HOSTED BY Available online at www.sciencedirect.com ScienceDirect Soils and Foundations 62 (2022) 101181 www.elsevier.com/locate/sandf Technical Paper A simplified method for evaluating temperature effect on the behavior of layered soil with a time-varying cylindrical heat source Lujun Wang Center for Hypergravity Experimental and Interdisciplinary Research, MOE Key Laboratory of Soft Soils and Geoenvironmental Engineering, College of Civil Engineering and Architecture, Zhejiang University, Hangzhou 310058, China Received 23 July 2021; received in revised form 2 June 2022; accepted 9 June 2022 Available online 1 July 2022 Abstract The engineering application of cylindrical heat source usually influences the characteristics of the surrounding soil and induces significant coupling responses of temperature, seepage and stress fields. This paper presents a numerical investigation into the thermohydraulic-mechanical (THM) coupling behavior of layered soils surrounding a cylindrical heat source by combining the analytical layer element method (ALEM) and the finite element method (FEM). The FEM is utilized to analyze the soil-cylindrical heat source interaction and the ALEM is used to obtain the fundamental solution of the layered soil. Then the THM coupling solution between the heat source and the soil is obtained by the simplified FEM-ALEM coupling method. Detailed validation against experimental result, and verifications against semi-analytical solution and numerical results by Comsol Multiphysics are performed to confirm the robustness of the present method, followed by extensive parametric studies to examine the effects of decaying half-life and type of the time-varying heat source, and the soil layered characteristics on the behavior of soils. It’s evident that the present method is accessible and can be implemented via common programming language, e.g. Fortran. Ó 2022 Production and hosting by Elsevier B.V. on behalf of The Japanese Geotechnical Society. This is an open access article under the CC BYNC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/). Keywords: Temperature effect; Layered soil; Simplified method; Consolidation; Time-varying load; Heat source 1. Introduction Shallow geothermal energy is one of the most promising options in addressing the energy shortage problem and has become the fastest growing clean energy resource over the world in recent years, contributing to the ever-increasing number of ground source heat pump (GSHP) systems (Brandl, 2006; Laloui et al., 2006; Faizal et al., 2019; Loveridge et al., 2020). During the service period, the temperature variation of GSHP usually leads to various processes and changes in soil–water system including the interaction between the thermal, hydraulic and mechanical fields in the surrounding soils. The temperature gradient Peer review under responsibility of The Japanese Geotechnical Society. E-mail address: lujunwang@zju.edu.cn and pore pressure gradient, in these engineering applications, cause the movement of pore water and conduction of heat energy within the soils and then affect the engineering behavior of soils (Selvadurai and Suvorov, 2014; Shahrokhabadi et al., 2020; Song et al., 2019; Wang and Wang, 2020; Zagorščak et al., 2017). Generally, a GSHP system can be taken as a special form of heat source or sink for the ground. For the stable temperature and high specific heat of the surrounding soils, the energy efficiency of GSHP system is higher than that of traditional air source heat pump (Self et al., 2013; Pan et al., 2020). Ground heat exchanger (GHE) is one of the key parts in GSHP system, whose thermal behavior and heat transfer performance are significant to the operation of GSHP system (Pan et al., 2020), and can be treated as a vertical cylindrical heat source (VCHS) (Faizal et al., https://doi.org/10.1016/j.sandf.2022.101181 0038-0806/Ó 2022 Production and hosting by Elsevier B.V. on behalf of The Japanese Geotechnical Society. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/). L. Wang Soils and Foundations 62 (2022) 101181 Nomenclature c C E E G K n p s t T u uj permeability coefficient (m/s) specific heat capacity (J/kg°C) Young’s modulus (kPa) constitutive tensor depending on E and l elastic shear modulus (kPa) heat conduction coefficient (W/m°C) Porosity of soil pore fluid pressure (kPa) Laplace transform parameter with respect to t time (s) temperature increment (°C) solid desplacement (m) displacement in j-th direction (m), j = r, z a af l j q f n e r cw linear thermal expansion coefficient of solid matrix (/°C) linear thermal expansion coefficient of pore fluid (/°C) Poisson’s ratio thermal diffusivity coefficient (m2/s) mass density (kg/m3) Laplace transform parameter with respect to z Hankel transform parameter with respect to r strain stensor total stress tensor (kPa) pore fluid unit weight (kN/m3) oped, which is usually different from that in engineering practice. Up to now, several methods or theories, such as the finite layer method (FLM) (Small and Booker, 1986; Mei et al., 2004; Bao et al., 2015) and the transfer matrix method (TMM) (Pan, 1999; Ai et al., 2002; Zheng et al., 2021), have been generally developed to investigate the behavior of layered media. It is worth mentioning that the stiffness matrix in FLM or TMM usually contains positive exponential functions, leading to the problem of computation overflow. To effectively avoid the computation overflow problem, a newly developed method named the analytical layer element method (ALEM) (Ai and Hu, 2015; Wang and Ai, 2018), whose matrices include none of positive exponential function, is employed to obtain the fundamental solution for the THM coupling problems of layered soils. The ALEM has been utilized in investigating the response of layered thermoelastic media (Wang and Ai, 2018) and poroelastic media (Ai and Hu, 2015; Wang et al., 2020). The classical numerical methods, e.g. FEM (Rotta Loria et al., 2015; Sutman et al., 2019) and the boundary element method (BEM) (Smith and Booker, 1996; EI-Zein, 2006), are able to implement the THM coupling behaviors of layered soils, large requirement of computational storage is needed and the solution is highly dependent on the discrete approach. In this paper, a simplified method is developed by coupling the ALEM and FEM to evaluate the temperature effects on the behavior of layered and normally consolidated soft soil subjected to a time-varying GHE. FEM is employed to analyze the cylindrical heat source with time-varying power output, and the ALEM (Ai and Wang, 2015) is used to obtain the fundamental solution of layered soil, then the THM coupling solution between the heat source and soil is acquired by coupling FEM and ALEM. Applicability and robustness of the simplified coupling method are confirmed through extensive comparisons with available experimental data, semi-analytical 2019; Wang et al., 2019). Regarding the complex properties of soils and loading characteristics of the vertical cylindrical heat source, distinctly different behavior of the surrounding soil can be expected, which greatly affects the soil-VCHS interaction. This is required to comprehensive understand the thermal response of the soil-VCHS system. Thermo-mechanical soil–VCHS interaction has been investigated by different means, namely, numerical approach (e.g. Dupray et al., 2014; Suryatriyastuti et al., 2014; Di Donna and Laloui, 2015; Ozudogru et al., 2015; Zagorščak et al., 2017; Sutman et al., 2019; Maiorano et al., 2019; Moradshahi et al., 2020; Moradshahi et al., 2021), analytical method (e.g. Loveridge et al., 2015; Chen and McCartney, 2016; Pasten and Santamarina, 2014; Olgun et al., 2014), field test (e.g. Kalantidou et al., 2012; Murphy and McCartney, 2014; Kramer et al., 2015; Murphy and McCartney, 2015; Faizal et al., 2019), and centrifuge test (e.g. McCartney and Rosenberg, 2011; Stewart and McCartney, 2014; Goode III and McCartney, 2015; Ng et al., 2015; Rotta Loria et al., 2015; Wang et al., 2015). It is worth noting that the previous studies assume the soil is homogeneous media. However, due to long term geological sedimentation processes, natural soils are normally distributed in layers and necessary to take the layered characteristics into consideration. Thus, it is necessary to extend the research scope from homogeneous soil to layered ones. Therefore, several studies for examining the behavior of layered soil surrounding GHE are developed by experimental studie (Guo et al., 2018; Li et al., 2019; Pan et al., 2020), numerical method (Florides et al., 2013; Luo et al., 2014) and analytical method (Erol and Francois, 2018; Pan et al., 2020; Wang et al., 2015; Zhou et al., 2016; Wang, 2022). In engineering practice, the temperature of GSHP is usually periodical variation with time (Laloui et al., 2006; Selvadurai and Suvorov, 2014; Shahrokhabadi et al., 2020). Several heat source models, treating the source intensity as constant, have been devel2 L. Wang Soils and Foundations 62 (2022) 101181 div r ¼ 0 solutions and numerical results using Comsol Multiphysics. Extensive parametric studies are performed to investigate the effects of temperature increment, heat source’s type and layered characteristic on the soil behavior. The main innovations of this paper can be summarized as follows: (1) the ALEM imposes no restriction on the layer thickness and no limitation on the number of layers, and avoids the exponential overflow and computational instability caused by ill-conditioned matrices; (2) temperature and excess pore pressure responses of homogeneous and multilayered soil surrounding a vertical GHE are compared; (3) the time-dependent behavior of layered soils subjected to a time-varying heat source is explored; (4) a generalized cylindrical heat source model, which relates to solid cylindrical heat source in energy piles, tubular and hollow cylindrical heat source in GSHP systems, is developed to study the responses of soils in different engineering applications. ð1Þ where div denotes the divergence. According to the generalized thermoelastic Hooke’s law and the effective stress principle, the constitutive equations can be written as (Biot, 1956; Laloui, et al., 2006). div r ¼ div ðE : eÞ bgradT gradp ð2Þ where grad represents the gradient and e ¼ grad u, b ¼ 2Gagð1 þ lÞ, here g ¼ 1=ð1 2lÞ and G ¼ E=2ð1 þ lÞ. Considering Fourier’s law and Darcy’s law, the total heat flow QT and the pore water flow Qf along z- direction are given as. Z t @T dt ð3aÞ K QT ¼ @z 0 Z t c@p Qf ¼ dt ð3bÞ 0 cw @z 2. Methodology Considering Fourier’s law and energy balance equation, one can obtain the heat conduction equation as. 2.1. Fundamental solution of layered soils with a heat source @T ¼ divðjgradT Þ @t ð4Þ where j ¼ K=qC. According to mass balance equation and Darcy’s law, the continuity equation of seepage is derived as (Ai and Wang, 2018). @ev @T c ¼ div þ au gradp ð5Þ @t cw @t Illustrated in Fig. 1 is a cylindrical source embedded in a multi-layered saturated soil system. The heat source changes temperature field by transferring the heat to the surrounding soil and subsequently influences the excess pore pressure (EPP) dissipation and the soil deformation. To evaluate the temperature effect of the heat source on the behavior of layered soil, ALEM is firstly employed to obtain the soil fundamental solution, and then based on the principle of FEM, a simplified method by coupling the ALEM and FEM is developed to solve this problem. For saturated normally consolidated homogeneous isotropic soft soils, the equilibrium equation for the thermal consolidation problem ignoring the body force in the cylindrical coordinate system is (Biot, 1956; Booker and Savvidou, 1985). r z þ urr þ @u . where au ¼ 3að1 nÞ þ 3naf and ev ¼ @u @r @z To solve the previous partial differential equations (PDEs), the integral transform approach is utilized to convert them into ordinary differential equations (ODEs). The mth-order Laplace-Hankel transform of f ðr; z; tÞ and its inversion are (Sneddon, 1972). Z þ1 f m ðn; z; tÞ ¼ f ðr; z; tÞJ m ðnrÞrdr ð6aÞ 0 Z þ1 f m ðn; z; sÞ ¼ est f m ðn; z; tÞds ð6bÞ 0 where the overbars ‘‘–” and ‘‘” are hereafter used to denote Hankel transform (HT) and Laplace transform (LT) of a given function, respectively. Taking the application of LT and HT to Eqs. (1)–(5), the PDEs are transformed into ODEs, one obtains (Ai and Wang, 2015). ð7Þ K ðn; z; sÞ ¼ ½ A1 ðn; z; sÞ A2 ðn; z; sÞ Rðn; 0; sÞ where " #T K ðn; f; sÞ ¼ ur ðn; f; sÞ; uz ðn; f; sÞ; p ðn; f; sÞ; T ðn; f; sÞ " Fig. 1. A layered soil system with a cylindrical heat source in the cylindrical coordinate system. , #T Rðn; 0; sÞ ¼ ur ðn; 0; sÞ; uz ðn; 0; sÞ; p ðn; 0; sÞ; T ðn; 0; sÞ; rrz ðn; 0; sÞ; rz ðn; 0; sÞ; Qf ðn; 0; sÞ; QT ðn; 0; sÞ 3 , L. Wang 2 where A1 ðn; z; sÞ and A2 ðn; z; sÞ refer to matrices of order 4 4. From Eq. (7), the following expression can be obtained. 3 2 6 K ðn; 0; sÞ 7 5¼ 4 K ðn; z; sÞ I44 044 A1 ðn; z; sÞ A2 ðn; z; sÞ The 2 here 044 and I44 , respectively, denote 0 matrix and the unit matrix of order 4 4. Taking the application of HT and LT to Eqs. (2) and (3), one obtains. 3 6 C ðn; 0; sÞ 7 5¼ 4 C ðn; z; sÞ 044 B1 ðn; z; sÞ " 3 6 K ðn; 0; sÞ 7 4 5 K ðn; z; sÞ and C ðn; z; sÞ I44 Rðn; 0; sÞ ¼ Nðn; z; sÞ Rðn; 0; sÞ B2 ðn; z; sÞ between 6 C ðn; 0; sÞ 7 6 K ðn; 0; sÞ 7 4 5 ¼ Uðn; z; sÞ 4 5 ð10Þ K ðn; z; sÞ 1 where Uðn; z; sÞ ¼ Nðn; z; sÞ Mðn; z; sÞ is a 8 8 stiffness matrix establishing the relationships between the generalized stresses and displacements in the transform domain of a single layer, as shown in Fig. 2. The elements of Uðn; z; sÞ are listed in Appendix II and detailed derivation can refer to the work by Ai and Wang (2015). ð9aÞ here relationship 3 6 C ðn; 0; sÞ 7 4 5 linked by the coefficients vector Rðn; 0; sÞ C ðn; z; sÞ can be derived as. 2 3 2 3 Rðn; 0; sÞ ¼ Mðn; z; sÞ Rðn; 0; sÞ ð8Þ 2 Soils and Foundations 62 (2022) 101181 #T C ðn; z; sÞ ¼ rrz ðn; z; sÞ; rz ðn; z; sÞ; Qf ðn; z; sÞ; QT ðn; z; sÞ 2.2. Simplified method for layered soil ð9bÞ where B1 ðn; z; sÞ and B2 ðn; z; sÞ denote matrices of order 4 4. A layered soil system, consisting of n layers with different material properties and thicknesses, embedded with a cylindrical source is illustrated in Fig. 3. The thickness of the ith layer is hi ¼ H i H i1 , here H i and H i1 denote the depths from the surface to the bottom and top of the ith layer. Applying Eq. (10) to the ith layer produces. 2 3 2 3 6 C ðn; H i1 ; sÞ 7 6 K ðn; H i1 ; sÞ 7 4 5 ¼ UðiÞ 4 5 C ðn; H i ; sÞ ð11Þ K ðn; H i ; sÞ here UðiÞ ¼ Uðn; hi ; sÞ is the element of the ith layer. As illustrated in Fig. 3(a) and (b), the length and radius of the source are L and r0 , respectively. The total heat flow Fig. 2. The generalized stresses and displacements for a single soil layer. Fig. 3. Diagram of a cylindrical heat source buried in layered soil. 4 L. Wang of the heat source is QT 0 and it is uniformly distributed along the length. The cylindrical heat source is divided into N 0 subsegments of micro heat source along the depth and each micro heat source is replaced by a uniform micro plane heat source in the middle of the subsegment whose heat flow is qzm ðtÞ. Assuming the length of the micro heat source is Dl yields. L¼ N0 X Dl where P ðn; H i ; sÞ is the general micro external load vector in LT and HT domain. The detailed expressions for dif ferent types of P ðn; H i ; sÞ are provided in Appendix I. Considering the continuity conditions of the soil layers’ interfaces and merging the analytical layer elements, the relationship between generalized stresses and displacements for the behavior of multilayered soil embedded with a cylindrical heat source can be established. For the case where the heat source is embedded in an arbitrary depth, as shown in Fig. 3(a), the global stiffness matrix is assembled as. ð12aÞ m¼1 QT 0 ¼ pr20 N0 Z X m¼1 mDl qz ðtÞdz ¼ N0 X qzm ðtÞ Soils and Foundations 62 (2022) 101181 ð12bÞ m¼1 ðm1ÞDl Eq. (12b) presents the expression of the total heat flow of a cylindrical heat source. For a hollow heat source whose external and internal radii are r1 and r2 respectively, the expression of the total heat flow is. N 0 Z mDl Z r2 N0 X X QT 0 ¼ 2pr qz ðtÞdrdz ¼ qzm ðtÞ ð12cÞ m¼1 ðm1ÞDl r1 m¼1 ð15aÞ here qz ðtÞ denotes the density of heat flow along the depth. Supposing that the surface is free and the bottom at z ? 1 is fixed and impermeable, produces. rzz ðn; 0; sÞ ¼ rrz ðn; 0; sÞ ¼ u ðn; 0; sÞ ¼ 0 For the case where the surface of a cylindrical heat source is flush with the surface of layered soil, as shown in Fig. 3(b), the global stiffness matrix is assembled as. ð13aÞ ur ðn; z; sÞ ¼ uz ðn; z; sÞ ¼ Qf ðn; z; sÞ ¼ QT ðn; z; sÞ ¼ 0; as z ! 1 ð13bÞ It’s worth noting that Eqs. (13a) and (13b) can be regarded as a general form of boundary conditions. For other application of practice, the boundary conditions presented here are available by replacing the corresponding expression in Eqs. (13a) and (13b) with the needed condi tions, e.g., when the bottom is permeable, Qf ðn; z; sÞ is just needed replaced by u ðn; z; sÞ ¼ 0 in Eq. (13b). Considering the continuity condition between any two adjacent layers without a heat source, the boundary condition between the (j + 1)th layer and jth layer can be given as. þ C n; H þ ; s ¼ C n; H ; s ; K n; H ; s ¼ K n; H ; s j j j j ð15bÞ According to Eqs. (15a) and (15b) and considering the boundary conditions, solution for the behavior of layered saturated soil with a time-varying heat source in LT and HT domain can be obtained. To obtain the solutions in physical domain, the Talbot algorithm (Talbot, 1979) is utilized to inverse LT and the Gauss–Legendre quadrature (Ai et al., 2002; Wang and Ai, 2018) is employed to inverse HT. Regarding a heat pump system, the number of the cylindrical heat source is ns , and each heat source is divided into N s subsegments of micro heat sources. The general load– displacement relationship of a single heat source is given as that of Eq. (15), which can be rewritten as. ð14aÞ where H þ j and H j are the depth H j of the ðj þ 1Þth layer and jth layer; respectively: At the interface plane z ¼ H i , the continuity conditions of a micro plane heat source are. C n; H þ i ; s ¼ C n; H i ; s þ P ðn; H i ; sÞ; ð14bÞ K n; H þ i ; s ¼ K n; H i ; s KjT ¼ Uj 1 CjT ð16Þ here CjT and KjT denote global vectors of the nodal displacement and interaction force of soil by the jth heat source, 1 respectively. Uj denotes the global flexibility matrix of 5 L. Wang layered soil surrounding a cylindrical heat source by Pan et al. (2020). The present solution and the experiment result provided by Pan et al. (2020) are compared to confirm the method, as shown in Fig. 4. The soil properties are listed in Table 2. The surface temperature boundary condition is 23.5 °C, before starting the experiment, the soil was left to stabilize until its initial average temperature was 23.5 ° C. The length and diameter of the cylindrical heat source are 200 mm and 6 mm, respectively. The soil depth was a half length greater than the length of the heat source to minimize the bottom boundary effect on the temperature in the short term test. Reasonably good agreement can be observed between the present results and the experiment data by Pan et al. (2020). It’s worth noting that the slight difference between the two results may be owing to the effects of surface boundary condition and pore water, while the trends are the same. soils, which is obtained by ALEM as shown in Eq. (11) and Appendix II. The global matrix equations of the heat pump system can be obtained by applying Eq. (16) to each single heat source in the heat pump system, which is expressed as. KjT ¼ Uj 1 CjT Soils and Foundations 62 (2022) 101181 ð17Þ where KTG and CTG are the global column vectors of order ns ðN s þ 1Þ 1. Based on Eq. (17), the relationship among the nodal temperature and nodal displacement of the heat pump system can be established and then further obtain the solutions. It’s known that the discretization of the cylindrical heat source is a key factor for the accuracy and stability of the solution. Four values of discretization number N were discussed: N = 10, 16, 20, 50. The relative error is defined as jT T j ee ¼ NT 50 100%, here T N is the dimensionless tempera50 ture when the cylindrical heat source is divided into N subsegments. As shown in Table 1, ee is as high as 9% between N = 10 and N = 50, ee decreases with the increase of N. ee is less than 0.51% when N = 20. On the consideration of the computational efficiency and accuracy, N = 20 is chosen in the following examples. The main process for the presented method can be summarized as: (1) based on the basic equations for THM coupling problems, the relationship for the state vectors of a single layer is described by an analytical layer element, which is obtained using LT and HT; (2) dividing the cylindrical heat source into a series of subsegments of micro heat sources along the depth, each micro heat source is replaced by a uniform micro plane heat source in the middle of the subsegment; (3) considering the initial and continuity conditions of the multilayered soil, the solution for the behavior of layered soil with a cylindrical heat source is obtained in the transform domain by coupling ALEM and FEM; (4) the actual solution in the physical domain is further obtained by inversing LT and HT. 2.3. Validation and verification (1) Case 1: compared with experiment A laboratory-scale experiment has been carried out to investigate the variation of temperature filed of a twoTable 1 The relative error with different discretization number N. N r=r0 10 16 20 50 ee 0.42766 9.243 0.40494 3.439 0.39347 0.509 0.39148 0.000 4 T N ee 0.16983 3.908 0.17365 1.748 0.17598 0.426 0.17673 0.000 10 T N ee 0.06166 5.416 0.06349 2.602 0.06488 0.469 0.06519 0.000 1 T N Fig. 4. Comparison of temperature distribution along depth surrounding a cylindrical source in a two-layered soil. 6 L. Wang Soils and Foundations 62 (2022) 101181 (2) Case 2: compared with FLM Another example is presented to compare the present solution with those by FLM (Savvidou and Booker, 1989). In the example, a solid cylindrical heat source with a length of h and a radius of r0 is embedded in the homogeneous soil, the soil thickness is 200h and the depth from the soil surface to the heat source top is 100h, here h ¼ 20r0 . The soil parameters and normalized variables are taken as: l ¼ 0:25, s ¼ jt=r0 , T N ¼ 0:02387QT 0 t=r0 , QT 0 denotes the total heat flow of the heat source. All the adopted parameters are the same as those given by Savvidou and Booker (1989). Fig. 5 illustrates the temperature variation curves at the mid-depth of the source with time, and a good match between the present results and the results by FLM (Savvidou and Booker, 1989) can be observed as well. Fig. 6. Comparison of EPP along z-direction surrounding a cylindrical source in a two-layered soil. (3) Case 3: compared with FEM Comsol Multiphysics) is divided into 9720 quadrilateral elements, the temperature change and infiltration condition at the surface boundary of the model are 0 °C and permeable, the bottom boundary conditions are insulated and impermeable. Here c1 ¼ c2 , c1 and c2 are the permeability coefficients of the two layers, the normalized variable pN 2G is defined as pN ¼ vT N , here v ¼ 12l ½au ð1 lÞ að1 þ lÞ and T N ¼ 0:02387q0 t=r0 . Reasonable agreement is observed between the present results and those by FEM (by Comsol Multiphysics). The present method has advantage in solving more complex coupled THM problems, as it solves the problems by coupling the ALEM and FEM approach and has no restriction on the number of layers and the layer thickness. Furthermore, an example is given to investigate the response of a two-layered soil with an buried cylindrical heat source, as shown in Fig. 6, The length and radius of the heat source are h and r0 respectively, the total thickness of the soil is 200h and the depth from the soil surface to the heat source top is 100h, here h ¼ 20r0 . The FEM model (by Table 2 Soil properties. Property Layer Layer 1 Layer 2 Mass density (kg/m3) Heat conduction coefficient (W/m°C) Heat capacity (J/kg°C) Thickness (mm) 1835 1.09 1657 100 1989 2.12 1558 200 3. Results and discussions This section focuses on the soil response in term of EPP and vertical displacement induced by a time-varying cylindrical heat source. Three scenarios are considered, i.e., time-varying heat source, different types of heat source and a three-layered soil with a cylindrical heat source. The temperature change and infiltration condition at the surface boundary of the model are 0 °C and permeable, the bottom boundary conditions are insulated and impermeable. Values for parameters are provided in Table 3, the effect of one parameter, the others are determined by default values. Permeability and heat conduction coefficient of the soil are 0:83 108 m=s and 1w/m°C, respectively. Poisson’s ratio is chosen as 0.25. The thickness of the soil is 1000m and the length of the cylindrical heat source is h ¼ 10m, h ¼ 20r0 , the source is in the middle of the soil. The total initial strength of the heat source is q0 . The normalized parameters of s, T N and pN are chosen as those in Section 2.3. Fig. 5. Comparison of temperature over time surrounding a cylindrical heat source in a homogeneous soil. 7 L. Wang Soils and Foundations 62 (2022) 101181 Table 3 Soil properties for parametric study. Parameters Normalized half-life of the decaying heat source, t Heat source types Number of layers 3.1. Behavior of saturated soil with a time-varying heat source Other values 0.01, 0.1 Solid cylindrical 3 0.001, 1, 10, 1 Tubular, hollow cylindrical 1 perature tends to be stable eventually for the source with constant strength (t ¼ 1), consistent with the observations made in previous works (e.g., Small and Booker, 1986; Ai and Wang, 2017). Fig. 8 illustrates that EPP achieves its peak value at early time and tends to be stable finally. It is observed that negative EPP appear at later time for the case of t ¼ 0:01, which is related to the cooling of the decaying heat source. In this subsection, the calculated point is at the middepth of the source, and r=r0 ¼ 1. Fig. 7 shows the time variation of temperature induced by decaying heat sources with different half-lives, here a normalized parameter t is given to reflect the decaying half-life. t ¼ j=xL2 Default values ð18Þ 3.2. Effect of heat source types The temperature increases to the peak before decreasing to zero in the final stage. The larger the decaying half-life is, the higher the peak value will achieve. Moreover, the tem- Figs. 9 and 10 illustrate the time variations of temperature and EPP of the point at r=r0 ¼ 1 and z=h ¼ 0:5 Fig. 7. Time variation of temperature induced by decaying heat sources with different half-life. Fig. 9. Time variation of temperature subjected to three types of heat source. Fig. 8. Time variation of EPP induced by decaying heat sources with different half-life. Fig. 10. Time variation of EPP subjected to three types of heat source. 8 L. Wang 3.3. Behaviors of layered soil with a cylindrical source induced by three types of heat source, e.g. solid cylindrical, tubular and hollow cylindrical source. Fig. 9 indicates that the curves corresponding to the three cases are quite different from each other in the initial stage s 1. When the radii are the same, the temperature variation caused by the tubular source is larger than that caused by the solid or hollow ones. Additionally, the temperature increases over time and almost unaffected by the heat source type when s > 1. From Fig. 10, it is observed that the curves of EPP increase along time firstly and then tend to converge and achieve the maximum. The curves for the three types are nearly overlapped after EPP achieve the maximum values, and the difference in EPP is quite small. It is probably attributed to the combined effect of the increase of EPP induced by the increasing temperature and the dissipation of EPP induced by seepage flow. In practical engineering, natural soils are mostly layered media owing to long periods of sedimentation process. The effect of stratification on the behavior of a typical threelayered soil system with a no-decaying cylindrical source is discussed here, the thickness relationship of three layers is h1 : h2 : h3 ¼ 3 : 4 : 3. Owing to long term geological sedimentation processes, soils are normally distributed in layers, various permeability and heat conduction coefficients in different layers should be considered, the relationships of permeability and heat conduction coefficient among the layers are provided in Table 4, other parameters can refer to Table 3. Four cases are investigated here, in case 1, the soil is homogeneous medium; in case 2, permeability coefficients are changed, while other parameters are the same; in case 3, heat conduction coefficients are changed; in case 4, both permeability coefficients and heat conduction coefficients are changed. Figs. 11 and 12 illustrate the distributions of temperature increment and EPP along depth in a three-layered soil surrounding a solid cylindrical heat source at different s, respectively. Fig. 11 shows that the temperature changes almost the same in the relatively early stage (s ¼ 0:1) for Table 4 Parameter relationships among different layers. Parameters Case 1 Case 2 Case 3 Case 4 c1:c2:c3 K1:K2:K3 1:1:1 1:1:1 1:3:9 1:1:1 1:1:1 1:1.2:1.5 1:3:9 1:1.2:1.5 Soils and Foundations 62 (2022) 101181 Fig. 11. Distribution of temperature increment along depth at r=r0 ¼ 1 in a three-layered soil. 9 L. Wang Soils and Foundations 62 (2022) 101181 Fig. 12. Distribution of EPP along depth at r=r0 ¼ 1 in a three-layered soil. of soil have significant effects both on the variations of temperature and EPP. different cases, then the differences appear and get larger with time. The curves of case 1 and case 2 are almost coincident, this is due to that the effects of permeability on temperature changes can almost be neglected. The heat conduction coefficients relationship in case 3 and case 4 are the same and the corresponding curves are well coincident, which is different from those of case 1 and case 2. It can be learned that the temperature change differences would be larger when the heat conductivity difference between the soil layers is larger. Owing to the various permeability in different layers, Fig. 12 shows sudden changes of EPP at the interfaces of adjacent layers. It is interesting that the EPP at layer 1 is obviously larger than that at layers 2 and 3 for cases 2 and 4 for the permeability of layer 1 is the smallest one. With the same permeability, the EPP of case 2 is slightly larger than that of case 4, and the EPP of case 1 is slightly larger than that of case 3. A possible reason is that extra pore pressure dissipating induced by the heat conduction results in changing of EPP. In the early stages (s ¼ 0:1; 1; 10), the EPP difference caused by the different permeability gradually increases with time, however, the EPP difference decreases, since the pore water dissipation is gradually in dominant in long term. Consequently, it can be reasonably to conclude that the layered properties 4. Summary This paper develops a simplified method to examine the temperature effect on the behavior of layered soil subjected to a time-varying cylindrical heat source. Comparisons with experimental result, semi-analytical solution and numerical result by Comsol Multiphysics are given to confirm the robustness of the method, followed by extensive parametric studies to investigate the effects of decaying half-life and types of the heat source, and the soil layered characteristics on the soil behavior. The main findings are drawn as follows. (1) Under the decaying heat source’s influence, the temperature increases to the peak at first and then decreases to zero over time. The peak value increases with the decaying half-life. EPP achieves its peak value early and then tends to be stable. Due to the cooling of the decaying heat source, the negative EPP appears at later time. 10 L. Wang Soils and Foundations 62 (2022) 101181 Fig. 13. Time-varying heat sources: (a) a constant heat source, (b) a decaying heat source. Table 5 Expressions for different types of general micro external loads. P ðn; H i ; sÞ Constant heat source Time-varying heat source Type of heat source Solid cylindrical source qm (r = r0) Tubular soure qm (r = r0) Hollow cylindrical source qm (rext = r1, rint = r2) h iT 0Þ 0; 0; 0; qmprJ 10ðnr ns2 h iT J 1 ðnr0 Þ 0; 0; 0; prqm0 nsðsþxÞ h iT 0 ðnr 0 Þ 0; 0; 0; qm J2ps 2 h iT qm r2 J 1 ðnr2 Þr1 J 1 ðnr1 Þ 0; 0; 0; pns 2 r2 r2 T m J 0 ðnr 0 Þ ½0; 0; 0; q2psðsþxÞ T qm 1 J 1 ðnr 1 Þ ½0; 0; 0; pnsðsþxÞ ðr2 J 1 ðnrr22Þr Þ r21 2 2 1 coelastic correspondence principle, the long-time dependent behaviors of clayey soil caused by creep and consolidation can be described based on the developed explicit solution. (2) The temperature change caused by a tubular heat source is larger than that caused by a solid or a hollow ones when the radii of the heat source are the same. The temperature increases over time almost unaffected by the heat source type when s > 1. The EPP first increases along time and then tends to converge and achieves the maximum. The EPP curves are nearly overlapped after achieve the maximum. (3) The layered characteristics show significant influences on the variations of temperature and EPP of the soil. A sharp gradient of EPP forms near the interface owing to the significant difference in permeability of three layers. The thermal interaction is a significant process that should be considered as refined calculations of temperature and EPP of GHE installed in layered soils. (4) This paper develops a fundamental solution for THM coupling behaviors of isotropic, poroelastic and layered soils, and several extensions can be improved in the future. Firstly, the anisotropic mechanical and thermal characteristics of natural soils can be considered by replacing the governing equations (2)–(5) with the corresponding equation of anisotropic soils. Secondly, with the aid of typical viscoelastic models (e.g., Maxwell, Merchant) and the elastic–vis- Declaration of Competing Interest The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. Acknowledgements This work was supported by the Zhejiang Provincial Natural Science Foundation of China (No. LY21E080026) and the National Natural Science Foundation of China (No. 52078458). Appendix I. The expressions for different types of P ðn; H i ; sÞ (Fig. 13) are listed in Table 5. Here qm denotes the strength of the mth heat source, x ¼ ln2=c, c is the half-life of a decay heat source. 11 L. Wang Soils and Foundations 62 (2022) 101181 Appendix II. (a) When j ¼ cðk þ 2GÞ=cw U11 ¼ U55 ¼ 2Gsdna2 b1 ðna1 d 1 d 2 g1 2Gjnd 1 d 5 f 1 d 3 y 4 Þ=‘, U12 ¼ U21 ¼ U56 ¼ U65 ¼ 2Gn2 ða1 ð2Gsdja2 b1 d 1 ðdd 4 d 5 d 2 f 2 Þ þ nd 2 l3 Þ , 2Gjnð2dd 5 f 1 g1 þ f 21 g3 þ dd 25 g4 ÞÞ=‘ U13 ¼ U31 ¼ U57 ¼ U75 ¼ 2Gdjna2 ðnðd 1 ðf 1 g1 þ d 5 g4 Þ d 3 ðdd 5 g1 þ f 1 g3 ÞÞ sa1 a2 b1 d 1 f 4 Þ=‘, U14 ¼ U58 ¼ Gnðd 4 ða1 d 2 ðd2 ng1 g2 þ l2 l4 sx1 y 3 Þ 2Gd2 jnd 5 f 1 g2 2Gdjnd 25 l2 Þþ , dð4sza1 a22 x1 y 3 na1 d 22 g1 l1 þ nd 2 ð2Gjd 5 f 1 l1 þ g2 y 4 ÞÞÞ=d‘d 2 U15 ¼ U51 ¼ 4Gsdna1 a2 b1 y 4 =‘, U16 ¼ U61 ¼ U25 ¼ U52 ¼ 4Gsdn2 a1 a2 b1 ða1 d 2 g1 2Gjd 5 f 1 Þ=‘, U17 ¼ U71 ¼ U35 ¼ U53 ¼ 4Gdjna1 a2 y 3 =‘, U18 ¼ U54 ¼ 2Gna2 ðd2 ng2 ð2Gjd 5 f 1 a1 d 2 g1 Þ þ sa1 x1 y 3 ðd 2 zdd 4 Þ þ l2 y 4 Þ=d‘d 2 , U22 ¼ U66 ¼ 2Gsna1 a2 b1 ð2Gjnf 1 f 4 þ d 2 ðdnd 1 g1 d 3 l5 ÞÞ=‘, U23 ¼ U32 ¼ U67 ¼ U76 ¼ 2Gdjna2 ðd 1 y 3 d 3 y 2 Þ=‘, U24 ¼ U68 ¼ Gnða1 d 22 ðl1 l5 dn2 g1 g2 Þ 2Gjnd 4 f 1 ðdf 1 g2 þ d 5 l2 Þ 4szda1 a22 x1 y 2 þ , d 2 ð2Gdjn2 d 5 f 1 g2 þ 2Gjnf 21 l1 þ a1 d 4 ðng1 l2 dg2 l5 þ sx1 y 2 ÞÞÞ=d‘d 2 U26 ¼ U62 ¼ 4Gsna1 a2 b1 ð2Gjnf 21 þ a1 d 2 l5 Þ=‘, U27 ¼ U72 ¼ U36 ¼ U63 ¼ 4Gdjna1 a2 y 2 =‘, U28 ¼ U64 ¼ 2Gna2 ð2Gdjnf 21 g2 þ 2Gjnd 5 f 1 l2 þ a1 ðd 2 ðdg2 l5 ng1 l2 Þ þ sx1 y 2 ðzdd 4 , d 2 ÞÞÞ=d‘d 2 U33 ¼ U77 ¼ djð2Gnðs2 a1 a22 b1 d 1 d 2 d 3 þ jnðdnd 25 g1 f 1 f 2 g1 þ d 5 ðnf 1 g3 f 2 g4 ÞÞÞ þ a1 d 4 y 1 Þ=s‘b1 , U34 ¼ U78 ¼ ð8Gs3 zdna21 a32 b1 d 1 d 2 x1 2Gna2 ðd2 jnf 4 g1 g2 þ s3 a21 b1 d 1 d 22 d 4 x1 Þ þ 4szda1 a22 x1 y 1 , d 2 ðsa1 d 4 x1 y 1 þ 2Gdjnðnf 1 g2 l4 þ l1 y 2 ÞÞ þ 2Gjnd 4 ðd2 d 5 g2 l5 l2 y 3 ÞÞ=2sd‘b1 d 2 U37 ¼ U73 ¼ 2dja1 a2 ð2Gs2 na1 a2 b1 d 1 d 2 þ y 1 Þ=s‘b1 , U38 ¼ U74 ¼ a2 ðsa1 x1 ðd 2 zdd 4 Þð2Gs2 na1 a2 b1 d 1 d 2 þ y 1 Þ þ 2Gjnðl2 y 3 d2 g2 y 2 ÞÞ=sd‘b1 d 2 , U41 ¼ U42 ¼ U43 ¼ U45 ¼ U46 ¼ U47 ¼ 0, U44 ¼ U88 ¼ Kdd 4 =sd 2 , U48 ¼ U84 ¼ 2Kda2 =sd 2 , U81 ¼ U82 ¼ U83 ¼ U85 ¼ U86 ¼ U87 ¼ 0. where d2 ¼ n2 þ s=j, b1 ¼ k þ 2G, c1 ¼ c=cw , x1 ¼ b þ b1 au , x2 ¼ b b1 au , k ¼ 2Gl=ð1 2lÞ, a1 ¼ ezn , a2 ¼ ezd , d 1 ¼ 1 a21 , d 2 ¼ 1 a22 , d 3 ¼ 1 þ a21 , d 4 ¼ 1 þ a22 , d 5 ¼ a2 d 3 a1 d 4 , f 1 ¼ na1 d 2 da2 d 1 , f 2 ¼ na2 d 1 da1 d 2 , f 3 ¼ d2 a1 d 4 n2 a2 d 3 , f 4 ¼ dd 1 d 4 nd 2 d 3 , g1 ¼ sza2 b1 d 1 þ 2Gjnd 5 , g2 ¼ sza1 d 2 x1 þ 2djb1 d 5 au , g3 ¼ sza2 b1 d 3 þ 2Gjf 2 , g4 ¼ szda2 b1 d 3 2Gjnf 1 , g5 ¼ szda1 d 4 x1 2jnb1 f 1 au , g6 ¼ sa1 d 2 x1 þ szda1 d 4 x1 þ 2djb1 f 2 au , l1 ¼ sba1 d 2 þ g5 sa1 b1 d 2 au , l2 ¼ ng5 sna1 d 2 x1 2sb1 f 1 au , l3 ¼ dg21 g3 g4 , l4 ¼ sa2 b1 d 1 ng3 , l5 ¼ sda2 b1 d 1 þ ng4 , y 1 ¼ s2 da22 b21 d 21 þ n2 l3 , y 2 ¼ nf 1 g1 þ d 5 l5 , y 3 ¼ dnd 5 g1 f 1 l4 , y 4 ¼ 2Gdjnd 25 a1 d 2 l4 , ‘ ¼ 2Gs2 na21 a2 b1 d 1 d 22 þ a1 d 2 y 1 2Gjnðdd 5 y 2 þ f 1 y 3 Þ. (b) When j–cðk þ 2GÞ=cw . U11 ¼ U55 ¼ 2Gscna2 ðnd 1 l5 d 4 y 6 Þ=‘ U12 ¼ U21 ¼ U56 ¼ U65 ¼ 2Gn2 ð2Gnwf 21 g3 þ a1 b1 d 2 ðsca2 d 1 g3 þ nl1 Þ þ sca2 d 4 l5 þ2Gcwd 7 ðnf 1 g1 þ y 3 ÞÞ=‘ U13 ¼ U31 ¼ U57 ¼ U75 ¼ 2Gcnwa2 ðd 1 y 3 þ d 4 y 5 Þ=‘, U14 ¼ U58 ¼ 2Gsnðd 6 ð2Gcnwd 7 l6 þ a1 b1 d 2 y 4 Þ þ wa1 a3 f 7 x1 y 5 þ cnd 3 ðg5 l5 þ g4 y 6 ÞÞ=ðc2 d2 Þw‘a3 d 3 U15 ¼ U51 ¼ 4Gscna1 a2 y 6 =‘, U16 ¼ U61 ¼ U25 ¼ U52 ¼ 4Gscn2 a1 a2 l5 =‘ U17 ¼ U71 ¼ U35 ¼ U53 ¼ 4Gcnwa1 a2 y 5 =‘ U18 ¼ U54 ¼ 4Gsnð2Gcnwd 7 l6 þ a1 ðb1 d 2 y 4 þ wf 6 x1 y 5 ÞÞ=ðd2 c2 Þw‘d 3 , U22 ¼ U66 ¼ 2Gsna1 a2 ð2Gnwf 1 f 8 þ b1 d 2 ðcnd 1 g1 d 4 l4 ÞÞ=‘, U23 ¼ U32 ¼ U67 ¼ U76 ¼ 2Gcnwa2 ðd 4 y 3 þ d 1 y 5 Þ=‘, U24 ¼ U68 ¼ 2Gsnðd 3 ð2Gcn2 wd 7 f 1 g4 þ 2Gnwf 21 g5 þ a1 b1 d 2 ðsca2 d 1 g5 þ nl3 ÞÞþ , d 6 ð2Gnwf 1 l6 a1 b1 d 2 y 1 Þ wa1 a3 f 7 x1 y 3 Þ=ðd2 c2 Þw‘a3 d 3 U26 ¼ U62 ¼ 4Gsna1 a2 ð2Gnwf 21 þ a1 b1 d 2 l4 Þ=‘, 12 L. Wang Soils and Foundations 62 (2022) 101181 U27 ¼ U72 ¼ U36 ¼ U63 ¼ 4Gcnwa1 a2 y 3 =‘, U28 ¼ U64 ¼ 4Gsnð2Gnwf 1 l6 a1 ðb1 d 2 y 1 þ wf 6 x1 y 3 ÞÞ=ðc2 d2 Þw‘d 3 , U33 ¼ U77 ¼ cc1 ð2Gnwðf 2 y 3 þ nd 7 y 5 Þ þ a1 b1 d 5 q1 Þ=s‘, U34 ¼ U78 ¼ ð2Gnðsc2 a2 d 1 d 3 d 7 g5 þ cnd 3 ðf 1 ðg4 l2 g1 g5 Þ þ d 7 l3 Þ d 6 ðcd 7 y 1 þ f 1 y 4 ÞÞ , þa1 a3 f 7 x1 q1 Þ=ðc2 d2 Þ‘a3 d 3 2 U37 ¼ U73 ¼ 2cwa1 a2 q1 =s‘, U38 ¼ U74 ¼ ð4Gnðcd 7 y 1 þ f 1 y 4 Þ 2a1 f 6 x1 q1 Þ=ðc2 d Þ‘d 3 , U41 ¼ U42 ¼ U43 ¼ U45 ¼ U46 ¼ U47 ¼ 0, U44 ¼ U88 ¼ Kdd 6 =sd 3 , U48 ¼ U84 ¼ 2Kda3 =sd 3 , U81 ¼ U82 ¼ U83 ¼ U85 ¼ U86 ¼ U87 ¼ 0. where c2 ¼ n2 þ s=w, d2 ¼ n2 þ s=j, w ¼ b1 c1 , b1 ¼ k þ 2G, c1 ¼ c=cw , x1 ¼ b þ b1 au , a1 ¼ ezn , a2 ¼ ezc , a3 ¼ ezd , d 1 ¼ 1 a21 , d 2 ¼ 1 a22 , d 3 ¼ 1 a23 , d 4 ¼ 1 þ a21 , d 5 ¼ 1 þ a22 , d 6 ¼ 1 þ a23 , d 7 ¼ ða2 a1 Þð1 a1 a2 Þ, d 8 ¼ ða3 a2 Þð1 a2 a3 Þ, d 9 ¼ ða3 a1 Þð1 a1 a3 Þ, f 1 ¼ ca2 d 1 na1 d 2 , f 2 ¼ na2 d 1 ca1 d 2 , f 3 ¼ da1 d 3 na3 d 1 , f 4 ¼ da3 d 1 na1 d 3 , f 5 ¼ da2 d 3 ca3 d 2 , f 6 ¼ da3 d 2 ca2 d 3 , f 7 ¼ dd 2 d 6 cd 3 d 5 , f 8 ¼ nd 2 d 4 cd 1 d 5 , g1 ¼ sza2 d 1 þ 2Gnc1 d 7 , g2 ¼ szca2 d 4 þ 2Gnc1 f 1 ,g3 ¼ sza2 d 4 þ 2Gc1 f 2 , g4 ¼ ba1 c1 d 8 wa3 d 7 au þ ja2 d 9 au , g6 ¼ bna1 c1 f 6 dwa3 f 1 au þ cja2 f 4 au , l1 ¼ cg21 g2 g3 , l2 ¼ sa2 d 1 ng3 , g5 ¼ ba1 c1 f 5 þ wa3 f 2 au þ ja2 f 3 au , l3 ¼ cng1 g4 þ g2 g5 , l4 ¼ sca2 d 1 þ ng2 , l5 ¼ 2Gwd 7 f 1 þ a1 b1 d 2 g1 , l6 ¼ df 1 g4 d 7 g6 , y 1 ¼ ng1 g6 dg4 l4 , y 2 ¼ l4 cng3 , y 3 ¼ nf 1 g1 d 7 l4 , y 4 ¼ cdng1 g4 þ g6 l2 , y 5 ¼ cnd 7 g1 þ f 1 l2 , y 6 ¼ 2Gcnwd 27 a1 b1 d 2 l2 , q1 ¼ n2 l1 þ sa2 d 1 y 2 , 2 2 ‘ ¼ 2Gnwð2cnd 7 f 1 g1 þ f 1 l2 cd 7 l4 Þ þ a1 b1 d 2 q1 . Faizal, M., Bouazza, B., McCartney, J.S., Haberfield, C., 2019. Effects of cyclic temperature variations on thermal response of an energy pile under a residential building. J. Geotech. Geoenviron. Eng., ASCE 145 (10), 04019066. Florides, G.A., Christodoulides, P., Pouloupatis, P., 2013. Single and double U-tube ground heat exchangers in multiple-layer substrates. Appl. Energy 102, 364–373. Goode III, J.C., McCartney, J.S., 2015. Centrifuge modeling of boundary restraint effects in energy foundations. J. Geotech. Geoenviron. Eng., ASCE 141 (8), 04015034. Guo, Y., Zhang, G., Liu, S., 2018. Investigation on the thermal response of full-scale PHC energy pile and ground temperature in multi-layer strata. Appl. Therm. Eng. 143, 836–848. Kalantidou, A., Tang, A.M., Pereira, J.-M., Hassen, G., 2012. Preliminary study on the mechanical behaviour of heat exchanger pile in physical model. Géotechnique 62 (11), 1047–1051. Kramer, C.A., Ghasemi-Fare, O., Basu, P., 2015. Laboratory thermal performance tests on a model heat exchanger pile in sand. Geotech. Geol. Eng. 33 (2), 253–271. Laloui, L., Nuth, M., Vulliet, L., 2006. Experimental and numerical investigations of the behaviour of a heat exchanger pile. Int. J. Numer. Anal. Meth. Geomech. 30 (8), 763–781. Li, W.X., Li, X.D., Du, R.Q., Wang, Y., Tu, J.Y., 2019. Experimental investigations of the heat load effect on heat transfer of ground heat exchangers in a layered subsurface. Geothermics 77, 75–82. Loveridge, F., McCartney, J.S., Narsilio, G.A., Sanchez, M., 2020. Energy geostructures: A review of analysis approaches, in situ testing and model scale experiments. Geomech. Energy Environ. 22 100173. Loveridge, F., Olgun, C.G., Brettmann, T., Powrie, W., 2015. The thermal behaviour of three different auger pressure grouted piles used as heat exchangers. Geotech. Geol. Eng. 33, 273–289. Luo, J., Rohn, J., Bayer, M., Priess, A., Xiang, W., 2014. Analysis on performance of borehole heat exchanger in a layered subsurface. Appl. Energy 123 (15), 55–65. Maiorano, R.M.S., Marone, G., Russo, G., Girolamo, L.D., 2019. Experimental behavior and numerical analysis of energy piles. Proceedings of the XVII ECSMGE-2019. https://doi.org/10.32075/ 17ECSMGE-2019-0819. Mei, G.X., Yin, J.H., Zai, J.M., Yin, Z.Z., Ding, X.L., Zhu, G.F., Chu, L. M., 2004. Consolidation analysis of a cross-anisotropic homogeneous elastic soil using a finite layer numerical method. Int. J. Numer. Anal. Meth. Geomech. 28 (2), 111–129. References Ai, Z.Y., Hu, Y.D., 2015. Multi-dimensional consolidation of layered poroelastic materials with anisotropic permeability and compressible fluid and solid constituents. Acta Geotech. 10, 263–273. Ai, Z.Y., Wang, L.J., 2015. Axisymmetric thermal consolidation of multilayered porous thermoelastic media due to a heat source. Int. J. Numer. Anal. Meth. Geomech. 39, 1912–1931. Ai, Z.Y., Wang, L.J., 2017. Thermal performance of stratified fluid-filled geomaterials with compressible constituents around a deep buried decaying heat source. Meccanica 52, 2769–2788. Ai, Z.Y., Wang, L.J., 2018. Precise solution to 3D coupled thermohydromechanical problems of layered transversely isotropic saturated porous media. Int. J. Geomech., ASCE 18 (1), 04017121. Ai, Z.Y., Yue, Z.Q., Tham, L.G., Yang, M., 2002. Extended Sneddon and Muki solutions for multilayered elastic materials. Int. J. Eng. Sci. 40, 1453–1483. Bao, H.Y., Hou, X.M., Yang, C., 2015. Finite layer method to calculate the dynamic compliance of a foundation. 7th International Conference on Measuring Technology and Mechatronics Automation (ICMTMA), 2015(1), 496–498.. Biot, M.A., 1956. Thermoelasticity and irreversible thermodynamics. J. Appl. Phys. 27 (3), 240–253. Booker, J.R., Savvidou, C., 1985. Consolidation around a point heat source. Int. J. Numer. Anal. Meth. Geomech. 9, 173–184. Brandl, H., 2006. Energy foundations and other thermos-active ground structures. Géotechnique 56 (2), 81–122. Chen, D., McCartney, J.S., 2016. Calibration parameters for load transfer analysis of energy piles in uniform soils. Int. J. Geomech., ASCE 17 (7), 04016159. Dupray, F., Laloui, L., Kazangba, A., 2014. Numerical analysis of seasonal heat storage in an energy pile foundation. Comput. Geotech. 55 (1), 67–77. Di Donna, A., Laloui, L., 2015. Numerical analysis of the geotechnical behaviour of energy piles. Int. J. Numer. Anal. Meth. Geomech. 39 (8), 861–888. EI-Zein, A., 2006. Laplace boundary element model for the thermoelastic consolidation of multilayered media. Int. J. Geomech., ASCE 6 (2), 136–140. Erol, S., Francois, B., 2018. Multilayer analytical model for vertical ground heat exchanger with groundwater flow. Geothermics 71, 294– 305. 13 L. Wang McCartney, J.S., Rosenberg, J.E., 2011. Impact of heat exchange on side shear in thermo-active foundations. Geo-Frontiers 2011, 488–498. Moradshahi, A., Khosravi, A., McCartney, J.S., Bouazza, A., 2020. Axial load transfer analyses of energy piles at a rock site. Geotech. Geol. Eng. 38, 4711–4733. Moradshahi, A., Faizal, M., Bouazza, A., McCartney, J.S., 2021. Effect of nearby piles and soil properties on the thermal behaviour of a fieldscale energy pile. Can. Geotech. J. 38, 4711–4733. Murphy, K.D., McCartney, J.S., 2014. Thermal borehole shear device. Geotech. Test. J., ASTM 37 (6), 1040–1055. Murphy, K.D., McCartney, J.S., 2015. Seasonal response of energy foundations during building operation. Geotech. Geol. Eng. 33, 343– 356. Ng, C.W.W., Shi, C., Gunawan, A., Laloui, L., Liu, H., 2015. Centrifuge modelling of heating effects on energy pile performance in saturated sand. Can. Geotech. J. 52 (8), 1045–1057. Olgun, C.G., Ozudogru, T.Y., Abdelaziz, S.L., Senol, A., 2014. Long-term performance of heat exchanger pile groups. Acta Geotech. 10 (5), 553– 569. Ozudogru, T.Y., Olgun, C.G., Arson, C.F., 2015. Analysis of friction induced thermo-mechanical stresses on a heat exchanger pile in isothermal soil. Geotech. Geol. Eng. 33, 357–371. Pan, A., McCartney, J.S., Lu, L., You, T., 2020. A novel analytical multilayer cylindrical heat source model for vertical ground heat exchangers installed in layered ground. Energy 200 117545. Pasten, C., Santamarina, J., 2014. Thermally induced long-term displacement of thermoactive piles. J. Geotech. Geoenviron. Eng., ASCE 140 (5), 06014003. Pan, E., 1999. Green’s functions in layered poroelastic half-spaces. Int. J. Numer. Anal. Meth. Geomech. 23 (13), 1631–1653. Rotta Loria, A.F., Gunawan, A., Shi, C., Laloui, L., Ng, C.W.W., 2015. Numerical modelling of energy piles in saturated sand subjected to thermo-mechanical loads. Geomech. Energy Environ. 1, 1–15. Savvidou, C., Booker, J.R., 1989. Consolidation around a heat source buried deep in a porous thermoelastic medium with anisotropic flow properties. Int. J. Numer. Anal. Meth. Geomech. 13 (1), 75–90. Self, S.J., Reddy, B.V., Rosen, M.A., 2013. Geothermal heat pump systems: status review and comparison with other heating options. Appl. Energy 101, 341–348. Selvadurai, A.P.S., Suvorov, A.P., 2014. Thermo-poromechanics of a fluid-filled cavity in a fluid-saturated geomaterial. Proc. R. Soc. A Math. Phys. Eng. Sci. 470 (2163), 20130634. Shahrokhabadi, S., Cao, D.T., Vahedifard, F., 2020. Thermal effects on hydromechanical response of seabed-supporting hydrocarbon pipelines. Int. J. Geomech., ASCE 20 (1), 04019143. Small, J.C., Booker, J.R., 1986. Behaviour of layered soil or rock containing a decaying heat source. Int. J. Numer. Anal. Meth. Geomech. 10 (5), 501–509. Soils and Foundations 62 (2022) 101181 Smith, D.W., Booker, J.R., 1996. Boundary element analysis of linear thermoelastic consolidation. Int. J. Numer. Anal. Meth. Geomech. 20 (7), 457–488. Sneddon, I.N., 1972. The Use of Integral Transform. McGraw-Hill, New York. Song, Z., Liang, F.Y., Chen, S.L., 2019. Thermo-osmosis and mechanocaloric couplings on THM responses of porous medium under point heat source. Comput. Geotech. 112, 93–103. Stewart, M.A., McCartney, J.S., 2014. Centrifuge modeling of soil– structure interaction in energy foundations. J. Geotech. Geoenviron. Eng., ASCE 140 (4), 04013044. Suryatriyastuti, M.E., Mroueh, H., Burlon, S., 2014. A load transfer approach for studying the cyclic behavior of thermo-active piles. Comput. Geotech. 55, 378–391. Sutman, M., Olgun, C.G., Laloui, L., 2019. Cyclic load–transfer approach for the analysis of energy piles. J. Geotech. Geoenviron. Eng., ASCE 145 (1), 04018101. Talbot, A., 1979. The accurate numerical inversion of Laplace transforms. IMA J. Appl. Math. 23 (1), 97–120. Wang, L.J., 2022. An analytical model for 3D consolidation and creep process of layered fractional viscoelastic soils considering temperature effect. Soils Found. 62 (2) 101124. Wang, L.J., Ai, Z.Y., 2018. Transient thermal response of a multilayered geomaterial subjected to a heat source. KSCE J. Civ. Eng. 22 (9), 3292–3301. Wang, L.J., Liu, X.T., Wan, L., Wang, L.H., 2020. Semianalytical solution for evaluating viscoelastic consolidation of saturated soils with overlying dry layers. Int. J. Geomech., ASCE 20 (9), 04020142. Wang, L.J., Wang, L.H., 2020. Semianalytical analysis of creep and thermal consolidation behaviors in layered saturated clays. Int. J. Geomech., ASCE 20 (4), 06020001. Wang, L.J., Zhu, B., Chen, Y.M., Kong, D.Q., Chen, R.P., 2019. Coupled consolidation and heat flow analysis of layered soils surrounding cylindrical heat sources using a precise integration technique. Int. J. Numer. Anal. Meth. Geomech. 43, 1539–1561. Wang, W., Regueiro, R., McCartney, J.S., 2015. Coupled axisymmetric thermoporo-elasto-plastic finite element analysis of energy foundation centrifuge experiments in partially saturated silt. Geotech. Geol. Eng. 33 (2), 373–388. Zagorščak, R., Sedighi, M., Thomas, H.R., 2017. Effects of thermoosmosis on hydraulic behavior of saturated clays. Int. J. Geomech., ASCE 17 (3), 04016068. Zheng, L., Deng, T., Liu, Q.J., 2021. Buckling of piles in layered soils by transfer matrix method. Int. J. Struct. Stab. Dyn. 21 (8), 2150109. Zhou, G.Q., Zhou, Y., Zhang, D.H., 2016. Analytical solutions for two pile foundation heat exchanger models in a double-layered ground. Energy 112 (1), 655–668. 14