Open access
Technical Papers
Mar 10, 2020

Soil Response to Repetitive Changes in Pore-Water Pressure under Deviatoric Loading

Publication: Journal of Geotechnical and Geoenvironmental Engineering
Volume 146, Issue 5

Abstract

Soils often experience repetitive changes in pore water pressure. This study explores the volumetric and shear response of contractive and dilative sand specimens subjected to repetitive changes in pore water pressure, under constant deviatoric stress in a triaxial cell. The evolution towards a terminal void ratio eT characterizes the volumetric response. The terminal void ratio eT for pressure cycles falls below the critical state line, between emin<eT<ecs. Very dense specimens only dilate if they reach high stress obliquity ηmax during pressurization. The terminal void ratios for very dense and medium dense specimens do not converge to a single trend. The shear deformation may stabilize at shakedown, or continue in ratcheting mode. The maximum stress obliquity ηmax is the best predictor of the asymptotic state; shakedown prevails in all specimens subjected to stress obliquity ηmax<0.95·ηcs and ratcheting takes place when the maximum stress obliquity approaches or exceeds ηmax0.95·ηcs. Volumetric and shear strains can accumulate when the strain level during pressure cycles exceeds the volumetric threshold strain (about 5×104 in this study). A particle-level analysis of contact loss and published experimental data show that the threshold strain increases with confinement po.

Introduction

Soils experience repetitive changes in pore water pressure during groundwater level oscillations associated with tidal and river level fluctuations, and engineered structures such as docks and managed reservoirs (O’Reilly and Brown 1991; Chu et al. 2003; Orense et al. 2004; Leroueil et al. 2009; Page et al. 2010; Nakata et al. 2013; Shi et al. 2016). Coupled processes may also cause pore-fluid pressure oscillations, for example, in the case of a soft clay subjected to temperature cycles (Abuel-Naga et al. 2007).
Pore-fluid pressure fluctuations affect a wide range of geotechnical systems from foundations and slope stability to pumped-storage hydroelectric power stations, aquifer storage and recovery systems, compressed air energy storage, enhanced oil recovery by cyclic water flooding and cyclic steam injection, and repetitive CO2 injection (Premchitt et al. 1986; Olson et al. 2000; Gambolati and Teatini 2015; Huang 2016; Chang et al. 2017).
Soils gradually deform in response to all kinds of repetitive excitations. Repetitive changes in water pressure imply effective stress cycles that can lead to the accumulation of plastic volumetric and shear strains. This study explores the volumetric and shear response of contractive and dilative sands subjected to repetitive changes in pore water pressure under constant deviatoric stress. The following section presents a detailed review of the state of the art and identifies salient gaps in knowledge.

Previous Studies: Asymptotic States

Pore-Fluid Pressure Oscillation

Previous studies explored the effects of repetitive changes in pore-fluid pressure in the context of engineering needs, such as slope failures (Nakata et al. 2013) or aquifer oscillations (Hung et al. 2012). The selected test boundary conditions reflected field situations: triaxial stress (clays, Ohtsuka and Miyata 2001; Ohtsuka 2007), plane strain conditions (sand, Nakata et al. 2013), and Ko-conditions (sand-silt mixtures, Chang et al. 2017). In all cases, the strain accumulation induced by repetitive changes in pore-fluid pressure became more significant with an increasing pressure amplitude Δuw. However, previous studies did not separate the volumetric response from the shear response; these are analyzed next.

Volumetric Asymptotic State: Terminal Void Ratio

All soils evolve towards an asymptotic terminal void ratio during repetitive loading (Narsilio and Santamarina 2008—Refer to the p-e quadrant in Fig. 1). The tendency towards a terminal state is apparent in published data for all types of repetitive loads: pore-water pressure cycles (Chang et al. 2017), Ko-loading and deviatoric stress cycles (Triantafyllidis et al. 2004; Wichtmann et al. 2005; Chong and Santamarina 2016), freeze-thaw (Viklander 1998), dry-wet (Albrecht and Benson 2001), and chemical cycles (Musso et al. 2003). There are irreversible structural changes during repetitive loading and soil properties adapt as the soil transitions towards the terminal void ratio; for example, the fabric changes of clays during dry-wet cycles (Croney and Coleman 1954), permeability increases in freeze-thaw cycles and dry-wet cycles (Chamberlain et al. 1990; Albrecht and Benson 2001), shear strength increases in freeze-thaw cycles (Ono and Mitachi 1997; Qi et al. 2006), and there is gradual stiffening in cyclic Ko loading (Park and Santamarina 2019).
Fig. 1. Anticipated soil response to pore water pressure cycles under constant deviatoric loading. The plot captures asymptotic conditions in shear strain (shakedown or ratcheting) and volumetric strain (terminal void ratio eT).

Shear Asymptotic State: Shakedown or Ratcheting?

The asymptotic condition in shear governs the response of all structures, from pavements (Sharp and Booker 1984) to metals (Johnson 1986). The shear response falls into one of three asymptotic regimes, as observed on the q-γ quadrant in Fig. 1 (Alonso-Marroquin and Herrmann 2004; Werkmeister et al. 2005)
Elastic shakedown: non-hysteretic, fully recoverable deformation in every cycle.
Plastic shakedown: hysteretic stress-strain response without permanent deformation at the end of each cycle.
Ratcheting: the stress strain response is hysteretic, and there is continued plastic strain accumulation in every cycle.
The asymptotic condition depends on the stress amplitude ratio, the cyclic stress ratio Δσzamp/2po, and the cyclic shear stress level Δτzθamp/Δσzamp (Wu et al. 2017; Cai et al. 2018; Gu et al. 2018). Ratcheting should be expected at large stress amplitudes and high stress obliquity η=q/p. It may also develop when a large number of cycles reaches a fatigue-induced tipping point, or when the stress level causes particle crushing (see data in Werkmeister 2003; Alonso-Marroquin and Herrmann 2004; Werkmeister et al. 2005; Wichtmann et al. 2005; da Fonseca et al. 2013).

Experimental Study

Tested Sand: Properties

Table 1 summarizes the main characteristics of the “KAUST 20/30 sand” used throughout this study and includes index properties such as the particle shape, the coefficient of uniformity Cu, and the extreme void ratios emax and emin. Measured values are compared against predicted values from index properties for self-consistent verification (refer to Table 1 for details).
Table 1. Tested sand—Properties
PropertyKAUST 20/30 sandObservations (Verifications using index test data)
Particle diameterd=0.600.85 
RoundnessR=0.60Image analysis—Roundness R=ri/N. The average radius of curvature of surface features divided by the radius of the largest inscribed sphere rmax
Coefficient of uniformityCu=1.20 
Specific gravityGs=2.65 
Maximum void ratioemax=0.786Estimated maximum void ratio: emax=0.76 (Youd 1973)
Minimum void ratioemin=0.533Estimated minimum void ratio: emin=0.54 (Cho et al. 2006)
Friction angle at constant volume shearϕcs=31°Angle of repose method: ϕcs=32° (Santamarina and Cho 2001) Inferred from roundness is ϕcs=31°±2° (Cho et al. 2006)
Critical state line CSL in e-logpΓ=0.845Intercept of CSL at 1 kPa
λ=0.074Slope of CSL
Shear wave velocity parametersα=89  m/sCorresponds to Hertzian-based power model Vs=α·[po/kPa]β
β=0.21Note: α and β values for eo=0.65
Estimated threshold strain for contact loss in monotonic loadingγth|loss=5×104Based on contact-loss analysis: γth|loss=1.3·(σo/Gg)2/3
Confining stress σ=250  kPa
Mineral: Gg=30  GPa and υ=0.25 (assumed in analysis)

Note: Measured values are compared against values predicted from index properties for self-consistent verification.

The critical state provides a “reference asymptotic state” for this study. Fig. 2 shows data for a set of conventional consolidated-undrained CU triaxial tests projected onto p-q-εz-e-u planes. Table 1 lists the critical state parameters obtained for the KAUST 20/30 sand. Measured values are consistent with values estimated from index properties (Refer to Table 1 for details).
Fig. 2. Conventional consolidated, undrained CU triaxial test data projected onto uw-εz-q-p-e planes (strain rate: εz=0.01/min). Notation: p=(σ1+σ3)/2,q=(σ1+σ3)/2, ϕcs=sin1(tanα), and stress obliquity η=q/p. Critical state parameters for the KAUST 20/30 sand: friction angle ϕcs=31°, intercept of CSL at 1 kPa in e-logp=0.845, and slope of CSL in e-log p=0.074. For reference, the maximum and minimum void ratios are emax=0.786 and emin=0.533.

Experimental Devices and Configuration

The triaxial system used to conduct the repetitive pressure cycles consists of (1) a triaxial cell with an LVDT (Linear Variable Differential Transformer) to track the vertical displacement, (2) a loading frame to apply a constant deviatoric stress, and (3) a pressure panel that generates cyclic changes in pore water pressure and measures volume changes.

Sample Preparation

We prepare loose, medium dense, and dense specimens using a combination of raining and tamping techniques to obtain different initial relative densities between Dr=15% and 70%.

Loading Histories

We can simulate the effects of changes in water pressure through changes in either the back pressure or the confining pressure (Brenner et al. 1985; Anderson and Sitar 1995; Farooq et al. 2004; Orense et al. 2004). The two test procedures yield the same results if the Biot’s coefficient χ=1Bsk/Bg remains close to χ1.0, that is at a relatively low confining effective stress (note: Bsk is the bulk modulus of the soil skeleton, and Bg is the bulk modulus of the mineral that makes the grains, Skempton 1961; Santamarina et al. 2001). In this study, we control the back pressure uw.
Fig. 3 presents a subset of the stress paths explored in this study. Typical loading histories consist of five stages (1) isotropic consolidation, (2) drained deviatoric loading to stress obliquity η=0.33, (3) a decrease in back pressure uw to reach η=0.20, (4) repetitive changes in pore water pressure from η=0.20 to ηmax=0.50 for N=100 loading cycles (shown in red), and (5) strain-controlled undrained axial compression from η=0.20 to failure at a vertical strain rate of εz=0.01/min. Table 2 summarizes the experimental study. Test parameters include the initial void ratio eo, cyclic pressure amplitude Δuw, and maximum stress obliquity ηmax=q/pmin. Cyclic pressure amplitudes Δuw selected for this study represent various field conditions, such as tidal action (<170  kPa at Burntcoat Head and Leaf Basin in North America), seasonal fluctuations in ground water levels (70–100 kPa,—Hung et al. 2012; Huang 2016), injection pressures for injection wells and injection-recovery wells used in aquifer storage (<500  kPa,—Shi et al. 2016; Page et al. 2010), and some coupled processes (e.g., geothermally-induced Δuw<250  kPa,—Laloui 2001; Abuel-Naga et al. 2007).
Fig. 3. Stress paths on the p-q space (a subset of cases is shown here. Refer to Table 1 for a complete description). The loading history consists of five stages: (1) isotropic consolidation, (2) drained deviatoric loading to stress obliquity η=0.33, (3) decrease back pressure uw to reach η=0.20, (4) repetitive change in pore water pressure from η=0.20 to ηmax=0.50 (shown in red), and (5) strain-controlled undrained axial compression from η=0.20 to failure at a strain rate of εz=0.01/min. Notation: p=(σ1+σ3)/2, q=(σ1+σ3)/2, ϕcs=sin1(tanα).
Table 2. Experimental study: Test conditions
Specimen characteristicsPressure cyclesCU—AC
Stress conditionsVoid ratio eShear strain γa
Conditions before pressure cycles relative to critical stateTest No.Void ratio after specimen preparationB-valueq [kPa]Δuw [kPa]pmax [kPa]ηmaxeoeTmbN*bγ1abcdVolume change tendency during final undrained shear
Contractive side10.74610.972501002500.330.70650.70001.0493.01×1040.0900.01500.00150Contractive
20.73900.981501102500.360.70270.69600.95438.74×1050.1130.02050.00130No CU-AC
30.74190.976501252500.400.70310.68781.0222.11×1030.2340.02500.00600Contractive
40.72080.977501402500.450.68280.66210.85194.70×1030.4100.02400.00800Dilative
50.70920.976501502500.500.67220.64480.88197.70×1030.7750.02300.01600Dilative
60.71800.981752253750.500.66600.63610.79137.92×1030.8000.022000Dilative
Dilative side70.69580.98425751250.500.67830.65981.0262.90×1030.6250.02200.0160Dilative
80.61380.95050552500.260.59000.59150.912.86×1040.0030.04000.000030No CU-AC
90.61710.95250852500.300.60210.60321.013.02×1040.0020.028000No CU-AC
100.61510.951501252500.400.60150.60300.914.45×1040.0100.01500.00010No CU-AC
110.61960.966501402500.450.60250.60291.012.57×1030.0750.02000.00010Dilative
120.61450.977501532500.520.60000.60081.013.06×1030.0800.02000.00010Dilative
130.61330.965501552500.530.59650.60521.072.03×1030.1700.015000No CU-AC
140.62000.967501602500.560.60320.64151.0243.54×1030.2000.050004×104No CU-AC

Note: Fitting parameters correspond to models introduced in the text. q = deviatoric stress; Δuw = cyclic pressure amplitude; pmax = maximum mean stress at the end of depressurization cycle; ηmax = maximum stress obliquity (=q/pmin); eo = initial void ratio at i=0; eT = terminal void ratio at i; m = model parameter; N* = characteristic number; γ1 = shear strain at the end of first cycle i=1; a, b, and c = model parameters; and d = ratcheting parameter.

a
Shear strain accumulation model: γi=γ1+a(1ib)c(1i1)+d(i1).
b
Void ratio evolution model: ei=eT+(eoeT)[1+(i/N*)m]1.

Experimental Results

This section reports detailed experimental results for two sets of tests designed to explore the effects of maximum stress obliquity ηmax and initial confinement po. We analyze the complete dataset gathered in this study in the following section.

Study 1: Maximum Stress Obliquity ηmax

Fig. 4 illustrates the load-deformation response of loose and medium dense sands initially loaded to the same p=250  kPa and ηmin=0.20, and subjected to repetitive fluid pressure cycles to different maximum stress obliquities ηmax=0.33, 0.40, 0.45, and 0.50 (Fig. 3). The results show
Pre-loading. The void ratio decreases during isotropic confinement (p=100  kPa, q=0) and deviatoric loading (p=150  kPa, q=50  kPa). The vertical strain is very similar in all specimens as the stress obliquity reaches the initial value of ηo=0.33 and during the first decrease in pore water pressure to reach ηmin=0.20.
Repetitive pressure cycles. All specimens exhibit volume dilation every time the pore pressure increases (the mean stress decreases, and obliquity increases from ηminηmax); however, there is residual contraction at the end of the cycle. The vertical strain increases during pressurization (ηminηmax) and accumulates at the end of every cycle. Volumetric contraction and vertical strain accumulation are more pronounced in specimens that reach a higher maximum stress obliquity ηmax during pressure cycles. Note that the initial void ratio eo of all specimens falls in the contractive zone just before repetitive loading; thereafter, the two specimens subjected to large pressure cycles (ηmax=0.50 and ηmax=0.45) become denser than at the critical state.
Undrained shear. All specimens reach the critical state line during the undrained deviatoric loading that followed the N=100 pressure cycles (p-q-e space in Fig. 4). These results confirm that in the absence of overt localization the critical state line is not affected by the monotonic or cyclic loading history (Taylor 1948; Schofield and Wroth 1968; Castro et al. 1982; Mohamad and Dobry 1986).
Fig. 4. Maximum stress obliquity: Loose and medium dense sands subjected to repetitive fluid pressure cycles to different maximum stress obliquities ηmax. In all four specimens, the pressure cycles begin at po=250  kPa and ηmin=0.20. Tests end with undrained axial compression from the same initial stress condition at ηmin=0.20. Notation: p=(σ1+σ3)/2, q=(σ1σ3)/2, ϕcs=sin1(tanα), and stress obliquity η=q/p. Numbers in square brackets [#] indicate the Test number in Table 2.

Study 2: Confining Effective Stress p

Fig. 5 shows the p-q-e-εz load-deformation response of three medium dense specimens subjected to different initial mean stress values po. Initial conditions include specimens above and below the critical state line. Details of the loading history before repetitive loading is shown in Fig. 3. Pressure cycles cause changes in obliquity from ηmin=0.20 to ηmax=0.50 in all cases. The changes in void ratio and the vertical strain accumulation during the repetitive pressure cycles are more significant in the one specimen subjected to high initial mean stress po. Once again, all specimens shear and dilate as the pressure increases. However the overall void ratio trend is contractive at the end of every cycle. All three specimens land on the dilative side of the critical state at the end of cyclic loading and exhibit a dilative tendency during the final undrained shear.
Fig. 5. Confining effective stress: Medium dense sand specimens subjected to repetitive fluid pressure cycles between ηmin=0.20 and ηmax=0.50. Tests end with undrained axial compression from the same obliquity ηmin=0.20. Figure 2 shows all stress paths in detail. Notation: p=(σ1+σ3)/2, q=(σ1σ3)/2, ϕcs=sin1(tanα), and stress obliquity η=q/p. Numbers in square brackets [#] indicate the Test number in Table 2.

Analysis of the Complete Dataset

This section analyzes the results of all tests conducted in this study (Table 2), with an emphasis on the shear strains and volume changes that occur during repetitive pressure cycles. Within triaxial boundary conditions, the shear strain γ=(3εzεvol)/2 combines the vertical strain εz and the volumetric strain εvol. System compliance and inadequate saturation bias both the measured peak-to-peak volumetric strain and the computed peak-to-peak shear strain. Therefore, figures and analyses in this section place emphasis on incremental and cumulative strains determined at the same pressure at the end of each cycle.

Shear Deformation

Fig. 6 presents the shear strain accumulation as a function of pressure cycles. The initial mean stress is the same for all specimens, po=250  kPa, but pressure cycles reach different maximum stress obliquities ηmax. The shear strain accumulation model below fits data trends in all tests (modified from Chong and Santamarina 2016)
γi=γ1+a(1ib)c(1i1)+d(i1)
(1)
where a, b, c, and d are fitting parameters, and i is the number of loading cycles. The shakedown response corresponds to d=0, while d>0 implies ratcheting. Table 2 summarizes the fitted model parameters for all tests. Results indicate that
The shear strain accumulation induced by pressure cycles is more pronounced in earlier cycles, in loose sands, in specimens that experience a higher maximum stress obliquity ηmax (for tests with the same initial po), and in specimens subjected to a higher initial mean stress po (for tests that reach the same ηmax).
Shakedown is unmistakable for specimens with small ηmax. In general, all specimens subjected to stress obliquity ηmax0.50 exhibit a shakedown response regardless of their initial density. For reference, the obliquity at critical state is ηcs=0.52.
The dense specimen subjected to pressure cycles above the critical state (ηmax=0.56) shows a ratcheting response (d=4×104). This specimen gradually dilates during pressure cycles. Hence the frictional resistance evolves from ϕpeak towards ϕcs; eventually, a pressure cycle above critical state obliquity will cause the soil to fail.
Overall, the initial packing density determines the different failure modes when soils are subjected to pressure fluctuations. Loose soil will contract. Dense soil will experience dilation when pressure cycles reach high stress obliquity (above values corresponding to ϕcs).
Fig. 6. Shear deformation: Cumulative shear strain γ versus number of cycles: (a) loose and medium dense specimens; and (b) dense specimens. The initial mean effective stress po=250  kPa and minimum stress obliquity is ηmin=0.20 in all tests. The maximum stress obliquity ηmax is indicated in each case. Dotted lines: Shear strain accumulation model fitted to test results [Eq. (1), Table 2 summarizes model parameters]. The obliquity at critical state is ηcs=0.52. Numbers in square brackets [#] indicate the Test number in Table 2.
These observations indicate that shear strain accumulation is a function of the initial void ratio eo, the initial confinement po, and the maximum stress obliquity ηmax reached during pressure cycles.

Volume Change

Void Ratio

Fig. 7 presents the evolution of the void ratio with the number of cycles for all specimens where pressure cycles start at po=250  kPa. Specimens in Fig. 7 have distinct initial void ratios eo (from eo=0.59 to eo=0.71; for reference, emin=0.533 and emax=0.786) and reach different maximum stress obliquity values ηmax. The highest rate of change in void ratio occurs during earlier pressure cycles and is more pronounced as the maximum stress obliquity increases. The void ratio ei measured at the end of the ith cycle evolves towards an asymptotic terminal void ratio eT in all specimens. The following accumulation model properly fits all datasets (Park and Santamarina 2019):
ei=eT+(e0eT)[1+(iN*)m]1for  m>0
(2)
where the m-exponent varies between m=0.8 to 1.0. The model parameter N* is the number of cycles required for a given specimen to reach half of the asymptotic volume change (eoeT)/2. Table 2 lists fitted model parameters for all tests.
Fig. 7. Volume change: Void ratio versus number of cycles for loose, medium, and dense specimens subjected to fluid pressure oscillations. The initial mean effective stress po=250  kPa and minimum stress obliquity is ηmin=0.20 is common to all tests. The maximum stress obliquity ηmax is indicated in each case. Dotted lines: void ratio evolution model fitted to test results [Eq. (2), Table 2 summarizes model parameters]. The obliquity at critical state is ηcs=0.52. Numbers in square brackets [#] indicate the Test number in Table 2.

Discussion

Particle-Scale Deformation Mechanisms: Threshold Strain

In the absence of grain crushing, particle-scale deformation mechanisms relate to the strain level γ the soil experiences. There are two threshold strains under monotonic loading conditions
1.
all deformations take place at contacts until the elastic threshold strain γγth|el that is selected at G/Gmax0.99, and
2.
there are minimal fabric changes until the volumetric threshold strain γγth|v. Typically, γth|v30·γth|el (Sands: Vucetic 1994, Ishihara 1996. Clays: Díaz-Rodríguez and Santamarina 2001). Slip-down, grain roll-over, and high frictional losses take place at strains above the volumetric threshold (Ishihara 1996; Mueth et al. 2000).
Let’s consider three spherical particles arranged in a triangular configuration and subjected to a normal force N [Fig. 8(a), inset]. The shear force T increases until the contact force F13 between particles ① and ③ becomes F13=0, which indicates contact loss. The extension of the 1 and 3 contact and the contraction of the 2 and 3 contact follow Hertzian behavior. Then, the horizontal displacement δ* of the top particle ③ relative to the interlayer height d·cos30° yields the equivalent shear strain for contact loss γth|loss as a function of the mineral shear modulus Gg and the applied confining stress σ estimated from the applied force as σN/d2 (Santamarina et al. 2001)
γth|loss=1.3(σGg)2/3
(3)
Fig. 8. Strain and obliquity thresholds: (a) incremental volumetric strain versus peak-to-peak vertical strain; and (b) incremental shear strain as a function of maximum stress obliquity at the i=100 pressure cycle. Numbers in square brackets [#] indicate the Test number in Table 2.
This analysis anticipates that the threshold strain at contact loss increases with confining stress σ in agreement with experimental evidence (Dyvik et al. 1984; Kim et al. 1991; Vucetic 1994). The threshold strain estimated using Eq. (3) is γth|loss5×104 at po=250  kPa (Table 1; see data in Silver and Seed 1971; Dobry et al. 1982; Vucetic 1994; Santamarina and Shin 2009).
Clearly, there can be no volumetric strain accumulation when the cyclic strain level is too low for contact loss and fabric change. But, what is the threshold strain for repetitive pressure cycles? Let us compute the incremental volumetric strain in a given cycle Δεvol|i as a function of the change in void ratio between two consecutive cycles i and i+1 (taken at the same fluid pressure at the end of each cycle)
Δεvol|i=eiei+11+ei
(4)
Fig. 8(a) shows the absolute value of the incremental volumetric strain Δεvol|i for contractive and dilative specimens plotted against the peak-to-peak vertical strain εzpp for all cycles. Data trends show that (1) volumetric changes diminish as the number of pressure cycles increases, and (2) volumetric changes vanish Δεvol|i0 as the peak-to-peak vertical strain εzpp2-to5×104.

Shakedown or Ratcheting?

The initial state of stress (po, qo) and void ratio eo together with the amplitude of pressure cycles Δuw and the maximum stress obliquity ηmax determine the shear strain response of a soil subjected to repetitive changes in pore water pressure under constant deviatoric stress. The incremental shear strain Δγi between two consecutive cycles i and i+1 scales with the maximum stress obliquity when ηmax<0.95·ηcs, and gradually diminishes towards shakedown [Figs. 6(a) and 8(b)]. Ratcheting takes place when the maximum stress obliquity approaches or exceeds ηmaxηcs [Figs. 6(b) and 8(b)]. Note that Wu et al. 2017 report the onset of ratcheting behavior at η=0.50, i.e., close to failure.

Minimum Volumetric Strain

The volumetric strain εvol=Δu/Bmax computed using the small-strain maximum skeletal bulk stiffness Bmax provides a lower bound estimate of the volumetric strain the soil will experience during a given pressure cycle Δuw. The maximum skeletal bulk stiffness can be computed from the in situ shear wave velocity Bmax=2·(Vs2ρ)(1+ν)/[3·(12ν)], where ν is the small-strain Poisson’s ratio. For example, consider a KAUST 20/30 specimen subjected to po=250  kPa and Δuw=100  kPa where the shear wave velocity for KAUST 20/30 sand increases with confining stress as Vs=89  m/s(po/1  kPa)0.21 and the small-strain Poisson’s ratio is ν0.15 (Note eo0.65 in—Table 1). Then, the minimum peak-to-peak volumetric strain is εvol6×104.

Maximum Volumetric Strain

Terminal Void Ratio

Fig. 9(a) compares the initial void ratio eo and the terminal void ratio eT for specimens with different eo, po, and ηmax (Note: po=250  kPa for the eight specimens in the dotted box, but symbols are p-shifted to facilitate the visualization). Previous studies have suggested that there is a characteristic “terminal void ratio” for each loading condition (Narsilio and Santamarina 2008). Note that the critical state CS is the terminal state for monotonic shear. Results reported in this study show that loose to medium dense specimens contract to reach terminal void ratios that are denser than CS. However, very dense specimens only dilate if pressurization causes high stress obliquity ηmax, and may rapidly evolve to failure without reaching a unique terminal state.
Fig. 9. Asymptotic volumetric response: (a) evolution of void ratio for medium dense (= blue) and dense sand (= red) specimens subjected to fluid pressure oscillations at different mean effective stress po. Empty symbols show the initial void ratio eo at the beginning of repetitive pressure cycles, while filled symbols show the terminal void ratio eT. The repetitive changes in pore water pressure begin at ηmin=0.20 in all specimens shown in this figure. The critical state line CSL is ecs=0.8450.074log(p); and (b) normalized volume change (eoeT)/(eoemin) caused by fluid pressure oscillations versus maximum stress obliquity ηmax—Loose, medium dense, and dense sand specimens. Numbers in square brackets [#] indicate the Test number in Table 2.

Potential Volume Change: Obliquity

Let us define the normalized asymptotic volume change (eoeT)/(eoemin) in terms of the initial void ratio eo at the beginning of pressure cycles (i=0), the terminal void ratio eT(), and the minimum void ratio emin. Results discussed above suggest that the normalized volume change caused by fluid pressure cycles depends on the maximum stress obliquity ηmax [Fig. 9(b)]. Contractive specimens experience volume change when obliquity exceeds ηmax>0.3, and it is proportional to ηmax thereafter. On the other hand, significant volumetric dilation in dense specimens requires a stress obliquity ηmax greater than the critical state stress obliquity ηmaxηcs=0.52. The minimum void ratio emin, the void ratio at critical state ecs, and the terminal void ratio for pressure cycles eT are all “asymptotic states” for a given sand (where ecs and eT are initial stress dependent). The preceding results show that terminal void ratios fall below the critical state line between emin<eT<ecs. Together, Figs. 79 suggest that the balance between internal deformation mechanisms depends on initial stress conditions po and qo, the maximum obliquity ηmax reached in pressure cycles and the initial void ratio eo.

Design Guidelines

The volumetric strain εT associated with the maximum asymptotic change in void ratio Δe=eoeT induced by pressure cycles as i is
εT=e0eT1+e0
(5)
We cannot propose a definitive approach to estimate the terminal volumetric strain εT for pressure cycles due to the limited dataset available at this time. However, the results in Fig. 9(b) suggest
The terminal change in void ratio for loose and medium dense sands is a μ-fraction of the void ratio difference eoemin. In other words, ΔeT=eoeT=μ·(eoemin).
The μ-fraction is relatively low (i.e., μ0.3) and is a function of the maximum stress obliquity ηmax.

Comparison between Pressure Cycles versus Ko-Loading Cycles

The terminal void ratio evolves to its asymptotic state when the sand is subjected to repetitive vertical loading under zero lateral strain (previously reported in Park and Santamarina 2019). While boundary conditions are very different, both studies show that
There is a minimum strain required for plastic strain accumulation. The vertical threshold strain in the Ko cell varies in the range of 2 to 7×104, which is similar to estimated values in this study.
All specimens contract in Ko-loading cycles, but not in the pore-water pressure cycles with deviatoric loads (Fig. 7). Yet, the terminal void ratio falls between eo>eT>(0.7·eo+0.3·emin) in both Ko-loading and pressure cycle studies.

Ratio between Horizontal-to-Vertical Plastic Strains

The shear strain accumulation model γi [Eq. (1)] and the void ratio evolution model ei [Eq. (2)] allow us to compute the incremental plastic vertical strain Δεzpl and plastic volumetric strain Δεvolpl between two consecutive cycles i and i+1. This approach avoids the inherent error magnification in incremental computations using experimental data
Δεzpl|i=εzpl|i+1εzpl|i
(6)
Δεvolpl|i=ei+1ei1+ei
(7)
For small strains, the ratio ν* between the incremental horizontal-to-vertical plastic strains in axisymmetric conditions is
ν*|i=ΔεΔε//|plastic=Δεvolpl|iΔεzpl|i2Δεzpl|i=12(1Δεvolpl|iΔεzpl|i)
(8)
A ratio ν*=0.5 implies vertical deformation at constant volume (i.e., accumulation of vertical deformation at the terminal density). A ratio ν*0 corresponds to volume contraction under zero-lateral strain. A negative ratio ν*<0 indicates that both vertical and horizontal contraction take place during repetitive loading; in fact, ν*=1 implies isotropic volume contraction. Finally, a positive ratio ν*>0 indicates Δεvolpl<Δεzpl.
Fig. 10 shows the evolution of the plastic strain ratio ν* with the number of cycles for loose, medium dense, and dense specimens. All trends exhibit an early dip into lower values of the plastic strain ratio (i.e., towards global contraction), followed by a gradual evolution to asymptotic trends.
Fig. 10. Ratio between incremental horizontal-to-vertical plastic strains ν*=Δεpl/Δε//pl versus number of cycles. Sketches capture deformation modes. Numbers in square brackets [#] indicate the Test number in Table 2.

Conclusions

Repetitive changes in pore water pressure can lead to the accumulation of plastic volumetric and shear strains. The initial state of stress and void ratio (po, qo, eo), the amplitude of pressure cycles Δuw, and the maximum stress obliquity ηmax determine the volumetric and shear strain response.

Volumetric Response

The void ratio evolves towards an asymptotic terminal void ratio eT as the number of pressure cycles increases; the rate of change is more pronounced for high stress obliquity ηmax.
The terminal void ratio for pressure cycles eT falls below the critical state line. The void ratio at critical state ecs for the same initial stress po and the minimum void ratio emin of the sand “bound” the terminal void ratio for pressure cycles emin<eT<ecs.
The terminal void ratios for dilative and contractive specimens do not converge to a single trend.
The terminal change in the void ratio (eoeT) in loose and medium dense sands increases with stress obliquity ηmax and is a fraction of (eoemin); for reference, (eoeT)0.3·(eoemin) in this study.
Dense dilative sands experience minimal void ratio changes and only dilate when ηmax approaches the critical state, ηmax0.95·ηcs. Consequently, the frictional resistance evolves from ϕpeak towards ϕcs and soils may fail in shear during subsequent pressure cycles.

Shear Response

The shear strain accumulation is more pronounced in earlier cycles, in loose sands, in specimens subjected to higher initial mean stress po and in specimens that experience a higher maximum stress obliquity ηmax.
The shear deformation may stabilize at shakedown, or continue in ratcheting mode. The maximum stress obliquity ηmax is the best predictor of shakedown or ratcheting.
Shakedown should be expected as long as pressure amplitudes keep the stress obliquity below ηmax<0.95·ηcs. Conversely, ratcheting takes place when the maximum stress obliquity approaches or exceeds ηmax0.95·ηcs.
Volumetric and shear strain accumulation during repetitive pressure cycles requires a minimum threshold strain which is estimated to be γ5×104 in this study. A particle-level analysis of contact loss and published experimental data show that the threshold strain increases with confinement p.

Acknowledgments

The KAUST endowment funded this research. Gabrielle E. Abelskamp edited the manuscript.

References

Abuel-Naga, H. M., D. T. Bergado, and A. Bouazza. 2007. “Thermally induced volume change and excess pore water pressure of soft Bangkok clay.” Eng. Geol. 89 (1–2): 144–154. https://doi.org/10.1016/j.enggeo.2006.10.002.
Albrecht, B. A., and C. H. Benson. 2001. “Effect of desiccation on compacted natural clays.” J. Geotech. Geoenviron. Eng. 127 (1): 67–75. https://doi.org/10.1061/(ASCE)1090-0241(2001)127:1(67).
Alonso-Marroquin, F., and H. J. Herrmann. 2004. “Ratcheting of granular materials.” Phys. Rev. Lett. 92 (5): 054301. https://doi.org/10.1103/PhysRevLett.92.054301.
Anderson, S. A., and N. Sitar. 1995. “Analysis of rainfall-induced debris flows.” J. Geotech. Eng. 121 (7): 544–552. https://doi.org/10.1061/(ASCE)0733-9410(1995)121:7(544).
Brenner, R. P. 1985. “Field stress path simulation of rain-induced slope failure.” In Vol. 2 of Proc. 11th ICSMFE, 991–996. Rotterdam, Netherlands: A.A. Balkema.
Cai, Y., T. Wu, L. Guo, and J. Wang. 2018. “Stiffness degradation and plastic strain accumulation of clay under cyclic load with principal stress rotation and deviatoric stress variation.” J. Geotech. Geoenviron. Eng. 144 (5): 04018021. https://doi.org/10.1061/(ASCE)GT.1943-5606.0001854.
Castro G., S. J. Poulos, J. W. France, and J. L. Enos. 1982. Liquefaction induced by cyclic loading. Winchester, MA: Geotechnical Engineers Inc.
Chamberlain, E. J., I. Iskandar, and S. E. Hunsicker. 1990. “Effect of freeze-thaw cycles on the permeability and macrostructure of soils.” Cold Region Res. Eng. Lab. 90 (1): 145–155.
Chang, W. J., S. H. Shou, and A. B. Huang. 2017. “Mechanism of subsidence from pore pressure fluctuation in aquifer layers.” In Proc. 19th Int. Conf. Soil Mechanics and Geotechnical Engineering. Seoul: International Society for Soil Mechanics and Geotechnical Engineering.
Chod, G. C., J. Dodds, and J. C. Santamarina. 2006. “Particle shape effects on packing density, stiffness, and strength: Natural and crushed sands.” J. Geotech. Geoenviron. Eng. 132 (5): 591–602. https://doi.org/10.1061/(ASCE)1090-0241(2006)132:5(591).
Chong, S. H., and J. C. Santamarina. 2016. “Sands subjected to vertical repetitive loading under zero lateral strain: Accumulation models, terminal densities, and settlement.” Can. Geotech. J. 53 (12): 2039–2046. https://doi.org/10.1139/cgj-2016-0032.
Chu, J., S. Leroueil, and W. K. Leong. 2003. “Unstable behaviour of sand and its implication for slope instability.” Can. Geotech. J. 40 (5): 873–885. https://doi.org/10.1139/t03-039.
Croney, D., and J. D. Coleman. 1954. “Soil structure in relation to soil suction (pF).” Eur. J. Soil Sci. 5 (1): 75–84. https://doi.org/10.1111/j.1365-2389.1954.tb02177.x.
da Fonseca, A. V., S. Rios, M. F. Amaral, and F. Panico. 2013. “Fatigue cyclic tests on artificially cemented soil.” Geotech. Test. J. 36 (2): 227–235. https://doi.org/10.1520/GTJ20120113.
Díaz-Rodríguez, J. A., and J. C. Santamarina. 2001. “Mexico City soil behavior at different strains: Observations and physical interpretation.” J. Geotech. Geoenviron. Eng. 127 (9): 783–789. https://doi.org/10.1061/(ASCE)1090-0241(2001)127:9(783).
Dobry, R., R. S. Ladd, F. Y. Yokel, R. M. Chung, and D. Powell. 1982. “Prediction of pore water pressure buildup and liquefaction of sands during earthquakes by the cyclic strain method.” In Building science series 138. Washington, DC: US Department of Commerce, National Bureau of Standards.
Dyvik, R., R. Dobry, G. E. Thomas, and W. G. Pierce. 1984. Influence of consolidation shear stresses and relative density on threshold strain and pore pressure during cyclic straining of saturated sand. Washington, DC: Dept. of the Army, USACE.
Farooq, K., R. Orense, and I. Towhata. 2004. “Response of unsaturated sandy soils under constant shear stress drained condition.” Soils. Found. 44 (2): 1–13. https://doi.org/10.3208/sandf.44.2_1.
Gambolati, G., and P. Teatini. 2015. “Geomechanics of subsurface water withdrawal and injection.” Water Resour. Res. 51 (6): 3922–3955. https://doi.org/10.1002/2014WR016841.
Gu, C., Z. Gu, Y. Cai, J. Wang, and Q. Dong. 2018. “Effects of cyclic intermediate principal stress on the deformation of saturated clay.” J. Geotech. Geoenviron. Eng. 144 (8): 04018052. https://doi.org/10.1061/(ASCE)GT.1943-5606.0001924.
Huang, A. W. 2016. “The seventh James. K. Mitchell lecture: Characterization of silty/sand soils.” Aust. Geomech. J. 51 (4): 1–23.
Hung, W. C., C. Hwang, J. C. Liou, Y. S. Lin, and H. L. Yang. 2012. “Modeling aquifer-system compaction and predicting land subsidence in central Taiwan.” Eng. Geol. 147–148 (Oct): 78–90. https://doi.org/10.1016/j.enggeo.2012.07.018.
Ishihara, K. 1996. Soil behaviour in earthquake geotechnics. Oxford, New York: Clarendon Press; Oxford University Press.
Johnson, K. L. 1986. “Plastic flow, residual stress and shakedown in rolling contact.” In Proc. 2nd Int. Conf. on Contact Mechanics and Wear of Rail/Wheel Systems. Waterloo, ON: Univ. of Rhode Island.
Kim, D. S. 1991. “Deformational characteristics of soils at small to intermediate strains from cyclic test.” Ph.D. thesis, Dept. of Civil Architectural and Environmental Engineering, Univ. of Texas.
Laloui, L. 2001. “Thermo-mechanical behaviour of soils.” Rev. Française Génie Civ. 5 (6): 809–843. https://doi.org/10.1080/12795119.2001.9692328.
Leroueil, S., J. Chu, and D. Wanatowski. 2009. “Slope instability due to pore water pressure increase.” In Proc., 1st Italian Workshop on Landslides, 8–10. Naples, Italy: Doppiavoce.
Mohamad, R., and R. Dobry. 1986. “Undrained monotonic and cyclic triaxial strength of sand.” J. Geotech. Eng. 112 (10): 941–958. https://doi.org/10.1061/(ASCE)0733-9410(1986)112:10(941).
Mueth, D. M., G. F. Debregeas, G. S. Karczmart, P. J. Eng, S. R. Nagel, and H. M. Jaeger. 2000. “Signatures of granular microstructure in dense shear flows.” Nature 406 (6794): 385–389. https://doi.org/10.1038/35019032.
Musso, G., E. R. Morales, A. Gens, and E. Castellanos. 2003. “The role of structure in the chemically induced deformations of FEBEX bentonite.” Appl. Clay Sci. 23 (1–4): 229–237. https://doi.org/10.1016/S0169-1317(03)00107-8.
Nakata, Y., T. Kajiwara, and N. Yoshimoto. 2013. “Collapse behavior of slope due to change in pore water pressure.” In Proc., 18th Int. Conf. Soil Mechanics and Geotechnical Engineering, 2225–2228. Paris: International Society for Soil Mechanics and Geotechnical Engineering.
Narsilio, A., and J. C. Santamarina. 2008. “Terminal densities.” Géotechniqué 58 (8): 669–674.
Ohtsuka, S. 2007. “Shear behavior of clay in slope for pore water pressure increase.” In Progress in landslide science, 295–303. Berlin: Springer.
Ohtsuka, S., and Y. Miyata. 2001. “Consideration on landslide mechanism based on pore water pressure loading test.” In Vol. 2 of Proc., Int. Conf. Soil Mechanics and Geotechnical Engineering, 1233–1236. London: International Society for Soil Mechanics and Geotechnical Engineering.
Olson, S. M., T. D. Stark, W. H. Walton, and G. Castro. 2000. “1907 static liquefaction flow failure of the north dike of Wachusett dam.” J. Geotech. Geoenviron. Eng. 126 (12): 1184–1193. https://doi.org/10.1061/(ASCE)1090-0241(2000)126:12(1184).
Ono, T., and T. Mitachi. 1997. “Computer controlled triaxial freeze-thaw-shear apparatus.” In Ground freezing, 335–339. Rotterdam: A. A. Balkema.
O’Reilly, M. P., and S. F. Brown. 1991. Cyclic loading of soils: From theory to design. Glasgow, UK: Blackie.
Orense, R., K. Farooq, and I. Towhata. 2004. “Deformation behavior of sandy slopes during rainwater infiltration.” Soils. Found. 44 (2): 15–30. https://doi.org/10.3208/sandf.44.2_15.
Page, D., P. Dillon, J. Vanderzalm, S. Toze, J. Sidhu, K. Barry, K. Levett, S. Kremer, and R. Regel. 2010. “Risk assessment of aquifer storage transfer and recovery with urban stormwater for producing water of a potable quality.” J. Environ. Qual. 39 (6): 2029–2039. https://doi.org/10.2134/jeq2010.0078.
Park, J., and J. C. Santamarina. 2019. “Sand response to a large number of loading cycles under zero-lateral strain conditions: Evolution of void ratio and small strain stiffness.” Géotechniqué 69 (6): 501–513. https://doi.org/10.1680/jgeot.17.P.124.
Premchitt, J., E. W. Brand, and H. B. Phillipson. 1986. “Landslides caused by rapid groundwater changes.” Geol. Soc. London, Eng. Geol. Spec. Publ. 3 (1): 87–94. https://doi.org/10.1144/GSL.ENG.1986.002.01.09.
Qi, J., P. A. Vermeer, and G. Cheng. 2006. “A review of the influence of freeze-thaw cycles on soil geotechnical properties.” Permafrost Periglacial Processes 17 (3): 245–252. https://doi.org/10.1002/ppp.559.
Santamarina, J. C., and G. C. Cho. 2001. “Determination of critical state parameters in sandy soils-simple procedure.” Geotech. Test. J. 24 (2): 185–192. https://doi.org/10.1520/GTJ11338J.
Santamarina, J. C., K. A. Klein, and M. A. Fam. 2001. Soils and waves: Particulate materials behavior, characterization and process monitoring. Chichester, UK: Wiley.
Santamarina, J. C., and H. Shin. 2009. “Friction in granular media.” In Meso-scale shear physics in earthquake and landslide mechanics, 157–188. London: CRC Press.
Schofield, A. N., and P. Wroth. 1968. Critical state soil mechanics. London: McGraw-Hill.
Sharp, R. W., and J. R. Booker. 1984. “Shakedown of pavements under moving surface loads.” J. Transp. Eng. 110 (1): 1–14. https://doi.org/10.1061/(ASCE)0733-947X(1984)110:1(1).
Shi, X., S. Jiang, H. Xu, F. Jiang, Z. He, and J. Wu. 2016. “The effects of artificial recharge of groundwater on controlling land subsidence and its influence on groundwater quality and aquifer energy storage in Shanghai, China.” Environ. Earth. Sci. 75 (3): 195. https://doi.org/10.1007/s12665-015-5019-x.
Silver, M. L., and H. B. Seed. 1971. “Volume changes in sands during cyclic loading.” J. Soil Mech. Found. Div. 97 (9): 1171–1182.
Skempton, A. W. 1961. “Effective stress in soils, concrete and rocks.” In Proc., Pore Pressure and Suction in Soils Conf. 4–16. London: Butterworths.
Taylor, D. W. 1948. Fundamentals of soil mechanics. New York: Wiley.
Triantafyllidis, T., T. Wichtmann, and A. Niemunis. 2004. “On the determination of cyclic strain history.” In Proc., CBS04, 321–332. Rotterdam, Netherlands: A.A. Balkema.
Viklander, P. 1998. “Permeability and volume changes in till due to cyclic freeze/thaw.” Can. Geotech. J. 35 (3): 471–477. https://doi.org/10.1139/t98-015.
Vucetic, M. 1994. “Cyclic threshold shear strains in soils.” J. Geotech. Geoenviron. Eng. 120 (12): 2208–2228. https://doi.org/10.1061/(ASCE)0733-9410(1994)120:12(2208).
Werkmeister, S. 2003. “Permanent deformation behaviour of unbound granular materials in pavement constructions.” Ph.D. thesis, Fakultät Bauingenieurwesen, Technischen Universita¨t.
Werkmeister, S., A. R. Dawson, and F. Wellner. 2005. “Permanent deformation behaviour of granular materials.” Road. Mater. Pavement. 6 (1): 31–51. https://doi.org/10.1080/14680629.2005.9689998.
Wichtmann, T., A. Niemunis, and T. Triantafyllidis. 2005. “Strain accumulation in sand due to cyclic loading: Drained triaxial tests.” Soil Dyn. Earthquake Eng. 25 (12): 967–979. https://doi.org/10.1016/j.soildyn.2005.02.022.
Wu, T., Y. Cai, L. Guo, D. Ling, and J. Wang. 2017. “Influence of shear stress level on cyclic deformation behaviour of intact Wenzhou soft clay under traffic loading.” Eng. Geol. 228 (Oct): 61–70. https://doi.org/10.1016/j.enggeo.2017.06.013.
Youd, T. L. 1973. “Factors controlling maximum and minimum densities of sands.” In Evaluation of relative density and its role in geotechnical projects involving cohesionless soils, edited by E. Selig, and R. Ladd. 98–112. West Conshohocken, PA: ASTM.

Information & Authors

Information

Published In

Go to Journal of Geotechnical and Geoenvironmental Engineering
Journal of Geotechnical and Geoenvironmental Engineering
Volume 146Issue 5May 2020

History

Received: Jul 8, 2018
Accepted: Oct 30, 2019
Published online: Mar 10, 2020
Published in print: May 1, 2020
Discussion open until: Aug 10, 2020

Authors

Affiliations

Postdoctoral Fellow, Earth Science and Engineering, King Abdullah Univ. of Science and Technology, Thuwal 23955-6900, Saudi Arabia (corresponding author). ORCID: https://orcid.org/0000-0001-7033-4653. Email: [email protected]
J. Carlos Santamarina, A.M.ASCE
Professor, Earth Science and Engineering, King Abdullah Univ. of Science and Technology, Thuwal 23955-6900, Saudi Arabia.

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share