consolidation, values of the shear modulus increase rapidly at the beginning and begin to level off as the excess pore pressure dissipates. After the end of pri-mary consolidation, the modulus increases linearly with the logarithm of time during secondary compres-sion. The expected change in shear modulus due to void ratio change during secondary compression can be es-timated using the following empirical formula for shear modulus as a function of void ratio and confining pressure (Hardin and Black, 1968): (2.97 e)2 G A p0.5 (12.3) 1 e where A is a unit dependant material constant, e is the void ratio, and p is the mean effective stress. The dashed line in Fig. 12.16 shows the calculated in-creases in the shear modulus due to void ratio decrease using Eq. (12.3). It is evident that the change in void ratio alone does not provide an explanation for the sec-ondary time-dependent increase in shear modulus. This aging effect has been recorded for a variety of mate-rials, ranging from clean sands to natural clays (Afifi and Richart, 1973; Kokusho, 1987; Mesri et al., 1990, and many others). Further discussion of aging phenom-ena is given in Section 12.11. Time-Dependent Changes in Soil Fabric Changes in soil fabric with time under stress influence the stability of soil structure. Changes in sand fabric with time after load application in one-dimensional compression were measured by Bowman and Soga (2003). Resin was used to fix sand particles after var-ious loading times. Pluviation of the sand produced a horizontal preferred particle orientation of soil grains, and increased vertical loading resulted in a greater ori-entation of particle long axes parallel to the horizontal, which is in agreement with the findings of Oda (1972a, b, c), Mitchell et al. (1976), and Jang and Frost (1998). Over time, however, the loading of sand caused parti-cle long axes to rotate toward the vertical direction (i.e., more isotropic fabric). Experimental evidence (Bowman and Soga, 2003) showed that large voids became larger, whereas small voids became smaller, and particles group or cluster Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 12. 476 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 2.6 2.4 2.2 2.0 Material 1.8 Copyrighted A A B B C C D D 4 6 8 10 20 Vertical Effective Stress σv (kPa) 0.5 1 2 E Small-strain Stiffness G0 (MPa) F E F Stiffness Change During the Primary Consolidation Between B and C Normal Consolidation Line Creep for 120 days Increase in Stiffness during Creep (C-D) Quasi-preconsolidation Pressure Destructuring State (a) (b ) 5 Void Ratio e Figure 12.15 (a) Compression curve and (b) variation of the maximum shear modulus G0 with void ratio for artificially sedimented Jonquiere clay (from Leroueil et al., 1996). together with time. Based on these particulate level findings, it appears that the movements of particles lead to interlocking zones of greater local density. The interlocked state may be regarded as the final state of any one particle under a particular applied load, due to kinematic restraint. The result, with time, is a stiffer, more efficient, load-bearing structure, with areas of rel-atively large voids and neighboring areas of tightly packed particles. The increase in stiffness is achieved by shear connections obtained by the clustering. Then, when load is applied, the increased stiffness and strength of the granular structure provides greater re-sistance to the load and the observed aging effect is seen. The numerical analysis in Kuhn and Mitchell (1993) led to a similar hypothesis for how a more ‘‘braced’’ structure develops with time. For load appli-cation in a direction different to that during the aging period, however, the strengthening effect of aging may be less, as the load-bearing particle column direction differs from the load direction. Time-Dependent Changes in Physicochemical Interaction of Clay and Pore Fluid A portion of the shear modulus increase during sec-ondary compression of clays is believed to result from a strengthening of physicochemical bonds between particles. To illustrate this, Nakagawa et al. (1995) ex-amined the physicochemical interactions between clays and pore fluid using a special consolidometer in which the sample resistivity and pore fluid conductivity could be measured. Shear wave velocities were obtained us-ing bender elements to determine changes in the stiff-ness characteristics of the clay during consolidation. Kaolinite clay mixed with saltwater was used for the experiment, and changes in shear wave velocities and electrical properties were monitored during the tests. The test results showed that the pore fluid compo-sition and ion mobility changed with time. At each load increment, as the effective stress increased with pore pressure dissipation, the shear wave velocities, and therefore the shear modulus, generally increased with time as shown in Fig. 12.17. It may be seen, how-ever, that in some cases, the shear wave velocities at the beginning of primary consolidation decreased slightly from the velocities obtained immediately be-fore application of the incremental load, probably as a result of soil structure breakdown. During the subse-quent secondary compression stage, the shear wave ve-locity again increased. As was the case for the results in Fig.12.16, the increases in shear wave velocity dur- Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 13. TIME-DEPENDENT DEFORMATION–STRUCTURE INTERACTION 477 50 40 30 20 10 0 0.5 1.0 1.5 2.0 IG = ΔG per log time = 6.2 MPa Copyrighted Material Primary Consolidation Secondary Compression Sample Height Change (mm) 101 102 103 104 Time (min) 100 101 102 103 104 Time (min) Ball Kaolinite Possible Change in G by Void Ratio Decrease Only Estimated Using Eq. (12.3) Initial Consolidation Pressure = 70 kPa Initial Void Ratio e0 = 1.1 Shear Modulus of less than γ = 10-3% (MPa) Figure 12.16 Modulus and height changes as a function of time under constant confining pressure for kaolinite: (a) shear modulus and (b) height change (from Anderson and Stokoe, 1978). Figure 12.17 Changes in shear wave velocity during pri-mary consolidation and secondary compression of kaolinite. Consolidation pressures: (a) 11.8 kPa and (b) 190 kPa (from Nakagawa et al., 1995). ing secondary compression are greater than can be ac-counted for by increase in density. The electrical conductivity of the sample measured by filter electrodes increased during the early stages of consolidation, but then decreased continuously there-after as shown in Fig. 12.18. The electrical conductiv-ity is dominated by flow through the electrolyte solution in the pores. During the initial compression, a breakdown of structure releases ions into the pore wa-ter, increasing the electrical conductivity. With time, the conductivity decreased, suggesting that the released ions are accumulating near particle surfaces. Some of these released ions are expelled from the specimen as consolidation progressed as shown in Fig. 12.18b. A slow equilibrium under a new state of effective stress is hypothesized to develop that involves both small particle rearrangements, associated with decrease in void ratio during secondary compression, and devel-opment of increased contact strength as a result of pre-cipitation of salts from the pore water and/or other processes. Primary consolidation can be considered a result of drainage of pore water fluid from the macropores, whereas secondary compression is related to the de-layed deformation of micropores in the clay aggregates (Berry and Poskitt, 1972; Matsuo and Kamon, 1977; Sills, 1995). The mobility of water in the micropores is restricted due to small pore size and physicochem-ical interactions close to the clay particle surfaces. Ak-agi (1994) did compression tests on specially prepared clay containing primarily Ca in the micropores and Na in the macropores. Concentrations of the two ions in the expelled water at different times after the start of consolidation were consistent with this hypothesis. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 14. 478 12 TIME EFFECTS ON STRENGTH AND DEFORMATION Activation Frequency The energy to enable a flow unit to cross a barrier may be provided by thermal energy and by various applied potentials. For a material at rest, the potential energy– displacement relationship is represented by curve A in Fig. 12.21. From statistical mechanics it is known that Copyrighted Material Figure 12.18 Changes in electrical conductivity of the pore water during primary consolidation and secondary compres-sion of kaolinite. Consolidation pressures: (a) 95 kPa and (b) 190 kPa (from Nakagawa et al., 1995). Figure 12.19 Energy barriers and activation energy. 12.4 SOIL DEFORMATION AS A RATE PROCESS Deformation and shear failure of soil involve time-dependent rearrangement of matter. As such, these phenomena are amenable for study as rate processes through application of the theory of absolute reaction rates (Glasstone et al., 1941). This theory provides both insights into the fundamental nature of soil strength and functional forms for the influences of sev-eral factors on soil behavior. Detailed development of the theory, which is based on statistical mechanics, may be found in Eyring (1936), Glasstone et al. (1941), and elsewhere in the physical chemistry literature. Adaptations to the study of soil behavior include those by Abdel-Hady and Her-rin (1966), Andersland and Douglas (1970), Christen-sen and Wu (1964), Mitchell (1964), Mitchell et al. (1968, 1969), Murayama and Shibata (1958, 1961, 1964), Noble and Demirel (1969), Wu et al. (1966), Keedwell (1984), Feda (1989, 1992), and Kuhn and Mitchell (1993). Concept of Activation The basis of rate process theory is that atoms, mole-cules, and/or particles participating in a time-dependent flow or deformation process, termed flow units, are constrained from movement relative to each other by energy barriers separating adjacent equilib-rium positions, as shown schematically by Fig. 12.19. The displacement of flow units to new positions re-quires the acquisition of an activation energy F of sufficient magnitude to surmount the barrier. The po-tential energy of a flow unit may be the same following the activation process, or higher or lower than it was initially. These conditions are shown by analogy with the rotation of three blocks in Fig. 12.20. In each case, an energy barrier must be crossed. The assumption of a steady-state condition is implicit in most applications to soils concerning the at-rest barrier height between successive equilibrium positions. The magnitude of the activation energy depends on the material and the type of process. For example, val-ues of F for viscous flow of water, chemical reac-tions, and solid-state diffusion of atoms in silicates are about 12 to 17, 40 to 400, and 100 to 150 kJ/mol of flow units, respectively. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 15. SOIL DEFORMATION AS A RATE PROCESS 479 Material Figure 12.20 Examples of activated processes: (a) steady-state, (b) increased stability, and (c) decreased stability. Copyrighted Figure 12.21 Effect of a shear force on energy barriers. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 16. 480 12 TIME EFFECTS ON STRENGTH AND DEFORMATION the average thermal energy per flow unit is kT, where k is Boltzmann’s constant (1.38 1023 J K1) and T is the absolute temperature (K). Even in a material at rest, thermal vibrations occur at a frequency given by kT/h, where h is Planck’s constant (6.624 1034 J s1). The actual thermal energies are divided among the flow units according to a Boltzmann distribution. It may be shown that the probability of a given unit becoming activated, or the proportion of flow units that are activated during any one oscillation is given by F p( F) exp (12.4) NkT Material Copyrighted where N is Avogadro’s number (6.02 1023), and Nk is equal to R, the universal gas constant (8.3144 J K1 mol1). The frequency of activation then is kT F exp (12.5) h NkT In the absence of directional potentials, energy bar-riers are crossed with equal frequency in all directions, and no consequences of thermal activations are ob-served unless the temperature is sufficiently high that softening, melting, or evaporation occurs. If, however, a directed potential, such as a shear stress, is applied, then the barrier heights become distorted as shown by curve B in Fig. 12.21. If ƒ represents the force acting on a flow unit, then the barrier height is reduced by an amount (ƒ/2) in the direction of the force and in-creased by a like amount in the opposite direction, where represents the distance between successive equilibrium positions.3 Minimums in the energy curve are displaced a distance from their original positions, representing an elastic distortion of the material struc-ture. The reduced barrier height in the direction of force ƒ increases the activation frequency in that direction to kT F/N ƒ/2 → exp (12.6) h kT and in the opposite direction, the increased barrier height decreases the activation frequency to kT F/N ƒ/2 ← exp (12.7) h kT 3Work (ƒ / 2) done by the force ƒ as the flow unit drops from the peak of the energy barrier to a new equilibrium position is assumed to be given up as heat. The net frequency of activation in the direction of the force then becomes kT F (→) (←) 2 exp ƒ sinh h RT 2kT (12.8) Strain Rate Equation At any instant, some of the activated flow units may successfully cross the barrier; others may fall back into their original positions. For each unit that is successful in crossing the barrier, there will be a displacement . The component of in a given direction times the number of successful jumps per unit time gives the rate of movement per unit time. If this rate of movement is expressed on a per unit length basis, then the strain rate ˙ is obtained. Let X F (proportion of successful barrier cross-ings and ) such that ˙ X[( →) ( ←)] (12.9) Then from Eq. (12.8) kT F ƒ ˙ 2X exp sinh (12.10) h RT 2kT The parameter X may be both time and structure de-pendent. If (ƒ/2kT) 1, then sinh(ƒ/2kT) (ƒ/2kT), and the rate is directly proportional to ƒ. This is the case for ordinary Newtonian fluid flow and diffusion where 1 ˙ (12.11) where ˙ is the shear strain rate, is dynamic viscosity, and is shear stress. For most solid deformation problems, however, (ƒ/2kT) 1 (Mitchell et al., 1968), so ƒ 1 ƒ sinh exp (12.12) 2kT 2 2kT and kT F ƒN ˙ X expexp(12.13) h RT 2RT Equation (12.13) applies except for very small stress intensities, where the exponential approximation of the hyperbolic sine is not justified. Equations (12.10) and Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 17. BONDING, EFFECTIVE STRESSES, AND STRENGTH 481 (12.13) or comparable forms have been used to obtain dashpot coefficients for rheological models, to obtain functional forms for the influences of different factors on strength and deformation rate, and to study defor-mation rates in soils. For example, Kuhn and Mitchell (1993) used this form as part of the particle contact law in discrete element modeling as described in the previous section. Puzrin and Houlsby (2003) used it as an internal function of a thermomechanical-based model and derived a rate-dependent constitutive model for soils. Copyrighted Material Soil Deformation as a Rate Process Although there does not yet appear to be a rigorous proof of the correctness of the detailed statistical me-chanics formulation of rate process theory, even for simple chemical reactions, the real behavior of many systems has been substantially in accord with it. Dif-ferent parts of Eq. (12.13) have been tested separately (Mitchell et al., 1968). It was found that the tempera-ture dependence of creep rate and the stress depend-ence of the experimental activation energy [Eq. (12.14)] were in accord with predictions. These results do not prove the correctness of the theory; they do, however, support the concept that soil deformation is a thermally activated process. Arrhenius Equation Equation (12.13) may be written kT E ˙ X exp (12.14) h RT where ƒN E F (12.15) 2 is termed the experimental activation energy. For all conditions constant except T, and assuming that X(kT/h) constant A, E ˙ A exp (12.16) RT Equation (12.16) is the same as the well-known em-pirical equation proposed by Arrhenius around 1900 to describe the temperature dependence of chemical reaction rates. It has been found suitable also for characterization of the temperature dependence of processes such as creep, stress relaxation, secondary compression, thixotropic strength gain, diffusion, and fluid flow. 12.5 BONDING, EFFECTIVE STRESSES, AND STRENGTH Using rate process theory, the results of time-dependent stress–deformation measurements in soils can be used to obtain fundamental information about soil structure, interparticle bonding, and the mecha-nisms controlling strength and deformation behavior. Deformation Parameters from Creep Test Data If the shear stress on a material is and it is distributed uniformly among S flow units per unit area, then ƒ (12.17) S Displacement of a flow unit requires that interatomic or intermolecular forces be overcome so that it can be moved. Let it be assumed that the number of flow units and the number of interparticle bonds are equal. If D represents the deviator stress under triaxial stress conditions, the value of ƒ on the plane of max-imum shear stress is D ƒ (12.18) 2S so Eq. (12.13) becomes kT F D ˙ X expexp(12.19) h RT 4SkT This equation describes creep as a steady-state process. Soils do not creep at constant rate, however, because of continued structural changes during de-formation as described in Section 12.3, except for the special case of large deformations after mobiliza-tion of full strength. Thus, care must be taken in ap-plication of Eq. (12.19) to ensure that comparisons of creep rates and evaluations of the influences of differ-ent factors are made under conditions of equal struc-ture. The time dependency of creep rate and the possible time dependencies of the parameters in Eq. (12.19) are considered in Section 12.8. Determination of Activation Energy From Eq. (12.14) ln(˙/T) E (12.20) (1/T) R provided strain rates are considered under conditions of unchanged soil structure. Thus, the value of E can be determined from the slope of a plot of ln(˙ /T) ver-sus (1/T). Procedures for evaluation of strain rates for Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 18. 482 12 TIME EFFECTS ON STRENGTH AND DEFORMATION soils at different temperatures but at the same structure are given by Mitchell et al. (1968, 1969). Determination of Number of Bonds For stresses large enough to justify approximating the hyperbolic sine function by a simple exponential in the creep rate equation and small enough to avoid tertiary creep, the logarithm of strain rate varies directly with the deviator stress. For this case, Eq. (12.19) can be written ˙ K(t) exp(D) (12.21) Material 4A procedure for evaluation of from the results of a test at a succession of stress levels on a single sample is given by Mitchell Copyrighted et al. (1969). where kT F K(t) X exp (12.22) h RT (12.23) 4SkT Parameter is a constant for a given value of ef-fective consolidation pressure and is given by the slope of the relationship between log strain rate and stress. It is evaluated using strain rates at the same time after the start of creep tests at several stress intensities. With known, /S is calculated as a measure of the number of interparticle bonds.4 Activation Energies for Soil Creep Activation energies for the creep of several soils and other materials are given in Table 12.1. The free energy of activation for creep of soils is in the range of about 80 to 180 kJ/mol. Four features of the values for soils in Table 12.1 are significant: 1. The activation energies are relatively large, much 2. Variations in water content (including complete 3. The values for sand and clay are about the same. 4. Clays in suspension with insufficient solids to Table 12.1 Activation Energies for Creep of Several Materials Material higher than for viscous flow of water. drying), adsorbed cation type, consolidation pres-sure, void ratio, and pore fluid have no significant effect on the required activation energy. form a continuous structure deform with an ac-tivation Activation Energy energy equal to that of water. (kJ/ mol)a Reference 1. Remolded illite, saturated, water contents of 30 to 43% 105–165 Mitchell, et al. (1969) 2. Dried illite: samples air-dried from saturation, then evacuated 155 Mitchell, et al. (1969) 3. San Francisco Bay mud, undisturbed 105–135 Mitchell, et al. (1969) 4. Dry Sacramento River sand 105 Mitchell, et al. (1969) 5. Water 16–21 Glasstone, et al. (1941) 6. Plastics 30–60 Ree and Eyring (1958) 7. Montmorillonite–water paste, dilute 84–109 Ripple and Day (1966) 8. Soil asphalt 113 Abdel-Hady and Herrin (1966) 9. Lake clay, undisturbed and remolded 96–113 Christensen and Wu (1964) 10. Osaka clay, overconsolidated 120–134 Murayama and Shibata (1961) 11. Concrete 226 Polivka and Best (1960) 12. Metals 210 Finnie and Heller (1959) 13. Frozen soils 393 Andersland and Akili (1967) 14. Sault Ste. Marie clay, suspensions, discontinuous structures Same as water Andersland and Douglas (1970) 15. Sault Ste. Marie clay, Li, Na, K forms, in H2O and CCl4, consolidated 117 Andersland and Douglas (1970) aThe first four values are experimental activation energies, E. Whether the remainder are values of F or E is not always clear in the references cited. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 19. BONDING, EFFECTIVE STRESSES, AND STRENGTH 483 Number of Interparticle Bonds Evaluation of S requires knowledge of , the separation distance between successive equilibrium positions in the interparticle contact structure. A value of 0.28 nm (2.8 A˚ ) has been assumed because it is the same as the distance separating atomic valleys in the surface of a silicate mineral. It is hypothesized that deformation in-volves the displacement of oxygen atoms along con-tacting particle surfaces, as well as periodic rupture of Copyrighted Material bonds at interparticle contacts. Figure 12.22 shows this interpretation for schematically. If the above as-sumption for is incorrect, calculated values of S will still be in the same correct relative proportion as long as remains constant during deformation. Normally Consolidated Clay Results of creep tests at different stress intensities for different consolidation pressures enable computation of S as a function of con-solidation pressure. Values obtained for undisturbed San Francisco Bay mud are shown in Fig. 12.23. The open point is for remolded bay mud. An undisturbed specimen was consolidated to 400 kPa, rebounded to Figure 12.22 Interpretation of in terms of silicate mineral surface structure. 50 kPa, and then remolded at constant water content. The effective consolidation pressure dropped to 25 kPa as a result of the remolding. The drop in effective stress was accompanied by a corresponding decrease in the number of interparticle bonds. Tests on remolded illite gave comparable results. A continuous inverse re-lationship between the number of bonds and water content over a range of water contents from more than 40 percent to air-dried and vacuum-desiccated clay is shown in Fig. 12.24. The dried material had a water content of 1 percent on the usual oven-dried basis. The very large number of bonds developed by drying is responsible for the high dry strength of clay. Overconsolidated Clay Samples of undisturbed San Francisco Bay mud were prepared to overcon-solidation ratios of 1, 2, 4, and 8 following the stress paths shown in the upper part of Fig. 12.25. The sam-ple represented by the triangular data point was re-molded after consolidation and unloading to point d, where it had a water content of 52.3 percent. The un-drained compressive strength as a function of consol-idation pressure is shown in the middle section of Fig. 12.25, and the number of bonds, deduced from the creep tests, is shown in the lower part of the figure. The effect of overconsolidation is to increase the num-ber of interparticle bonds over the values for normally consolidated clay. Some of the bonds formed during consolidation are retained after removal of much of the consolidation pressure. Values of compressive strength and numbers of bonds from Fig. 12.25 are replotted versus each other in Fig. 12.26. The resulting relationship suggests that strength depends only on the number of bonds and is independent of whether the clay is undisturbed, re-molded, normally consolidated, or overconsolidated. Dry Sand Creep tests on oven-dried sand yielded results of the same type as obtained for clay, as shown in Fig. 12.27, suggesting that the strength-generating and creep-controlling mechanisms may be similar for both types of material. Composite Strength-Bonding Relationship Values of S and strength for many soils are combined in Fig. 12.28. The same proportionality exists for all the ma-terials, which may seem surprising, but which in reality should be expected, as discussed further later. Significance of Activation Energy and Bond Number Values The following aspects of activation energies and num-bers of interparticle bonds are important in the under-standing of the deformation and strength behavior of uncemented soils. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 20. 484 12 TIME EFFECTS ON STRENGTH AND DEFORMATION Material Figure 12.23 Number of interparticle bonds as a function of consolidation pressure for normally consolidated San Francisco Bay mud. Copyrighted Figure 12.24 Number of bonds as a function of water con-tent for illite. 1. The values of activation energy for deformation of soils are high in comparison with other ma-terials and indicate breaking of strong bonds. 2. Similar creep behavior for wet and dry clay and for wet and dry sand indicates that deformation is not controlled by viscous flow of water. 3. Comparable values of activation energy for wet and dry soil indicate that water is not respon-sible for bonding. 4. Comparable values of activation energy for clay and sand support the concept that interparticle bond strengths are the same for both types of material. This is supported also by the unique-ness of the strength versus number of bonds re-lationship for all soils. 5. The activation energy and presumably, there-fore, the bonding type are independent of con-solidation pressure, void ratio, and water content. 6. The number of bonds is directly proportional to effective consolidation pressure for normally consolidated clays. 7. Overconsolidation leads to more bonds than in normally consolidated clay at the same effective consolidation pressure. 8. Strength depends only on the number of bonds. 9. Remolding at constant water content causes a decrease in the effective consolidation pressure, which means also a decrease in the number of bonds. 10. There are about 100 times as many bonds in dry clay as in wet clay. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 21. BONDING, EFFECTIVE STRESSES, AND STRENGTH 485 Material Copyrighted Figure 12.25 Consolidation pressure, strength, and bond numbers for San Francisco Bay mud. Although it may be possible to explain these results in more than one way, the following interpretation ac-counts well for them. The energy F activates a mole of flow units. The movement of each flow unit may involve rupture of single bonds or the simultaneous rupture of several bonds. Shear of dilute montmoril-lonite– water pastes involves breaking single bonds (Ripple and Day, 1966). For viscous flow of water, the activation energy is approximately that for a single hy-drogen bond rupture per flow unit displacement, even though each water molecule may form simultaneously up to four hydrogen bonds with its neighbors. If the single-bond interpretation is also correct for soils, then consistency in Eq. (12.10) requires that shear force ƒ pertain to the force per bond. On this basis, parameter S indicates the number of single bonds per unit area. In the event activation of a flow unit requires simul-taneous rupture of n bonds, then S represents 1/nth of the total bonds in the system. That the activation energy for deformation of soil is well into the chemical reaction range (40 to 400 kJ/ mol) does not prove that bonding is of the primary Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 22. 486 12 TIME EFFECTS ON STRENGTH AND DEFORMATION (2) it is the same for wet and dry soils, and (3) it is essentially the same for different adsorbed cations and Copyrighted Material Figure 12.26 Strength as a function of number of bonds for San Francisco Bay mud. Figure 12.28 Composite relationship between shear strength and number of interparticle bonds (from Matsui and Ito, 1977). Reprinted with permission from The Japanese Society of SMFE. valence type because simultaneous rupture of several weaker bonds could yield values of the magnitude ob-served. On the other hand, the facts that (1) the acti-vation energy is much greater than for flow of water, Figure 12.27 Strength as a function of number of bonds for dry Antioch River sand. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 23. BONDING, EFFECTIVE STRESSES, AND STRENGTH 487 pore fluids (Andersland and Douglas, 1970) suggest that bonding is through solid interparticle contacts. Physical evidence for the existence of solid-to-solid contact between clay particles has been obtained in the form of photomicrographs of particle surfaces that were scratched during shear (Matsui et al., 1977, 1980) and acoustic emissions (Koerner et al., 1977). Activation energy values of 125 to 190 kJ/mol are of the same order as those for solid-state diffusion of oxygen in silicate minerals. This supports the concept that creep movements of individual particles could re-sult Copyrighted Material from slow diffusion of oxygen ions in and around interparticle contacts. The important minerals in both sand and clay are silicates, and their surface layers consist of oxygen atoms held together by silicon at-oms. Water in some form is adsorbed onto these sur-faces. The water structure consists of oxygens held together by hydrogen. It is not too different from that of the silicate layer in minerals. Thus, a distinct bound-ary between particle surface and water may not be dis-cernable. Under these conditions, a more or less continuous solid structure containing water molecules that propagates through interparticle contacts can be visualized. An individual flow unit could be an atom, a group of atoms or molecules, or a particle. The preceding arguments are based on the interpretation that individ-ual atoms are the flow units. This is consistent with both the relative and actual values of S that have been determined for different soils. Furthermore, by using a formulation of the rate process equation that enabled calculation of the flow unit volume from creep test data, Andersland and Douglas (1970) obtained a value of about 1.7 A˚ 3, which is of the same order as that of individual atoms. On the other hand, Keedwell (1984) defined flow units between quartz sand particles as consisting of six O2 ions and six Si4 ions and be-tween two montmorillonite clay particles as consisting of four H2O molecules. If particles were the flow units, not only would it be difficult to visualize their thermal vibrations, but then S would relate to the number of interparticle contacts. It is then difficult to conceive how simply drying a clay could give a 100-fold increase in the number of inter-particle contacts, as would have to be the case accord-ing to Fig. 12.27. A more plausible interpretation is that drying, while causing some increase in the number of interparticle contacts during shrinkage, causes mainly an increase in the number of bonds per contact because of increased effective stress. At any value of effective stress, the value of S is about the same for both sand and clay. The number of interparticle contacts should be vastly different; how-ever, for equal numbers of contacts per particle, the number per unit volume should vary inversely with the cube of particle size. Thus, the number of clay particles of 1-m particle size should be some nine orders of magnitude greater than for a sand of 1-mm average particle size. Each contact between sand particles would involve many bonds; in clay, the much greater number of contacts would mean fewer bonds per par-ticle. The contact area required to develop bonds in the numbers indicted in Figs. 12.23 to 12.27 is very small. For example, for a compressive strength of 3 kg/cm2 ( 300 kPa) there are 8 1010 bonds/cm2 of shear surface. Oxygen atoms on the surface of a silicate min-eral have a diameter of 0.28 nm. Allowing an area 0.30 nm on a side for each oxygen gives 0.09 nm2, or 9 1016 cm2, per bonded oxygen for a total area of 9 1016 8 1010 7.2 105 cm2/cm2 of soil cross section. Hypothesis for Bonding, Effective Stress, and Strength Normal effective stresses and shear stresses can be transmitted only at interparticle contacts in most soils.5 The predominant effects of the long-range physico-chemical forces of interaction are to control the initial soil fabric and to alter the forces transmitted at contact points from what they would be due to applied stresses alone. Interparticle contacts are effectively solid, and it is likely that both adsorbed water and cations in the con-tact zone participate in the structure. An interparticle contact may contain many bonds that may be strong, approaching the primary valence type. The number of bonds at any contact depends on the compressive force transmitted at the contact, and the Terzaghi–Bowden and Tabor adhesion theory of friction presented in Sec-tion 11.4, can account for strength. The macroscopic strength is directly proportional to the number of bonds. For normally consolidated soils the number of bonds is directly proportional to the effective stress. As a re-sult of particle rearrangements and contacts formed during virgin compression, an overconsolidated soil at a given effective stress has a greater number of bonds and higher strength than a normally consolidated soil. This effect is more pronounced in clays than in sands because the larger and bulky sand grains tend to re- 5 Pure sodium montmorillonite may be an exception since a part of the normal stress can be carried by physicochemical forces of inter-action. The true effective stress may be less than the apparent effec-tive stress by R A as discussed in Chapter 7. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 24. 488 12 TIME EFFECTS ON STRENGTH AND DEFORMATION cover their original shapes when unloaded, thus rup-turing most of the bonds in excess of those needed to resist the lower stress. The strength of the interparticle contacts can vary over a wide range, depending on the number of bonds per contact. The unique relationship between strength and num-ber of bonds for all soils, as indicted by Fig. 12.28, reflects the fact that the minerals comprising most soils are silicates, and they all have similar surface struc-tures. Copyrighted Material In the absence of chemical cementation, interparticle bonds may form in response to interparticle contact forces generated by either applied stresses, physico-chemical forces of interaction, or both. Any bonds ex-isting in the absence of applied effective stress, that is, when 0, are responsible for true cohesion. There should be no difference between friction and cohesion in terms of the shearing process. Complete failure in shear involves simultaneous rupture or slipping of all bonds along the shear plane. 12.6 SHEARING RESISTANCE AS A RATE PROCESS Deformation at large strain can approach a steady-state condition where there is little further structural change with time (such as at critical state). In this case, Eq. (12.19) can be used to describe the shearing resistance as a function of strain rate and temperature. If the max-imum shear stress is substituted for the deviator stress D, then 1 3 (12.24) 2 and kT ˙ X exp F exp (12.25) h RT 2SkT Taking logarithms of both sides of Eq. (12.25) gives kT F ln ˙ lnX (12.26) h RT 2SkT By assuming X(kT/h) is a constant equal to B (Mitchell, 1964), Eq. (12.26) can be rearranged to give 2S 2SkT ˙ F ln(12.27) N B From the relationships in Section 12.5, the following relationship between bonds per unit area and effective stress is suggested. S a b (12.28) ƒ where a and b are constants and ƒ is the effective normal stress on the shear plane. Thus, Eq. (12.27) becomes 2a F 2akT ˙ 2b F 2bkT ˙ ln ln N B N B ƒ (12.29) Equation (12.29) is of the same form as the Cou-lomb equation for strength: c tan (12.30) ƒ By analogy, 2a F 2akT ˙ c ln (12.31) N B 2b F 2bkT ˙ tan ln (12.32) N B These equations state that both cohesion and friction depend on the number of bonds times the bond strength, as reflected by the activation energy, and that the values of c and should depend on the rate of deformation and the temperature. Strain Rate Effects All other factors being equal, the shearing resistance should increase linearly with the logarithm of the rate of strain. This is shown to be the case in Fig. 12.9, which contains data for 26 clays. Additional data for several clays are shown in Fig. 12.29, where shearing resistance as a function of the speed of vane rotation in a vane shear test is plotted. Analysis of the relation-ship between shearing stress and angular rate of vane rotation shows that / log decreases with an increase in water content. This follows directly from Eq. (12.29) because d 2akT 2bkT 2kT (a b) d ln(˙ /B) ƒ ƒ (12.33) that is, d /d ln (˙ /B) is proportional to the number of bonds, which decreases with increasing water content. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 25. CREEP AND STRESS RELAXATION 489 Material Copyrighted Figure 12.29 Effect of rate of shear on shearing resistance of remolded clays as determined by the laboratory vane apparatus (prepared from the data of Karlsson, 1963). This interpretation of the data in Figs. 12.9 and 12.29 assumes that the effective stress was unaffected by changes in the strain rate, which may not necessarily be true in all cases. Effect of Temperature Assumptions of reasonable values for parameters show that the term (˙ /B) is less than one (Mitchell, 1964). Thus the quantity ln(˙ /B) in Eq. (12.29) is negative, and an increase in temperature should give a decrease in strength, all other factors being constant. That this is the case is demonstrated by Fig. 12.30, which shows deviator stress as a function of temperature for samples of San Francisco Bay mud compared under conditions of equal mean effective stress and structure. Other ex-amples of the influence of temperature on strength are shown in Figs. 11.6 and 11.133. 12.7 CREEP AND STRESS RELAXATION Although the designation of a part of the strain versus time relationship as steady state or secondary creep may be convenient for some analysis purposes, a true steady state can exist only for conditions of constant structure and stress. Such a set of conditions is likely only for a fully destructured soil, and a fully destruc-tured state is likely to persist only during deformation at a constant rate, that is, at failure. This state is often called ‘‘steady state,’’ in which the soil is deforming continuously at constant volume under constant shear and confining stresses (Castro, 1975; Castro and Poulos, 1977). Otherwise, bond making and bond breaking occur at different rates as a result of different internal time-and strain-dependent phenomena, which might include thixotropic hardening, viscous flows of water and ad- Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 26. 490 12 TIME EFFECTS ON STRENGTH AND DEFORMATION Copyrighted Material Figure 12.30 Influence of temperature on the shearing re-sistance of San Francisco Bay mud. Comparison is for sam-ples at equal mean effective stress and at the same structure. Figure 12.31 Variation of creep strain rate with deviator stress for drained creep of London clay (data from Bishop, 1966). sorbed films, chemical, and biological transformations, and the like. Furthermore, distortions of the soil struc-ture and relative movements between particles cause changes in the ratio of tangential to normal forces at interparticle contacts that may be responsible for large changes in creep rate. Because of these time depend-encies some of the parameters in Eq. (12.19) may be time dependent. For example, Feda (1989) accounted for the time dependency of creep rate by taking changes in the number of structural bonds into account. Therefore, application of Eq. (12.19) for the determi-nation of the bonding and effective stress relationships discussed in Section 12.5 required comparison of creep rates under conditions of comparable time and struc-ture. The influence of creep stress magnitude on the creep rate at a given time after the application of the stress to identical samples of a soil was shown in Fig. 12.3. At low stresses the creep rates are small and of little practical importance. The curve shape is compatible with the hyperbolic sine function predicted by rate process theory, as given by Eq. (12.10). In the mid-range of stresses, a nearly linear relationship is found between logarithm of strain rate and stress, also as pre-dicted by Eq. (12.10) for the case where the argument of the hyperbolic sine is greater than 1. At stresses approaching the strength of the material, the strain rate becomes very large and signals the onset of failure. Other examples of the relationships between logarithm of strain rate and creep stress corresponding to differ-ent times after the application of the creep stress are given in Fig. 12.31 for drained tests on London clay and Fig. 12.32 for undrained tests on undisturbed San Francisco Bay mud. Only values for the midrange of stresses are shown in Figs. 12.31 and 12.32. Effect of Composition In general, the higher the clay content and the more active the clay, the more important are stress relaxation and creep, as illustrated by Figs. 4.22 and 4.23, where creep rates, approximated by steady-state values, are related to clay type, clay content, and plasticity. Time-dependent deformations are more important at high water contents than at low. Deviatoric creep and sec-ondary compression are greater in normally consoli-dated than overconsolidated soils. Although the magnitude of creep strains and strain rates may be small in sand or dry soil, the form of the Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 27. CREEP AND STRESS RELAXATION 491 is a special case of volumetric creep. Deviatoric creep is often accompanied by volumetric creep. The ratio of volumetric to deviatoric creep fol- Copyrighted Material Figure 12.32 Variation of creep strain rate with deviator stress for undrained creep of normally consolidated San Fran-cisco Bay mud. 1 0.8 0.6 0.4 0.2 Drained Triaxial Test Confining Pressure σ3 = 414 kPa Deviator Stress from 344 to 377 kPa A B C D E 2820 min 1450 min 90 min 20 min 2 min 0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 Deviator strain (%) 1.0 0.8 0.6 0.4 0.2 0 (a) 0 1 2 3 4 Strain increment ratio dεv/dεs during creep (b) Volumetric Strain (%) Stress ratio q/p Figure 12.33 Dilatancy relationship obtained from drained creep tests on kaolinite: (a) development of volumetric and deviatoric strains with time and (b) effect of stress ratio on strain increment ratio d/ds (from Walker, 1969). behavior conforms with the patterns described and il-lustrated above. This is to be expected, as the basic creep mechanism is the same in all inorganic soils.6 Water may ‘‘lubricate’’ the particles and possibly in-crease the creep rate even though the basic mechanism of creep is the same for dry and wet materials (Losert et al., 2000). Takei et al. (2001) showed that the de-velopment of creep strains due to time-dependent breakage of talc specimens increased more for satu-rated specimens than dry ones. However, a negligible effect of water on creep rate was reported by Ahn-Dan et al. (2001) who performed creep tests on unsaturated and saturated crushed gravel and by Leung et al. (1996) who performed one-dimensional compression 6Volumetric creep and secondary compression of organic soils, peat, and municipal waste fills can develop also as a result of decompo-sition of organic matter. creep tests on sands. The conflicting evidence may be due to the presence or absence of impurities that may lubricate or cement the soil in the presence of water (Human, 1992; Bowman, 2003). Volume Change and Pore Pressures Due to the known coupling effects between shearing and volumetric plastic deformations in soils, an in-crease in either mean pressure or deviator stress can generate both types of deformations. Creep behavior is no exception. Time-dependent shear deformations are usually referred to as deviatoric creep or shear creep. Time-dependent deformations under constant stress re-ferred to as volumetric creep. Secondary compression Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 28. 492 12 TIME EFFECTS ON STRENGTH AND DEFORMATION lows a plastic dilatancy rule. Walker (1969) inves-tigated the time-dependent change of these two com-ponents from incremental drained triaxial creep tests on normally consolidated kaolinite. The increase in shear strains with increase in volumetric strains at dif-ferent times is shown in Fig. 12.33a. At the beginning of the triaxial test, the deviator stress was instantane-ously increased from 344 to 377 kPa and kept constant. pressure increase with time for consolidated undrained creep tests on illite at several stress intensities is shown in Fig. 12.34. Figure 12.35 shows a slow decrease in Copyrighted Material After an immediate increase in shear strains at constant volume (AB in Fig. 12.33a), section BD corresponds to primary consolidation that is controlled by the dis-sipation of pore pressures. After point D, creep oc-curred, and the ratio of volumetric to deviatoric strains was independent of time. This ratio decreased with in-creasing stress ratio as shown in Fig. 12.33b. This ob-servation led to the time-dependent flow rule, which is similar to the dilatancy rule described in Section 11.20. Sand deforms with time in a similar manner. Under progressive deviatoric creep, the volumetric creep re-sponse is highly dependent on density, the stress level, and the stress path before creep. The rate of both vol-umetric and deviatoric creep increases with confining pressure, particularly after particle crushing becomes important at high stresses (Yamamuro and Lade, 1993). For dense sand under high deviator stress, dilative creep is observed (Murayama et al., 1984; Mejia et al., 1988). The volumetric response of dense sand and gravel with time is a highly complex function of stress history, with some samples contracting or dilating (Lade and Liu, 1998; Ahn-Dan et al., 2001). Some dense sand samples contract initially but then dilate with time (Bowman and Soga, 2003). Further discus-sion of the creep behavior of sands in relation to me-chanical aging phenomena is given in Section 12.11. The fundamental process of creep strain develop-ment is therefore similar to that of time-independent plastic strains, and the same framework of soil plastic-ity can possibly be used. It can be argued whether it is necessary to separate the deformation into time-dependent and independent components. Rate-independent behavior can be considered as the limiting case of rate-dependent behavior at a very slow rate of loading. Volumetric-deviatoric creep coupling implies that rapid application of a stress or a strain invariably re-sults in rapid change of pore water pressures in a sat-urated soil under undrained conditions. For a constant total minor principal stress, the magnitude of the pore pressure change depends on the volume change ten-dencies of the soil when subjected to shear distortions. These tendencies are, in turn, controlled by the void ratio, structure, and effective stress, and can be quan-tified in terms of the pore pressure parameter A as dis-cussed in Chapters 8 and 10. An example showing pore Figure 12.34 Pore pressure development with time during undrained creep of illite. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 29. CREEP AND STRESS RELAXATION 493 Material Figure 12.35 Normalized pore pressure vs. time relationships during creep of kaolinite. Copyrighted pore pressure during the sustained loading of kaolinite. Similar behavior was demonstrated in the measured stress paths of undrained creep test on San Francisco Bay mud (Arulanandan et al., 1971). As shown in Fig. 12.36, the effective stress states shifted toward the fail-ure line. At higher stress levels, the specimens even-tually underwent creep rupture. However, soil strength in terms of effective stresses does not change unless there are chemical, biological, or mineralogical changes during the creep period. This is illustrated by the stress paths shown schematically in Fig. 12.37, where the pre- and postcreep strengths fall on the same failure envelope. Effects of Temperature An increase in temperature decreases effective stress, increases pore pressure, and weakens the soil structure. Creep rates ordinarily increase and the relaxation stresses corresponding to specific values of strain de-crease at higher temperature. These effects are illus-trated by the data shown in Figs. 12.38 and 12.39. Effects of Test Type, Stress System, and Stress Path Most measurements of time-dependent deformation and stress relaxation in soils have been done on sam-ples consolidated isotropically and tested in triaxial compression or by measurement of secondary com-pression in oedometer tests. However, most soils in nature have been subjected to an anisotropic stress his-tory, and deformation conditions conform more to plane strain than triaxial compression in many cases. Some investigations of these factors have been made. Although the general form of the stress–strain–time and stress–strain rate–time relationships are similar to those shown above for triaxial loading conditions, the actual values may differ considerably. For example, undisturbed Haney clay, a gray silty clay from British Columbia, with a sensitivity in the range of 6 to 10, was tested both in triaxial compres-sion and plane strain (Campanella and Vaid, 1974). Samples were normally consolidated both isotropically and under K0 conditions to the same vertical effective stress. Samples consolidated isotropically were tested in triaxial compression. Coefficient K0 consolidation was used for both K0 triaxial and plane strain tests. The results shown in Fig. 12.40 indicate that the pre-creep stress history had a significant effect on the deformations. The plane strain and K0 consolidated triaxial samples gave about the same creep behavior under the same deviatoric stress, which suggests that preventing strain in one horizontal direction and/or the intermediate principal stress were not factors of major importance for this soil under the test conditions used. Interaction Between Consolidation and Creep Experimental evidence suggests that creep occurs dur-ing primary consolidation (Leroueil et al., 1985; Imai and Tang, 1992). Following the initial large change following load application, the pore pressure may ei-ther dissipate, with accompanying volume change if drainage is allowed, or change slowly during creep or stress relaxation, if drainage is prevented. The devel-opment of complete effective stress and void ratio equilibrium may take a long time. One illustration of Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 30. 494 12 TIME EFFECTS ON STRENGTH AND DEFORMATION Material 0 10 20 30 40 50 60 70 80 Mean Pressure p (kPa) Copyrighted 0 80 160 240 320 400 480 560 640 Figure 12.36 Measured stress paths of undrained creep tests of San Francisco Bay mud. Initial confining pressure: (a) 49 kPa and (b) 392 kPa (from Arulanandan et al., 1971). 50 40 30 20 10 400 320 160 80 0 Normal Undrained Triaxial Compression Test: Effective Stress Path Total Stress Path of Triaxial Compression Test 1 3 Possible Critical State Line Effective Stress State After 1,000 min Creep After 20,000 min Creep Effective Stress Path of Undrained Creep 0 Normal Undrained Triaxial Compression Test: Effective Stress Path Total Stress Path of Triaxial Compression Test 1 3 Possible Critical State Line Effective Stress State After 1,000 min Creep After 20,000 min Creep Effective Stress Path of Undrained Creep 240 (a) (b) Deviator Stress q (kPa) Deviator Stress q (kPa) Mean Pressure p (kPa) Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 31. CREEP AND STRESS RELAXATION 495 Figure 12.37 Effects of undrained creep on the Material strength of normally consolidated clay. Copyrighted Figure 12.38 Creep curves for Osaka clay tested at different temperatures—undrained tri-axial compression (Murayama, 1969). this is given by Fig. 10.5, where it is shown that the relationship between void ratio and effective stress is dependent on the time for compression under any given stress. Another is given by Fig. 12.41, which shows pore pressures during undrained creep of San Francisco Bay mud. In each sample, consolidation un-der an effective confining pressure of 100 kPa was al-lowed for 1800 min prior to the cessation of drainage and the start of a creep test. The consolidation period was greater than that required for 100 percent primary consolidation. The curve marked 0 percent stress level refers to a specimen maintained undrained but not sub-ject to a deviator stress. This curve indicates that each of the other tests was influenced by a pore pressure that contained a contribution from the prior consoli-dation history. The magnitude and rate of pore pressure develop-ment if drainage is prevented following primary con-solidation depend on the time allowed for secondary compression prior to the prevention of further drain-age. This is illustrated by the data in Fig. 12.42, which show pore pressure as a function of time for samples that have undergone different amounts of secondary compression. In summary, creep deformation depends on the ef-fective stress path followed and any changes in stress Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 32. 496 12 TIME EFFECTS ON STRENGTH AND DEFORMATION Figure 12.39 Influence of temperature on the Material initial and final stresses in stress relaxation tests on Osaka clay—undrained triaxial compression (Murayama, 1969). Copyrighted Figure 12.40 Creep curves for isotropically and K0-consolidated samples of undisturbed Haney clay tested in triaxial and plane strain compression (from Campanella and Vaid, 1974). Reproduced with permission from the National Research Council of Canada. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 33. RATE EFFECTS ON STRESS–STRAIN RELATIONSHIPS 497 Figure 12.41 Pore pressure development during undrained creep of San Francisco Bay mud after consolidation at 100 kPa for 1800 min Material (from Holzer et al., 1973). Reproduced with permission from the National Research Council of Canada. Copyrighted Figure 12.42 Pore pressure development under undrained conditions following different periods of secondary compression (from Holzer et al., 1973). Reproduced with permission from the National Research Council of Canada. with time. Furthermore, time-dependent volumetric re-sponse is governed both by the rate of volumetric creep and by the rate of consolidation. The latter is a com-plex function of drainage conditions and material prop-erties, especially the permeability and compressibility. Because the effective stress path is controlled by the rate of loading and drainage conditions, the separation of consolidation and creep deformations can be diffi-cult in the early stage of time-dependent deformation as given by section BD in Fig. 12.33a. In some cases, a fully coupled analysis of soil–pore fluid interaction with an appropriate time-dependent constitutive model is necessary to reconcile the time-dependent deforma-tions observed in the field and laboratory. 12.8 RATE EFFECTS ON STRESS–STRAIN RELATIONSHIPS An increase in strain rate during soil compression is manifested by increased stiffness, as was noted in Sec-tion 12.3. In essence, the state of the soil jumps to the stress–strain curve that corresponds to the new strain rate. Commonly, this rate-dependent stress–strain Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 34. 498 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 5% /h 0.5% /h Copyrighted Material 0.05% /h Belfast Clay 4 m σ1c = σv0 16%/h 1%/h 0.25%/h Winnipeg Clay 11.5 m σ1c σv0 CAU Triaxial Compression Tests Relaxation Tests (R) 0 0.6 0.5 0.4 0.3 0.2 0.1 0 4 8 12 16 20 Axial Strain (σ1 – σ3)/2σ1c R R R Effective Stress σv (kPa) εv2 . SP1 test SP2 test εv1 . εv1 . = 2.70 10–6 s–1 εv2 = 1.05 10–7 s–1 εv2 . εv1 . . . εv1 εv1 εv1 . εv1 . . . εv3 . εv2 . εv2 ε.v3 = 1.34 10–5 s–1 0 5 10 15 20 25 30 0 50 100 150 200 250 Strain εv (%) Figure 12.43 Rate-dependent stress–strain relations of clays: (a) undrained triaxial compression tests of Belfast and Winnipeg clay (Graham et al., 1983a) and (b) one-dimensional compression tests of Batiscan clay (Leroueil et al., 1985). curve, noted by S˘ uklje (1957), is the same as if the soil had been loaded from the beginning at the new strain rate. This phenomenon is often observed in clays. Examples are given in Fig. 12.43a for undrained triaxial compression tests of Belfast and Winnipeg clays (Graham et al., 1983a) and Fig. 12.43b for one-dimensional compression tests of Batiscan clay (Ler-oueil et al., 1985). Yield and Strength Envelopes of Clays The undrained shear strength and apparent precon-solidation pressure of soils decrease with decreasing strain rate or increasing duration of testing. Preconsol-idation pressures obtained from one-dimensional con-solidation tests and undrained shear strengths obtained from triaxial tests are just two points on a soil’s yield envelope in stress space. For a given metastable soil structure, the degree of rate dependency of preconsol-idation pressure is similar to that of undrained shear strength (Soga and Mitchell, 1996). If the apparent pre-consolidation pressure depends on the strain rate at which the soil is deforming, then the same analogy can be expanded to the assumption that the size of the en-tire yield envelope is also strain rate dependent (Tav-enas and Leroueil, 1977). Figure 12.44 shows a family of strength envelopes corresponding to constant strain rates7 obtained from drained and undrained creep tests on stiff plastic Mascouche clay from Quebec (Leroueil and Marques, 1996). The effective stress failure line of soil is uniquely defined regardless of the magnitude of the strain rate applied in undrained compression. Figure 12.45a shows the failure line of Haney clay (Vaid and Cam-panella, 1977). The line represents the stress conditions at the maximum ratio of /. The data were obtained 1 3 by various undrained tests, and a unique failure line can be observed. Figure 12.45b shows the undrained stress paths and the critical state line of reconstituted mixtures of sand and clay with plasticity indices rang-ing from 10 to 30 (Nakase and Kamei, 1986). A unique critical state line can be observed although the rates of shearing are different. The change in undrained shear strength with strain rate results from a difference in generation of excess pore pressures. A decrease in strain rate leads to larger excess pore pressures at fail-ure due to creep deformation. 7The strain rate is defined as ˙ ˙2 ˙2, where ˙ is the volu- vs v s v metric strain rate and ˙s is the deviator strain rate (Leroueil and Mar-ques, 1996). Whether the use of this strain rate measure is appropriate or not remains to be investigated. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 35. RATE EFFECTS ON STRESS–STRAIN RELATIONSHIPS 499 Material Copyrighted Figure 12.44 Influence of strain rate on the yield surface of Masouche clay (from Leroueil and Marques, 1996). Excess Pore Pressure Generation in Normally Consolidated Clays Excess pore pressure development depends primarily on the collapse of soil structure. Accordingly, strain is the primary factor controlling pore pressure generation. This is shown in Fig. 12.46 by the undrained stress– strain–pore pressure response of normally consolidated natural Olga clay (Lefebvre and LeBouef, 1987). The natural clay specimens were normally consolidated un-der consolidation pressures larger than the field over-burden pressure and then sheared at different strain rates. Although the deviator stress at any strain in-creases with increasing strain rate, the pore pressure versus strain curves are about the same at all strain rates. At a given deviator stress, the pore pressure gen-eration was larger at slower strain rates as a result of more creep under slow loading. This is consistent with the observation made in connection with undrained creep of clays as discussed in Section 12.7. Strain-driven pore pressure generation was also suggested by Larcerda and Houston (1973) who showed that pore pressure does not change significantly during triaxial stress relaxation tests in which the axial strain is kept constant. Overconsolidated Clays Rate dependency of undrained shear strength decreases with increasing overconsolidation, since there is no contraction or collapse tendency observed during creep of heavily overconsolidated clays. Sheahan et al. (1996) prepared reconstituted specimens of Boston blue clay at different overconsolidation ratios and sheared them at different strain rates in undrained con-ditions. Figure 12.47 shows that the undrained stress path and the strength were much more strain rate de-pendent for lightly overconsolidated clay (OCR 1 and 2) than for more heavily overconsolidated clay (OCR 4 and 8). The results also show that the strength failure envelope is independent of strain rate as discussed earlier. The strain rate effects on stress–strain–pore pressure response of overconsolidated structured Olga clays are shown in Fig. 12.48 (Lefebvre and LeBouef, 1987). The natural samples were reconsolidated to the field Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 36. 500 12 TIME EFFECTS ON STRENGTH AND DEFORMATION Const. Stress Creep Const. Rate of Strain Shear Const. Rate of Loading Shear Aged Samples Const. Load Creep Step Creep Thixotropic Hardened 0.1 0 0 0.1 0.2 Material Line Copyrighted State Critical-Line K0-Critical-State Line 0 0.2 0.2 0.3 0.3 0.4 0.4 0.4 0.5 0.6 0.6 0.7 0.8 M-15 710-1: ε1 710-2: ε.2 710-3: ε3 0.8 (σ1 + σ3)/2σ1c (σ1 – σ3)/2σ1c (a) q/σvc (b) 1.0 0.8 0.6 0.4 0.2 0 –0.2 –0.4 –0.6 1.0 p/σvc M-10 . ε(%/min) . . 710-1: ε1 710-2: ε2 710-3: ε3 Critical-State Line K0-Line Critical-State Line 0 0.2 0.4 0.6 0.8 1.0 p/σvc q/σvc 1.0 0.8 0.6 0.4 0.2 0 –0.2 –0.4 –0.6 . ε(%/min) . . . M15 Soil (Plasticity Index = 15) M10 Soil (Plasticity Index = 10) Figure 12.45 Strain rate independent failure line: (a) Haney clay (from Vaid and Campa-nella, 1977) and (b) reconstituted mixtures of sand and clay (from Nakase and Kamei, 1986). overburden pressure. The deformation is brittle, with strain softening indicating development of localized shear failure planes. Up to the peak stress, the response follows what has been described previously, that is the stress–strain response is rate dependent and the pore pressure generation is strain dependent but indepen-dent of rate. However, after the peak, the pore pressure generation becomes rate dependent. This is due to local drainage within the specimens as the deformation be-comes localized. As the time to failure increased, there is more opportunity for local drainage toward the di-lating shear band and the measured pore pressure may not represent the overall behavior of the specimens. The difference in softening due to swelling at the fail- Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 37. RATE EFFECTS ON STRESS–STRAIN RELATIONSHIPS 501 Normally Consolidated Olga Clay Undrained Triaxial Compression Tests Initial Isotropic Confining Pressure =137 kPa Axial Strain Rate Material Copyrighted 1 2 3 4 5 6 7 8 Axial Strain (%) Axial Strain Rate 0.1 %/hr 0.5 %/hr 2.6 %/hr 12.3 %/hr 1 2 3 4 5 6 7 8 Axial Strain (%) 120 100 80 60 40 20 0 140 120 100 80 60 40 20 0 0.1 %/hr 0.5 %/hr 2.6 %/hr 12.3 %/hr Excess Pore Pressure Δu (kPa) Deviator Stress q (kPa) Figure 12.46 Stress–strain and pore pressure–strain curves for normally consolidated Olga clay (from Lefebvre and LeBouef, 1987). Overconsolidated Olga Clay Undrained Triaxial Compression Tests Initial Isotropic Confining Pressure = 17.6 kPa Axial Strain Rate 1 2 3 4 5 6 7 8 Axial Strain (%) Axial Strain Rate 0.1 %/hr 0.5 %/hr 2.6 %/hr 12.3 %/hr 1 2 3 4 5 6 7 8 Axial Strain (%) 70 60 50 40 20 10 0 10 0 0.1 %/hr 0.5 %/hr 2.5 %/hr 12.3 %/hr 30 20 Excess Pore Pressure Δu (kPa) Deviator Stress q (kPa) Figure 12.48 Stress–strain and pore pressure–strain curves for overconsolidated Olga clay (from Lefebvre and LeBouef, 1987). OCR=1 OCR=2 OCR=4 OCR=8 OCR=2 OCR=4 Effective Stress Path for Axial Strain Rate = 0.50 %/hr 0.2 0.4 0.6 0.8 0.4 0.3 0.2 0.1 0.0 -0.1 Effective Stress State at Peak for Axial Strain Rate = 0.051 %/hr Axial Strain Rate = 0.50 %/hr Axial Strain Rate = 5.0 %/hr Axial Strain Rate = 49 %/hr Initial K0 Consolidation State (σa + σr)/2σvm (σa – σr)/2σvm OCR=1 Large Rate OCR=8 Effect Negligible Rate Effect Figure 12.47 Rate dependency stress path and strength of overconsolidated Boston blue clay (from Sheahan et al., 1996). ure plane results in apparent rate dependency at large strains. Similar observations were made by Atkinson and Richardson (1987) who examined local drainage effects by measuring the angles of intersection of shear bands with very different times of failure. Rate Effects on Sands Similar rate-dependent stress–strain behavior is ob-served in sands (Lade et al., 1997), but the effects are quite small in many cases (Tatsuoka et al., 1997; Di Benedetto et al., 2002). An example of time depend-ency observed for drained plane strain compression tests of Hostun sand is shown in Fig. 12.49 (Matsushita et al., 1999). The stress–strain curves for three differ-ent strain rates (1.25 101, 1.25 102, and 1.25 103 %/min) are very similar, indicating very small rate effects when the specimens are sheared at a con-stant strain rate. On the other hand, the change from one rate to another temporarily increases or decreases the resistance to shear. The influence of acceleration rather than the rate is reflected by the significant creep deformation and stress relaxation of this rate-insensitive material as shown the figure. This is differ- Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 38. 502 12 TIME EFFECTS ON STRENGTH AND DEFORMATION 6.0 5.5 5.0 4.5 4.0 3.5 3.0 Variation of Stress-strain curve by Constant Strain Rate Tests at Axial Strain Rates = 0.125, 0.0125 and 0.00125 %/min. A Very Small Rate Effect Is Observed for Continuous Loading. CRS at 0.125%/min Creep Material Copyrighted Creep Creep CRS at 0.125%/min Stress Relaxation CRS at 0.00125%/min CRS at 0.125%/min Creep CRS at 0.125%/min CRS at 0.00125%/min Accidental Pressure Drop Followed by Relaxation Stage 0 1 2 3 4 5 6 7 8 Stress Ratio σa /σr Shear Strain γ = εa – εr (%) Figure 12.49 Creep and stress relaxation of Hostun sand (from Matsushita et al., 1999). NSF-Clay Isotropically Consolidated Esec = Δq/εa, Eeq = (Δq)SA / (εa)SA Emax = 239 MPa p0 = 300 kPa 300 200 100 0 Young's Modulus, Esec or Eeq (MPa) 10-3 10-2 10-1 100 Axial Strain, εa or Single Amplitude of Cyclic Axial Strain, (εa)SA (%) Figure 12.50 Clay stiffness degradation curves at three strain rates (from Shibuya et al., 1996). Coefficient of Strain Rate, (γ) Shear Strain, γ (%) Figure 12.51 Strain rate parameter G and strain level for several clays (from Lo Presti et al., 1996). ent from the observations made for clays as shown in Fig. 12.43 in which a unique stress–strain–strain rate relationship was observed. Hence, the modeling of stress–strain–rate behavior of sands appears to be more complicated than that of clays, and further in-vestigation is needed, as time-dependent behavior of sands can be of significance in geotechnical construc-tion as discussed further in Section 12.10. Stiffness at Small and Intermediate Strains Although the magnitude is small, the strain rate de-pendency of the stress–strain relationship is observed even at small strain levels for clays. The stiffness in-creases less than 6 percent per 10-fold increase in strain rate (Leroueil and Marques, 1996). The rate de-pendency on stiffness degradation curves measured by monotonic loading of a reconstituted clay is shown in Fig. 12.50 (Shibuya et al., 1996). At different strain levels, the increase in the secant shear modulus with shear strain rate ˙ is often expressed by the following equation (Akai et al., 1975; Isenhower and Stokoe, 1981; Lo Presti et al., 1996; Tatsuoka et al., 1997): G ( ) (12.34) G log ˙ G( , ˙ ) ref where G is the increase in secant shear modulus with increase in log strain rate log ˙ , and G( , ˙ ) is the ref secant shear modulus at strain and reference strain rate ˙ . The magnitude is large in clays, considerably ref less in silty and clayey sands, and small in clean sands (Lo Presti et al., 1996; Stokoe et al., 1999). The vari-ation of G with strain is shown in Fig. 12.51 for dif- Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 39. MODELING OF STRESS–STRAIN–TIME BEHAVIOR 503 ferent plasticity clays (Lo Presti et al., 1996). The magnitude of rate dependency increases with strain level, especially for strain levels larger than 0.01 per-cent, which is within the preplastic region zone 3 de-scribed in Section 11.17. Rate Effects During Cyclic Loading The frequencies of cyclic loading to which a soil is subjected can vary widely. For example, the frequency of sea and ocean waves is in the range of 102 to 101 Hz, earthquakes are in the range of 0.1 to a few hertz, and machine foundations are in the range of 10 to 100 Hz. Similarly to monotonic loading, the effect of load-ing problems requiring the determination of de- Copyrighted Material frequency on shear modulus degradation is small in clean, coarse-grained soils (Bolton and Wilson, 1989; Stokoe et al., 1995), but the effect becomes more significant in fine-grained soils (Stokoe et al., 1995; d’Onofrio et al., 1999; Matesic and Vucetic, 2003; Meng and Rix, 2004). An example of frequency effects on a shear modulus degradation curve for a clay ob-tained from cyclic loading is shown in Fig. 12.50 along with the monotonic data. Figure 12.52 shows the effect of frequency on shear modulus of several soils at very small shear strains (less than 103 percent) measured by torsional shear and resonant column apparatuses (Meng and Rix, 2004). The effect is 10 percent in-crease per log cycle at most. At a given frequency of cyclic loading, the strain rate applied to a soil increases with applied shear strain as shown by the equation below: ˙ 4ƒ (12.35) c 250 200 150 100 50 where ƒ is the frequency and c is the cyclic shear strain amplitude. Using Eq. (12.35), Matesic and Vu-cetic (2003) report values of G of 2 to 11 percent for clays and 0.2 to 6 percent for sands as the strain rate increases 10-fold. The values of G in general de-creased when the applied cyclic shear strain increased from 5 104 percent to 1 102 percent. It should be noted that the strain range examined is within the non-linear elastic range (zone 1 to zone 2 in Section 11.17). The monotonic loading data presented in Fig. 12.51 show that the rate effect becomes more pro-nounced at larger strain, that is, as plastic deformations become more significant. Hence, it is possible that the fundamental mechanisms of rate dependency are dif-ferent at small elastic strain levels than at larger plastic strains. Small strain damping shows more complex fre-quency dependency, as shown in Fig. 12.53 (Shibuya et al., 1995; Meng and Rix, 2004). At a frequency of more than 10 Hz, the damping ratio increases with increased frequency, possibly due to pore fluid viscos-ity effects. As the applied frequency decreases, the damping ratio decreases. However, at a frequency less than 0.1 Hz, the damping ratio starts to increase with decreasing frequency. This may result from creep of the soil (Shibuya et al., 1995). 12.9 MODELING OF STRESS–STRAIN–TIME BEHAVIOR Constitutive models are needed for the solution of geo-technical Sandy Elastic Silt 0 10-2 10-1 100 101 102 Frequency (Hz) Sandy Silty Clay Kaolin Vallencca Clay Kaolin Sandy Lean Clay Kaolin Subgrade Fat Clay Shear Modulus G (MPa) Figure 12.52 Rate dependency of cyclic small strain stiffness of a sandy elastic silt (from Meng and Rix, 2004). Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 40. 504 12 TIME EFFECTS ON STRENGTH AND DEFORMATION Clayey Subgrades Sandy Lean Clay 10-2 10-1 100 101 102 Frequency Material (Hz) Copyrighted 6 5 4 3 2 1 Sandy Silty Clay Sandy Elastic Silt Kaolin Vallencca Clay Fat Clay Pisa Clay Augusta Clay Damping (%) Figure 12.53 Effect of strain rate of damping ratio of soils (from Shibuya et al., 1995 and Meng and Rix, 2004). formations, displacements, and strength and stability changes that occur over time periods of different lengths. Various approaches have been used, including empirical curve fitting, extensions of rate process theory, rheological models, and advanced theories of viscoelasticity and viscoplasticity. Owing to the complexity of stress states, the many factors that influ-ence the creep and stress relaxation properties of a soil, and the difficulty of accounting for concurrent volu-metric and deviatoric deformations in systems that are many times undergoing consolidation as well as sec-ondary compression or creep, it is not surprising that development of general models that can be readily im-plemented in engineering practice is a challenging un-dertaking. Nonetheless, some progress has been made in estab-lishing functional forms and relationships that can be applied for simple analyses and comparisons, and one of these is developed in this section. A complete re-view and development of all recent theories and pro-posed relationships for creep and stress relaxation is beyond the scope of this book. Comprehensive reviews of many models for representation of the time-dependent plastic response of soils are given in Adachi et al. (1996). General Stress–Strain–Time Function Strain Rate Relationships between axial strain rate ˙ and time t of the type shown in Figs. 12.4 and 12.5 can be expressed by ˙t ln m ln (12.36) ˙(t ,D) t 1 1 or t ln ˙ ln ˙(t ,D) m ln (12.37) 1 t1 where ˙(t ,D) is the axial strain rate at unit time and 1 is a function of stress intensity D, m is the absolute value of the slope of the straight line on the log strain rate versus log time plot, and t1 is a reference time, for example, 1 min. Values of m generally fall in the range of 0.7 to 1.3 for triaxial creep tests; lower values are reported for undrained conditions than for drained con-ditions. For the development shown here, the stress intensity D is taken as the deviator stress (1 3). A shear stress or stress level could also be used. The same data plotted in the form of Figs. 12.3, 12.31, and 12.32 can be expressed by ˙ln D (12.38) ˙(t,D ) 0 or ln ˙ ln ˙(t,D ) D (12.39) 0 in which ˙(t,D ) is a fictitious value of strain rate at 0 D 0, a function of time after start of creep, and is the slope of the linear part of the log strain rate versus stress plot. From Eqs. (12.37) and (12.39) t ln ˙(t ,D) m ln 1 ln ˙(t,D ) 0 D (12.40) t 1 For D 0, Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com