Open access
Technical Papers
Apr 13, 2022

Extended TTS Model for Thermal and Mechanical Creep of Clay and Sand

Publication: Journal of Geotechnical and Geoenvironmental Engineering
Volume 148, Issue 6

Abstract

In geothermal engineering, the long-term thermomechanical responses of sand and clay are difficult to predict due to the lack of an essential understanding of thermally induced volume change. Recent experimental studies have measured thermal creep of dry sand, glass beads, and clays. However, this universal phenomena of thermal creep has not been properly included in constitutive models of soils. In this paper, the Tsinghua ThermoSoil (TTS) model, previously developed for clay, is extended to describe thermal and mechanical creep of sand. The elemental thermal creep and thermal cyclic behavior of Todi clay and Bangkok sand, reported under hydrostatic stress or oedometric conditions, are reproduced by the proposed formulation. These comparisons show the following key features of behavior described by the extended model (e-TTS): (1) thermal creep of clay can produce either compressive or s1welling behavior depending on the consolidation stress history, (2) the heating rate and temperature range have significant influences on thermal creep, but only small effects on swelling behavior, (3) thermal cycling induces compression of sand, which is dependent on effective stress level and relative density, and (4) the e-TTS model predicts stabilization of thermal creep strains for sands over within 50–100 thermal cycles due to equalization of elastic and locked-in hysteretic strains.

Introduction

Thermally induced volume change of clays and clay soils or shales has attracted much attention in applications ranging from geothermal resource development (Zymnis and Whittle 2020) to energy foundations (Brandl 2006), oil exploration (Hueckel and Borsetto 1990), energy storage (Laloui and Cekerevac 2003), and nuclear waste (Houston et al. 1985; Cui et al. 2009). In geothermal resource development, the continuous operation of ground source heat pump (GSHP) installations can result in significant long-term settlements, which could adversely affect the overlying structures or adjacent foundations (Zymnis and Whittle 2020). Similarly, thermal-induced strains in clays can cause stress changes in energy piles, reducing the safety associated with conventional design methods (Laloui and Di Donna 2011). In nuclear waste repository, thermal volume change of the boom clay affects stability of underground cravens (Cui et al. 2009). In applications to energy foundation in saturated sands, oil sands, sand and bentonite mixed buffering of nuclear waste repository, and backfilling of underground high-voltage electric cables, thermally induced volume change of sands has attracted attention (Hueckel and Borsetto 1990; Ng et al. 2016; Sittidumrong et al. 2019). In these applications, the soils undergo temperature fluctuations, resulting in unexpected monotonic and cyclic thermally induced volume changes. Better understanding of the complex thermal behavior of both clays and sands induced by daily or seasonal cyclic temperature changes is crucial for reliable analysis of these engineering problems.
Thermal volume change has been widely investigated in clays (Baldi et al. 1988; Burghignoli et al. 2000; Sultan et al. 2002; Laloui and Cekerevac 2003; Abuel-Naga et al. 2009; Di Donna and Laloui 2015), and to a lesser extent in sands (Demars and Charles 1982; Agar et al. 1986; Ng et al. 2016; Liu et al. 2018; Sittidumrong et al. 2019; Rotta Loria and Coulibaly 2021). Houston et al. (1985) studied the thermal consolidation of saturated marine clays through laboratory element tests, where excess pore pressures were generated by heating samples at constant water content and then allowed to dissipate. They observed thermal primary consolidation where volumetric strains occur due to the diffusion of excess pore pressures and thermal secondary consolidation (i.e., thermal creep) where strains occur without further change in effective stress. Thermal creep strains typically increased linearly with log time at rates controlled by the prevailing temperature.
The results of a large number of drained element tests show that normally consolidated (NC) clays exhibit thermal contraction under monotonic heating (Cekerevac and Laloui 2004), while heavily overconsolidated (OC) clays undergo thermal expansion (Sultan et al. 2002; Laloui and Cekerevac 2003; Abuel-Naga et al. 2009), and lightly OC soils show a transition from expansion to contraction when heated beyond a characteristic temperature (Baldi et al. 1988; Towhata et al. 1993; Sultan et al. 2002). Cui et al. (2009) show that thermal volumetric strains of OC Boom clay are dependent on the stress history, while creep strain rates increase significantly with temperature. A single cycle of temperature variation causes irreversible volume strains in clays (Di Donna and Laloui 2015). These thermal strains accumulate in each successive cycle but at a rate that decreases with the number of cycles.
Time-dependent compression behavior of clays is often characterized by the isotache framework (Šuklje 1957), which assumes a unique relation between effective stress, strain, and strain rate in 1D compression, shown as loci of constant strain rate in e-logp′ space, Fig. 1. This concept allows the creep of normally consolidated (NC) clay to be described quantitatively. The creep deformation at constant effective stress (p0; Fig. 1) corresponds to a decrease in strain rate of soil (path A to B′ in Fig. 1). Consolidation stress history, represented by swelling along the path A-C, causes a marked reduction in compressive creep rates at low OCR (C-D), while expansive/dilative creep strains occur at higher OCR (C-E-E) (Yao and Fang 2020).
Fig. 1. Unification of thermal and mechanical creep in 1D compression tests through isotache framework from Leroueil and Marques (1996).
The compressibility of clay is influenced by temperature at a given strain rate (Yashima et al. 1998). Leroueil and Marques (1996) show that isotache loci for normal consolidation are functions of strain rate and temperature. Increases in temperature cause additional compressive thermal creep strains for NC and lightly OC states (B-B and D=D; Fig. 1) and augment the swelling strains at higher OCR (E-E).
There is much more limited data on the thermal creep of sands or other granular materials. Chen et al. (2006) and Divoux et al. (2008) have reported the compression of columns of dry spherical beads due to thermal cycles. The data by Chen et al. (2006) show that the average accumulative compaction strain of granular packing reaches 5% after 30 heating and cooling cycles (T=107°C±2°C). The data by Divoux et al. (2008) show that the entire 1.7-m-high granular column compacts by 1.5 cm after 1,000 heating and cooling cycles (ΔT=10.8°C). The compaction of granular at each temperature decrement converges to a rate that varies linearly with log time. They resolve microscale deformations to show that the movements are independent of the dilation of the container and hence report that irreversible thermal creep is associated with local particle rearrangement.
More recently, Ng et al. (2016) observed the effect of initial relative density on the volumetric behavior of saturated Toyoura sand, heated and cooled (at a constant hydrostatic stress) in a triaxial apparatus, over a temperature range of T=23°C53°C. While the dense sand (Dr=70%) expanded linearly with the increase in temperature, specimens of loose sand (Dr=21%26%) showed maximum contractive strains (up to 0.15%) at T40°C and incremental expansion at higher temperatures. Pan et al. (2020) investigated that the 1D volumetric response of a saturated Foundry sand at very low confining stress. These tests show expansive strains for specimens with Dr=30%90% for heating from T=0°C40°C, while cooling back to the original temperature caused small net contractive strains (0.01%–0.05%). Sittidumrong et al. (2019) have measured the accumulation of cyclic thermal strains for saturated Bangkok sand (intact and reconstituted) for ΔT=22°C in a modified oedometer apparatus. These tests show always thermal contractive strains for all specimens after 100 thermal cycles, with the maximum contractive strains up to 0.6% at ΔT=22°C. Their results quantify the thermal strains as functions of the effective stress and relative density.
Several thermomechanical constitutive models have been proposed for soils assuming thermal hardening of the yield function to represent thermoplastic deformation of clay (Hueckel and Baldi 1990; Hueckel and Borsetto 1990; Cui et al. 2000; Laloui and Cekerevac 2003; Abuel-Naga et al. 2007; Zhang et al. 2012; Xiong et al. 2014). However, these thermomechanical constitutive models generally fail to reproduce the thermal and mechanical creep during heating and cooling phases as they neglect the viscous behavior of soils. Other models based on semiempirical or theoretical thermoviscoplasticity (TVP) were established to predict the combining effects of time and temperature (Laloui and Cekerevac 2003; Kurz et al. 2016; Qiao and Ding 2017). Di Donna and Laloui (2015) and Zhou et al. (2017) proposed constitutive models using more complex concepts for describing thermal cyclic stabilization. Zymnis et al. (2018) demonstrated that the Tsinghua ThermoSoil (TTS) model has the ability to predict the accumulation and stabilization of thermal strains for saturated clay. Zymnis and Whittle (2020) illustrate the geotechnical design optimization for an array of Borehole Heat Exchangers (BHEs) using the original TTS model. The original TTS model was developed for clay where thermal strains are related to the underlying physical mechanism of conversion of bound water to free water within the clay (Zhang and Cheng 2016, 2017).
Thermally-induced compression of sands and thermal strain rate dependency of granular media is not described by the existing TTS model or other aforementioned models. This paper describes an extended TTS (e-TTS) formulation that can present a unified model to explain both thermal strains in sands and clays. The conversion of bound water to free water is emphasized in clays during temperature elevation, while thermally-induced skeleton compaction is highlighted in cohesionless soils (sands, glass beads, etc). We illustrate model predictions through comparisons with existing laboratory experiments (Burghignoli et al. 2000; Sittidumrong et al. 2019) on two soils where there are measurements of both isothermal/mechanical creep and thermally induced creep strains.

Extended Tsinghua ThermoSoil (e-TTS) Model

The TTS model (Zhang and Cheng 2017) was developed to describe the thermo-hydro-mechanical coupling behavior of saturated soils based on a theory of nonequilibrium thermodynamics referred to as Granular Solid Hydrodynamics [(GSH) (Jiang and Liu 2003, 2007, 2009)]. The TTS model uses a hyperelastic relation to describe elasticity of granular media. Time-dependent inelastic behavior is attributable to the energy dissipation determined by Onsager’s reciprocal principle. A new intrinsic variable, granular entropy, replaces the plastic strain to describe the various irreversible granular-level processes of soils and to provide a more fundamental physical basis for characterizing the complex mechanical behavior of soils. The TTS model can describe major features of soil behavior including critical states, irreversible strains in undrained cyclic loading, and long-term thermal-mechanical response (Zhang and Cheng 2016, 2017). The analysis in this paper is based on an extension of the TTS formulation to unify the description of thermal creep, strain accumulation, and cyclic stabilization of clays and sands.
The key assumptions of the model can be summarized as follows:
1.
The saturated soil includes a solid phase, bound water phase, and free water phase. Bound water and free water are referred to as fluid phases, while the bound water is completely absorbed on the surface of the solid particles;
2.
It is assumed that soils are homogeneous and initially isotropic materials. All the phases are at the same temperature based on the local thermodynamic equilibrium postulate (LTE) and continuous in space; that is, for each point in space or each representative element of the volume (REV) in space;
3.
The fluid phase does not undergo solidification or vaporization. The bound water phase and the free water phase exhibit substance exchanges as temperature changes, while there is no mass exchange between the fluid phase and the solid phase.

Thermoelastic Behavior

In the e-TTS model, the stress-strain relationship for thermoelastic coupling is obtained from a function for the elastic potential energy density, ωe, to guarantee that the model is thermodynamically acceptable
πij=ωeεije
(1)
where εije = elastic strain tensor; and πij = elastic stress tensor.
Following Zheng and Cheng (2017), thermoelastic behavior of granular media (with Hertzian contact) can be represented by
ωe=B(εve)1.5[25(εve)2+ξ(εse)2]+3KeβTΔTdεve
(2)
where B=B0exp(B1ρd); B0, B1, and ξ = material parameters; ΔT=TT0 = temperature change; T0 = reference temperature; 3KeβTΔT = additional isotropic thermal loading to restrain the thermoelastic expansion induced by the temperature change; Ke = secant elastic bulk modulus of the solid skeleton, the formulation of which is shown in the Appendix; βT = linear thermal expansion coefficient for the soil skeleton; εve and εse=eijeeije = volumetric and 2nd invariant of deviator elastic strain tensor, respectively; and eij = components of the deviatoric strain tensor.
Zhang and Cheng (2017) established a generalized effective stress composition for saturated soils using the granular solid hydrodynamics approach
σij=πij+σijbw+α=s,f,bσijvα+σijg
(3)
where the first term on the right side of the effective stress equation is elastic stress, πij, which is the most important components of the effective stress; the partial stress σijbw = bound effective stress (due to disjoining pressures between bound water and clay particles); σijvα = viscous stress associated with dissipative flows of material viscosity; and σijg = granular effective stress, associated with kinetic energy fluctuation at the granular scale. Considering only the first term effective stress for saturated thermoelastic soils can then be expressed as
σijπij=1.4B(εve)2.5δij+1.5Bξ(εve)0.5(εse)2δij+3KeβT(TT0)δij+2Bξ(εve)1.5eije
(4)

Thermoviscoplastic Behavior

The total strain rate can be divided into the elastic strain rate and the inelastic strain rate
dtεij=dtεije+dtεijD
(5)
In classical plasticity theory, plastic strain is adopted to describe the inelastic strain for granular materials (where dtεijD represents the plastic strain rate). The variation of plastic strain is described by the plastic potential function. In the TTS model, the variation of inelastic strain is attributable to the energy dissipation according to GSH theory. The total energy dissipation rate R can be expressed as the product of dissipative forces and flows
R=Yijπij+IgTg+σijvsdtεij
(6)
where πij = dissipative forces corresponding to the effective stresses; Tg = granular temperature; dtεij = total strain rates; Yij = corresponding flows comprising the plastic strain rates; Ig = the rate of conversion of granular (local) entropy; and σijvs = viscous stresses of the soil skeleton.
The dissipative flows can be expressed as linear functions of the dissipative forces based on Onsager’s reciprocal principle. This well-known principle has been used to solve coupled fluid and heat flows in soil mechanics (Mitchell and Soga 2005). The TTS model describes incremental constitutive behavior as follows:
[YijIg]=[λijkl00γ][πijTg]
(7)
where λijkl and γ = migration coefficients.
It can be theoretically shown that the nonelastic strain rate (dtεijD) is the same as Yij (Zhang and Cheng 2017). Zhang and Cheng (2017) propose the following expression for the dissipative flow Yij (or plastic strain rate, dtεijD) to describe the transient elasticity of granular material
Yij=dtεijD=λsTga(eijeeijh)+λvTga(εkkeεkkh)δij
(8)
where εije, εijh = tensors of elastic and hysteretic strains tensors; and the parameter a controls rate dependence in material behavior.
The hysteretic strains are additional state parameters that improve model predictions of hysteretic stress-strain response under cyclic loading and enable the model to describe locked-in elastic strains (Coussy 1995; Collins 2005; Yan and Li 2011).

New Formulation for Granular Entropy

The e-TTS model accounts for interactions between the macrolevel and microlevel of particle or granular behavior by a double-entropy approach. In any irreversible process the total real entropy represents the irrecoverable deformations at the macro level, while the granular entropy is attributable to the energy dissipative mechanisms of soils at the microscopic level, such as sliding, rolling, and collision of particles, leading to a change in the kinetic energy and elastic potential energy. The energy evolution and dissipation at the microscopic level is similar to molecular motion and can be described through a granular entropy density sg (=bρdTg where b = constant; and ρd = dry mass density) and its conjugate variable, Tg granular temperature (Haff 1983; Jiang and Liu 2009). The e-TTS model assumes that granular entropy increases according to the following equation
Tgdtsg=RgIgTg
(9)
where Rg/Tg = granular entropy production rate; Tg can be understood as the dissipative force of the granular fluctuation; Ig = corresponding dissipative flow; and IgTg = energy dissipation rate induced by the granular fluctuation.
In addition, it should be noted that the reorganization of soil particles can be activated by mechanical and/or thermal loading, such that the strain rate and/or rate of temperature change can be understood as dissipative forces for granular fluctuation, the “dissipative flows” corresponding to dtεij and dtT are denoted as σijg and M, respectively. Hence, the granular entropy can be expressed as follows:
Rg=σijgdtεij+MdtT
(10)
with dissipation flow σijg
σijg=ηgTgdteij+ζgTgdtεkkδij
(11)
In order to describe the thermal-compaction of granular materials (sands, glass beads, etc.), the dissipative flow M in e-TTS model can be expressed as a linear function of the rate of temperature change following the Onsager’s reciprocal principle again, shown in [Eq. 12(a)].
In clays, this universal thermal-compaction mechanism is also present in addition to the conversion process from bound water to free water during temperature elevation, which is already modeled in the original TTS model (Zhang and Cheng 2016, 2017). So the dissipative flow M for clays has to be generalized into [Eq. 12(b)], in which the first term represents the effect of bound water to free water conversion and the second term represents the effect of thermal compaction mechanism of soil skeleton. This is similar to the concept used in the MIT-SR elasto-visco-plastic model for clay behavior (Yuan and Whittle 2018, 2021)
M=θtTgdtTfor  sand
(12a)
M=[ψgπkkαbfϕbw3ρd(1ϕ)+θtTg]dtTfor  clay
(12b)
where θt = migration coefficient for thermal-compaction of soil skeleton; ψg controls the direction of temperature change for clay: ψg>0 for heating and ψg=0 for cooling, πkk = mean effective stress; ϕbw(=Vbw/V) and ϕ = bound water porosity and total porosity, respectively; and αbf controls the conversion rate of bound water to free water. Zymnis et al. (2019) describe experimental methods to measure αbf.
Finally, the evolution of granular temperature can be derived as follows:
dtTg=ηg(dteij)2+ζg(dtεkk)2γTgbρd+[ψgπkkαbfϕbw3ρd(1ϕ)+θt](dtT)2
(13)
where ηg, ζg, γ, ψg, and θt = constant migration coefficients.
According to [Eq. 12(b)], there are two distinct mechanisms that relate changes in granular temperature to the temperature of the soil. The first involves the conversion from bound to free water, previously discussed by Zymnis et al. (2018) while the second, described by the migration coefficient, θt, describes thermally induced skeleton compaction within the granular materials. In addition, there is still one more thermoelastic related physical mechanism, characterized by the term of dtεkk in Eq. (13). If the temperature of soil is changed, this term becomes temperature dependent following Eqs. (4) and (8). This physical mechanism refers to the thermoelastic induced volumetric change, which is not purely thermal expansion/contraction any more according to the Eq. (8). This mechanism plays a less important role in complex thermal-mechanical coupled problems of both clays and sands, compared with the aforementioned two distinct mechanisms.
It is should be noted that there are other thermally-induced transport phenomena in soils including: (1) thermo-osmosis, and (2) thermal-filtration, which refers to the influence of a pressure gradient on heat flow (Mitchell and Soga 2005). These coupled processes have been discussed by Zhang and Cheng (2017), but are not defined within the constitutive law at the REV level in this paper.
The thermal creep or cyclic volumetric deformations are generally measured under 1D (oedometric) conditions, where there are two independent components of stress and strain. The Appendix summarizes the e-TTS model formulation for saturated soils in triaxial space. These relations define the effective stress-strain-temperature-time relations of soils at the elemental/point level and are used to simulate the thermal creep and cyclic thermomechanical response described in detail subsequently.
Within the e-TTS formulation, the inelastic strain is strain-rate dependent (Appendix) while the equations describing the evolution of granular temperature and inelastic strain are influenced by thermal effects [Eq. (13) in the Appendix]. Thus, the model simulates thermal creep behavior as a function of both strain rate and temperature.
Fig. 2 shows the solving procedure of thermal cyclic strain for the e-TTS model with a known rate of temperature (dtT), the current effective stress state and initial values of state variables. The rates of elastic strains, total strains, and mass conservation are updated using Eqs. (14), (16), and (19), and then the rates of granular temperature, hysteretic strains (locked entropy) are then updated using Eqs. (17) and (18). In the current implementation the key governing equation [granular entropy, Eq. (17)] is solved using Matlab (ode45 solver).
Fig. 2. Solution procedure of the e-TTS model under triaxial conditions with a known rate of temperature change.
There are 14 parameters (material constants) in the extended TTS model. The relationships between migration coefficients and model input parameters (b and m1,0 to m6) are presented in Table 1. These can be calibrated using conventional laboratory tests. Zymnis et al. (2018) illustrate the calibration procedure for clay, under the assumption of strain rate–independent behavior (a=0.5). In this work, the isothermal mechanical component parameters (B0, B1, ξ, a, h, m1, m2, m3, m4) can be calibrated by isotropic/1D compression and creep tests, while the thermal component parameters (βT, LT, αbf, m5, m6) are obtained from a thermal creep test. The initial elastic strains can be obtained from the initial consolidation stress state (p, q, e). We assume the initial locked strain is zero and initial granular temperature, Tg=104.
Table 1. e-TTS model parameters
ParameterSymbolPhysical meaningTodi clayBangkok sand
Isothermal mechanical parameter
B0kPaInterception of NCL10×1031.5×103
B1m3/kgSlope of NCL1×1037.7×103
ξRelated to K00.10.1
aControls rate effects0.450.5
HHysteretic strains (slope of unload curve)0.90.9
m2Controls elastic strain evolution and location of reload curve8×1046×105
m2=ηg/bηg: controls contribution of dteij on σijg
b: controls contribution of Tg on sg
m3Controls contribution of volumet ric and deviatoric strains on granular temperature production6×105
m3=ζg/bζg: controls contribution of dtεkk on σijg
m4kgs1°C1Rate of granular temperature production50010
m4=γ/bγ: dependency of energy dissipation rate
Thermal related parameter
βT°C1Volumetric thermal expansion of soils3.5×1058.1×106
Thermal creep mechanism I: thermally induced skeleton compaction
m6s2°C1Controls volume change of soil skeleton by temperature changes105
m6=θt/bθt: effect of thermal compaction on Tg
Thermal creep mechanism II: conversion from bound to free water
αbf°C1Affects conversion of bound water to free water during heating0.0237
m5s3m2°C1Controls amount of thermal volumetric strains by conversion of bound water2×105
m5=<ψg/b>ψg: rate of bound water changes on Tg0 (cooling)
Thermal creep mechanism III: thermoelastic induced volumetric change
m1,0Controls elastic strain evolution0.120.04
m1=λv/λsλv: controls contribution of εve on Yij
λs: controls contribution of εse on Yij
LT°C1Controls temperature dependence of NCL0.010.01
All model parameters listed in the Table 1 remain constant under temperature changes or gradients. It should be noted that m5 is constant for a heating process and m5=0 for cooling. The bound water content or bound water porosity ϕbw=Vbw/V changes with the temperature.

Thermal and Mechanical Creep

In order to describe the mechanical and thermal creep behaviors of clay, Burghignoli et al. (2000) conducted a series of heating and cooling tests on reconstituted Todi clay at constant vertical effective stress in a modified triaxial apparatus. The test specimens were all consolidated under isotropic conditions (σp=392  kPa), some were unloaded/swelled to 49 kPa (OCR=8), and others were reloaded to 98 kPa (OCR=4). In the first series, a heating-cooling cycle was applied to the specimen at each OCR state. Heating was performed for 520 min (with ΔT=26°C) and then the temperature was kept constant for the next 860 min. Subsequently, the temperature was decreased to the initial temperature (22°C) during a 520-min cooling period, and finally held constant at 22°C for 1,100 min. A temperature change rate of 3°C/h was adopted to avoid the generation of excess pore water pressure (measured Δu±2  kPa) and hence achieve constant effective stress conditions. In the second series, three specimens consolidated under the same hydrostatic stress of 392 kPa with different OCR values of 1, 4, and 8 experienced the creep process at the room temperature of 22°C. The creep process lasted 3,000 min.

Influence of Stress History on Thermal Creep of Clay

Fig. 3 compares the measured experimental results for creep due to thermal and isothermal phases of tests on reconstituted Todi clay reported by Burghignoli et al. (2000) with corresponding simulation results using the e-TTS model. The measured data show that isothermal creep is strongly influenced by stress history with contractive creep strains occurring for OCRs=1.0, 4.0, and volumetric expansion at OCR=8 [Fig. 3(a)]. The thermal cycle amplifies the magnitudes of compressive strains at OCR=1 and expansive strains at OCR=8 [Fig. 3(b)]. The e-TTS captures the measured effects of OCR for creep under isothermal conditions and the relative magnitudes of thermal creep strains associated with the imposed cycle of heating and cooling. The model tends to overestimate the measured compressive creep strains at OCR=1.0 and underestimates expansive strains at OCR=8.0 using the current calibration of model parameters. However, it should be noted that most existing thermomechanical constitutive models ignore the thermally activated creep (Hueckel and Borsetto 1990; Cui et al. 2000; Abuel-Naga et al. 2007; Di Donna and Laloui 2015). As a consequence, either the thermal deformation is significantly underestimated or the thermal creep is significantly overestimated to fit to the total creep strain. Increasing the heating time (or decreasing the heating rate) inevitably induces a larger difference between the thermal creep and isothermal creep strains.
Fig. 3. Model performance for thermal and mechanical creep on Todi clay: (a) mechanical creep strain; and (b) thermal creep strain. (Data from Burghignoli et al. 2000.)
Fig. 4(a) shows the experimental results and simulation results of the isotropic compressive curve for reconstituted Todi clay. A typical unloading/reloading hysteretic loop was reproduced by the e-TTS model. In the e-TTS model, the locked energy (i.e., hysteretic strain) represents the influence of the stress history on the isothermal/mechanical and thermal creep. In general, the magnitude of thermal creep is smaller in OC soils than in NC soils. The transition from compressive to expansive creep strains has been characterized in experimental studies at a critical OCR (Yao and Fang 2020). The critical OCR can also be used to estimate whether the thermal creep will induce contraction or expansion. Within the e-TTS model, the critical OCR is controlled by the unloading module through the parameter h that describes hysteretic response in unloading and reloading [Eq. (18)].
Fig. 4. Effect of consolidation stress history on creep strain: (a) effect of parameter h in e-TTS model; and (b) variations of elastic and hysteretic strain during unloading.
Fig. 4(a) illustrates e-TTS model predictions of the compressive and swelling behavior at selected values of h. When the Todi clay was unloaded from normally consolidated state (p=396  kPa) to 98 and 49 kPa, the corresponding OCR values were 4 and 8, respectively. Once the soil is unloaded, the locked strain increases, accompanied by increasing OCR, while the elastic strain decreases due to the reduction in effective stress [h=0.1, 0.3, and 0.9 simulated by e-TTS, Fig. 4(b)]. The difference between elastic and locked strain transit from negative to positive, resulting in a plastic compressive or swelling strain [Eq. (16)]. Therefore, the e-TTS model demonstrates that lower values of the OCR result in higher values of locked strain, indicating that the thermal creep will generate compressive strains. Conversely, higher values of the OCR result in lower values of the locked strain, and thermal creep predicts swelling behavior.

Effect Magnitude of Temperature Change on Thermal Creep

The e-TTS model reproduces a series of thermal isotaches for NC Todi clay to describe the isothermal and thermal creep. The magnitude of isothermal creep is controlled by the viscous parameter, a [Eq. (16)], which describes the effects of strain rate on the preconsolidation pressure, and in turn influences the location of the normal consolidation line. Smaller values of a increase the magnitude of the creep strains [Fig. 5(a)]. The magnitude of thermal creep is controlled by the viscous parameter, a, and thermal parameter LT [Eq. (16)], which describes the combining effects of strain rate and temperature on the pre-consolidation pressure. Smaller values of a and/or larger values of LT increase the magnitude of the thermal creep strains [Fig. 5(a)]. Fig. 5(b) shows the effects of heating rate and magnitude of temperature change on thermal creep. For a NC clay specimen consolidated to 200 kPa [Point A; Fig. 5(a)], isothermal creep at constant effective produces 0.34% compressive strain over a period of 3,000 mins [A-B Fig. 5(b)]. A temperature increase, ΔT=10°C, applied over the same time period increases the creep strain (A-C; 0.58%) at the same effective stress state, while ΔT=20°C, produces 0.87% compressive strain (A-D) over the same time period. This result clearly shows the role of strain rate and temperature on isotache states of NC clays, where the thermal creep depends on the viscous parameters α and LT [m1 in Eq. (16)]. In summary, increasing the magnitude of temperature rise induces a higher thermal creep. The greater the magnitude of temperature rise, the larger the contribution of thermal creep to the total deformation.
Fig. 5. Effect magnitude of temperature change on thermal creep produced by e-TTs model: (a) thermal isotaches for Todi clay; and (b) isothermal and thermal creep.

Thermal Creep Strain Accumulation and Stabilization in Saturated Sands

The e-TTS model considers the influences of porosity and effective stress on the initial inelastic strain (Appendix) and temperature-induced skeleton compaction through evolution of the granular temperature, Tg [Eq. (131)]. Sittidumrong et al. (2019) carried out cyclic heating-cooling tests and creep tests on Bangkok sand under 1D conditions in modified oedometer cell. Their experimental program comprised pairs of tests (done in parallel) on saturated specimens of reconstituted sand prepared at the same water content. One specimen in each pair was subjected to conventional incremental loading under isothermal conditions (T=28°C), while the second was subject a series of 100 thermal cycles (T=28°C50°C) each with 1 hr duration at the imposed levels of effective stress. Heating rates were selected to ensure there was no development of pore water pressure in the specimens. The authors then report thermal creep by subtraction of the isothermal creep strains assuming identical behavior for each pair of test specimens. They present results at four selected formation densities (γd=17.7, 16.5, 15.7, and 15.0  kN/m3) and three levels of vertical effective stress (σv=2.01, 3.21, and 4.01 MPa).
Fig. 6 shows that the computed and measured thermal creep strains for the first five temperature cycles of Bangkok sand at two different initial density states [Dr=30%, 70% in Figs. 5(a and b), respectively], using a single set of e-TTS model input parameters (Table 1). The e-TTS model presents thermal swelling strain at first two elevated temperature, but the accumulative thermal strain after two temperature cycles tends to be compressive, which shows the same tendency as the measured accumulative strain. The e-TTS model reproduces the thermally induced compaction phenomenon for sand after a single temperature cycle and the increase of accumulative strain with cycles of heating and cooling.
Fig. 6. Comparison of computed and measured thermal creep strains of reconstituted Bangkok sand at σv=3.21  MPa over five 1-h temperature cycles (data exclude isothermal creep by subtraction from parallel test): (a) loose sand (Dr=30%, γd=15.0  kN/m3); and (b) dense sand (Dr=70%, γd=16.5  kN/m3).
Fig. 7 summarizes the predicted and measured accumulation of isothermal and thermal creep strains over 100 (1 h) temperature cycles for the Bangkok sand. The e-TTS model shows very good agreement between the computed and measured thermal creep strains. As expected, the accumulated thermal creep strains increase with the level of effective stress and reduce with increasing relative density.
Fig. 7. Effect of initial density and stress level on the predicted and measured accumulation of isothermal and thermal creep strains for Bangkok sand: (a) effect of stress level (Dr=98%, γd=17.7  kN/m3); and (b) effect of initial density (σv=3.21  MPa).
The e-TTS model predicts that strain accumulation approaches a definite limit for each initial state. This behavior can be related to the underlying elastic and hysteretic/locked strains. Fig. 8 shows the convergence of these state parameters over 60 thermal cycles (εveεvh=0) for the Bangkok sand at Dr=30% and σv=3.21  MPa.
Fig. 8. e-TTs model simulations.

Conclusion

This paper describes the formulation of the extended e-TTS model to interpret the long-term thermal-mechanical behavior of soils, including thermal creep and cyclic stabilization for both clay and sand. The model has been validated for two reference datasets (Todi clay and Bangkok sand) which include independent measurements of isothermal and thermally-activated creep strains. The following key points should be noted:
1.
Thermal-induced compaction phenomena exist in granular materials (sands, glass beads, etc.) where irreversible deformations are associated with local particle re-arrangement. This can be contrasted with cohesive soils where thermal creep was previously linked to the exchange of bound and free water for clay particles with high specific surface area;
2.
Isothermal (mechanical) creep is a special case of creep induced by varying temperature. Stress history affects the sign of volumetric strains associated with isothermal and thermal creep. While normally and lightly overconsolidated soils exhibit compressive creep strains, more highly overconsolidated materials expand. In general, increases in temperature (and thermal cycles) augment the magnitude of creep strains that occur under isothermal conditions;
3.
The thermal creep of sands is first described by the e-TTS model. A new formulation for granular entropy, including a new expression for dissipation flow, M, is proposed for interpreting the thermally-induced skeleton compaction within the granular materials [Eqs. (10) and (12)]. The e-TTS model is able to provide two distinct mechanisms that relate changes in granular temperature to the temperature of the soil and unify the description of thermal strains in sands and clays. The e-TTS model is able to provide two distinct mechanisms which relate changes in granular temperature to the temperature of the soil and the third mechanism which relates changes in elastic strain energy relaxation to the temperature of soil. Based on these improved energy considerations the e-TTS model unifies the description of thermal strains in sands and clays; and
4.
The e-TTS model also produces the thermal cyclic stabilization for sand over large number of thermal cycles due to equalization of elastic and locked-in hysteretic strains. The initial value of elastic strain (an important state variable in e-TTS model) is strongly dependent on the dry density and effective stress, reproducing the effects of initial density and stress level on the isothermal and thermal strain.
The e-TTS model can be applied for the bearing performance analysis of energy foundations in saturated clayey and sandy soils. The long-term and time-dependent settlement of surrounding soils induced by cyclic thermal variations and mechanical load can be well analyzed by the e-TTS model. However, the e-TTS model is limited to in-situ ground conditions that can be approximated as homogeneous and isotropic, and does not represent other physicochemical mechanisms associated with materials such as expansive, frozen, highly sensitive, or organic soils.

Appendix. Formulation of the e-TTS Model in Triaxial Stress Conditions

To take into consideration the effect of bound water on the thermoelastic expansion of the soil skeleton, βT can be expressed as βT=βs+(ϕb/ϕs)βb, where βs and βb are the thermoelastic expansion coefficients for soil particle and bound water, ϕs=Vs/V and ϕb=Vb/V (where V=total volume).
According to effective stress [Eq. (4)], the mean effective stress p and deviator stress q under triaxial stress conditions can be expressed as
p=1.4B(εve)2.5+1.5Bξ(εve)0.5(εse)2+3KeβTΔTq=6Bξ(εve)1.5εse
(14)
where the secant elastic bulk modulus Ke under isothermal conditions can be expressed as
Ke=1.4B(εve)0.5εve+1.5Bξ(εse)2(εve)0.5
(15)
Combining Eqs. (5)–(8), the elastic strain can be expressed under triaxial stress conditions
dtεve=dtεv3m1(Tg)a(εveεvh)dtεse=dtεs(Tg)a(εseεsh)
(16)
In Eq. (16), the model parameter m1 (=λv/λs) is considered as a function of temperature, m1=m1,0[1+LT(TT0)], among which m1,0 and LT are material parameters.
Combining equations [Eqs. (9)–(13)], the evolution of granular temperature, Tg is obtained under triaxial stress conditions can be expressed as follows [cf. Eq. (13)]
dtTg=m2(dtεs)2+m3(dtεv)2ρd+(m5αbfϕbwpρd(1ϕ)+m6)(dtT)2m4Tgρd
(17)
where m2, m4, and m5 = material parameters m2=ηg/b, m3=ζg/b, m5=ψg/b, m4=γ/b, and m6=θt/b, among which m5 remains a positive constant for heating and equals to 0 for cooling.
In Eq. (18), εvh and εsh are the hysteretic volumetric and deviator strains, respectively, which are the two state variables for the description of the hysteretic loop induced by loading-unloading. The hysteretic strain accounts for the effect of the stress history of soil. Once the soil is subjected to unloading, the hysteretic strain increases with decreasing effective stress level or increasing OCR. The hysteretic volumetric and deviator strains can be expressed as (Zymnis et al. 2018)
dtεvh=dtεvDdtεvD·εvh/3+dtεdD·εdhh0.5[(εvh)2/3+(εdh)2]1.5εvh
(18a)
dtεdh=dtεdDdtεvD·εvh/3+dtεdD·εdhh0.5[(εvh)2/3+(εdh)2]1.5εdh
(18b)
where h = material parameter, which controls the slope of unload-reload curve.
Finally, the mass conservation equation for saturated soils is written as
dtρd=ρddtεkk
(19)

Data Availability Statement

Some or all data, models, or code that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

This study was supported by the National Natural Science Foundation of China (Fund Nos. 51778338 and 41907231), to which we hereby express our sincere gratitude.

References

Abuel-Naga, H. M., D. T. Bergado, and A. Bouazza. 2007. “Volume change behavior of saturated clays under drained heating conditions: Experimental results and constitutive modeling.” Can. Geotech. J. 44 (8): 942–956. https://doi.org/10.1139/t07-031.
Abuel-Naga, H. M., D. T. Bergado, and A. Bouazza. 2009. “Thermomechanical model for saturated clays.” Géotechnique 59 (3): 273–278. https://doi.org/10.1680/geot.2009.59.3.273.
Agar, J. G., N. R. Morgenstern, and J. D. Scott. 1986. “Thermal-expansion and pore pressure generation in oil sands.” Can. Geotech. J. 23 (3): 327–333. https://doi.org/10.1139/t86-046.
Baldi, G., T. Hueckel, and R. Pellegrini. 1988. “Thermal volume changes of the mineral-water system in low-porosity clay soils.” Can. Geotech. J. 25 (4): 807–825. https://doi.org/10.1139/t88-089.
Brandl, H. 2006. “Energy foundations and other thermo-active ground structures.” Géotechnique 56 (2): 81–122. https://doi.org/10.1680/geot.2006.56.2.81.
Burghignoli, A., A. Desideri, and S. Miliziano. 2000. “A laboratory study on the thermomechnical behavior of clayey soils.” Can. Geotech. J. 37 (4): 764–780. https://doi.org/10.1139/t00-010.
Cekerevac, C., and L. Laloui. 2004. “Experimental study of the thermal effects on the mechanical behaviour of a clay.” Int. J. Numer. Anal. Methods Geomech. 28 (3): 209–228. https://doi.org/10.1002/nag.332.
Chen, K., K. Cole, C. Conger, J. Draskovic, M. Lohr, K. Klein, T. Scheidemantel, and P. Schiffer. 2006. “Granular materials: Packing grains by thermal cycling.” Nature 442 (7100): 257. https://doi.org/10.1038/442257a.
Collins, I. F. 2005. “The concept of stored plastic work of frozen elastic energy in soil mechanics.” Géotechnique 55 (5): 373–382. https://doi.org/10.1680/geot.2005.55.5.373.
Coussy, O. 1995. Mechanics of porous continua. Chichester, UK: Wiley.
Cui, Y. J., T. T. Le, and A. M. Tang. 2009. “Investigating the time-dependent behaviour of Boom clay under thermo-mechanical loading.” Géotechnique 59 (4): 319–329. https://doi.org/10.1680/geot.2009.59.4.319.
Cui, Y. J., N. Sultan, and P. Delage. 2000. “A thermomechanical model for saturated clays.” Can. Geotech. J. 37 (3): 607–620. https://doi.org/10.1139/t99-111.
Demars, K. R., and R. D. Charles. 1982. “Soil volume changes induced by temperature cycling.” Can. Geotech. J. 19 (2): 188–194. https://doi.org/10.1139/t82-021.
Di Donna, A., and L. Laloui. 2015. “Response of soil subjected to thermal cyclic loading: Experimental and constitutive study.” Eng. Geol. 190 (14): 65–76. https://doi.org/10.1016/j.enggeo.2015.03.003.
Divoux, T., H. Gayvallet, and J. C. Géminard. 2008. “Creep motion of a granular pile induced by thermal cycling.” Phys. Rev. Lett. 101 (14): 148303. https://doi.org/10.1103/PhysRevLett.101.148303.
Haff, P. K. 1983. “Grain flow as a fluid-mechanical phenomenon.” J. Fluid Mech. 134 (Sep): 401–430. https://doi.org/10.1017/S0022112083003419.
Houston, S. L., W. N. Houston, and N. D. Williams. 1985. “Thermo-mechanical behavior of seafloor sediments.” J. Geotech. Geoenviron. Eng. 111 (11): 1249–1263. https://doi.org/10.1061/(ASCE)0733-9410(1985)111:11(1249).
Hueckel, T., and G. Baldi. 1990. “Thermoplasticity of saturated clays: Experimental constitutive study.” J. Geotech. Geoenviron. Eng. 116 (12): 1778–1796. https://doi.org/10.1061/(ASCE)0733-9410(1990)116:12(1778).
Hueckel, T., and M. Borsetto. 1990. “Thermoplasticity of saturated soils and shales: Constitutive equations.” J. Geotech. Geoenviron. Eng. 116 (12): 1765–1777. https://doi.org/10.1061/(ASCE)0733-9410(1990)116:12(1765).
Jiang, Y. M., and M. Liu. 2003. “Granular elasticity without the coulomb condition.” Phys. Rev. Lett. 91 (14): 144301–144304. https://doi.org/10.1103/PhysRevLett.91.144301.
Jiang, Y. M., and M. Liu. 2007. “From elasticity to hypoplasticity: Dynamics of granular solids.” Phys. Rev. Lett. 99 (10):105501. https://doi.org/10.1103/PhysRevLett.99.105501.
Jiang, Y. M., and M. Liu. 2009. “Granular solid hydrodynamics.” Granular Matter 11 (3): 139–156. https://doi.org/10.1007/s10035-009-0137-3.
Kurz, D., J. Sharma, M. Alfaro, and J. Graham. 2016. “Semi-empirical elastic–thermoviscoplastic model for clay.” Can. Geotech. J. 53 (10): 1583–1599. https://doi.org/10.1139/cgj-2015-0598.
Laloui, L., and C. Cekerevac. 2003. “Thermo-plasticity of clays: An isotropic yield mechanism.” Comput. Geotech. 30 (8): 649–660. https://doi.org/10.1016/j.compgeo.2003.09.001.
Laloui, L., and A. Di Donna. 2011. “Understanding the behaviour of energy geo-structures.” Proc. Inst. Civ. Eng. Civ. Eng. 164 (4):184–191. https://doi.org/10.1680/cien.2011.164.4.184.
Leroueil, S., and M. E. S. Marques. 1996. “Importance of strain rate and temperature effects in geotechnical engineering.” In Measurement and modeling time dependent soil behaviour. Reston, VA: ASCE.
Liu, H. L., Y. Xiao, J. S. McCartney, and H. Liu. 2018. “Influence of temperature on the volume change behavior of saturated sand.” Geotech. Test. J. 41 (4): 747–758. https://doi.org/10.1520/GTJ20160308.
Mitchell, J. K., and K. Soga. 2005. Fundamentals of soil behavior. Hoboken, NJ: Wiley.
Ng, C. W. W., S. H. Wang, and C. Zhou. 2016. “Volume change behaviour of saturated sand under thermal cycles.” Géotech. Lett. 6 (2): 124–131. https://doi.org/10.1680/jgele.15.00148.
Pan, Y. Z., J. B. Coulibaly, and A. F. Rotta Loria. 2020. “Thermally induced deformation of coarse-grained soils under nearly zero vertical stress.” Géotech. Lett. 10 (4): 486–491. https://doi.org/10.1680/jgele.20.00013.
Qiao, Y. F., and W. Q. Ding. 2017. “ACMEG-TVP: A thermoviscoplastic constitutive model for geomaterials.” Comput. Geotech. 81 (1): 98–111. https://doi.org/10.1016/j.compgeo.2016.08.003.
Rotta Loria, A. F., and J. B. Coulibaly. 2021. “Thermally induced deformation of soils: A critical overview of phenomena, challenges and opportunities.” Geomech. Energy Environ. 25 (2): 100193. https://doi.org/10.1016/j.gete.2020.100193.
Sittidumrong J., A. Jotisankasa, and K. Chantawarangul. 2019. “Effect of thermal cycles on volumetric behaviour of Bangkok sand.” Geomech. Energy Environ. 20 (Dec): 100127. https://doi.org/10.1016/j.gete.
Šuklje, L. 1957. “The analysis of the consolidation process by the isotaches method.” In Proc., 4th Int. Conf. Soil Mechanics. Foundation Engineering, 200–206. London: Butterworths.
Sultan, N., P. Delage, and Y. J. Cui. 2002. “Temperature effects on the volume change behaviour of Boom clay.” Eng. Geol. 64 (2–3): 135–145. https://doi.org/10.1016/S0013-7952(01)00143-0.
Towhata, I., P. Kuntiwattanakul, I. Seko, and K. Ohishi. 1993. “Volume change of clays induced by heating as observed in consolidation tests.” Soils Found. 33 (4): 170–183. https://doi.org/10.3208/sandf1972.33.4_170.
Xiong, Y. L., S. Zhang, G. L. Ye, and F. Zhang. 2014. “Modification of thermo-elasto-viscoplastic model for soft rock and its application to THM analysis of heating tests.” Soils Found. 54 (2): 176–196. https://doi.org/10.1016/j.sandf.2014.02.009.
Yan, W. M., and X. S. Li. 2011. “A model for natural soil with bonds.” Géotechnique 61 (2): 95–106. https://doi.org/10.1680/geot.8.P.061.
Yao, Y. P., and Y. F. Fang. 2020. “Negative creep of soils.” Can. Geotech. J. 57 (1): 1–16. https://doi.org/10.1139/cgj-2018-0624.
Yashima, A., S. Leroueil, F. Oka, and I. Guntoto. 1998. “Modelling temperature and strain rate dependent behaviour of clays: One dimensional consolidation.” Soils Found. 38 (2): 63–73. https://doi.org/10.3208/sandf.38.2_63.
Yuan, Y. X., and A. J. Whittle. 2018. “A novel elasto-viscoplastic formulation for compression behaviour of clays.” Géotechnique 68 (12): 1044–1055. https://doi.org/10.1680/jgeot.16.P.276.
Yuan, Y. X., and A. J. Whittle. 2021. “Formulation of a new elastoviscoplastic model for time-dependent behavior of clay.” Int. J. Numer. Anal. Methods Geomech. 45 (6): 843–864. https://doi.org/10.1002/nag.3174.
Zhang, S., W. M. Leng, F. Zhang, and Y. L. Xiong. 2012. “A simple thermo-elastoplastic model for geomaterials.” Int. J. Plast. 34 (Jul): 93–113. https://doi.org/10.1016/j.ijplas.2012.01.011.
Zhang, Z. C., and X. H. Cheng. 2016. “A thermo-mechanical coupled constitutive model for clay based on extended granular solid hydrodynamics.” Comput. Geotech. 80 (Dec): 373–382. https://doi.org/10.1016/j.compgeo.2016.05.010.
Zhang, Z. C., and X. H. Cheng. 2017. “A fully coupled THM model based on a non-equilibrium thermodynamic approach and its application.” Int. J. Numer. Anal. Methods Geomech. 41 (4): 527–554. https://doi.org/10.1002/nag.2569.
Zhou, C., K. Y. Fong, and C. W. W. Ng. 2017. “A new bounding surface model for thermal cyclic behavior.” Int. J. Numer. Anal. Methods Geomech. 41 (16): 1656–1666. https://doi.org/10.1002/nag.2688.
Zymnis, D. M., and A. J. Whittle. 2020. “Geotechnical considerations in the design of borehole heat exchangers in clay.” Can. Geotech. J. 58 (9): 1247–1262. https://doi.org/10.1139/cgj-2020-0118.
Zymnis, D. M., A. J. Whittle, and X. H. Cheng. 2018. “Simulation of long-term thermo-mechanical response of clay using an advanced constitutive model.” Acta Geotech. 14 (2): 295–311. https://doi.org/10.1007/s11440-018-0726-6.
Zymnis, D. M., A. J. Whittle, and J. T. Germaine. 2019. “Measurements of temperature-dependent bound water in clays.” Geotech. Test. J. 42 (1): 232–244. https://doi.org/10.1520/GTJ20170012.

Information & Authors

Information

Published In

Go to Journal of Geotechnical and Geoenvironmental Engineering
Journal of Geotechnical and Geoenvironmental Engineering
Volume 148Issue 6June 2022

History

Received: Mar 19, 2021
Accepted: Jan 25, 2022
Published online: Apr 13, 2022
Published in print: Jun 1, 2022
Discussion open until: Sep 13, 2022

Authors

Affiliations

Postdoctoral, Dept. of Civil Engineering, Tsinghua Univ., Beijing 100084, China. ORCID: https://orcid.org/0000-0003-2057-2607
Xiaohui Cheng, M.ASCE [email protected]
Associate Professor, Dept. of Civil Engineering, Tsinghua Univ., Beijing 100084, China (corresponding author). Email: [email protected]
Andrew J. Whittle, M.ASCE https://orcid.org/0000-0001-5358-4140
Edmund K. Turner Professor, Dept. of Civil and Environmental Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139. ORCID: https://orcid.org/0000-0001-5358-4140

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.

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share