You are on page 1of 39

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

net/publication/360124359

ScienceDirect A 3D fully thermo-hydro-mechanical coupling model for


saturated poroelastic medium-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/)

Article in Computer Methods in Applied Mechanics and Engineering · April 2022


DOI: 10.1016/j.cma.2022.114939

CITATION READS

1 363

2 authors:

Xin Cui Louis Ngai Yuen Wong


Massachusetts Institute of Technology The University of Hong Kong
8 PUBLICATIONS 62 CITATIONS 164 PUBLICATIONS 8,265 CITATIONS

SEE PROFILE SEE PROFILE

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

rock movement View project

Fracturing Mechanisms and Fracability Evaluation of Granite Considering Influences of Microstructure and Thermal Treatment View project

All content following this page was uploaded by Xin Cui on 23 April 2022.

The user has requested enhancement of the downloaded file.


Available online at www.sciencedirect.com
ScienceDirect

Comput. Methods Appl. Mech. Engrg. 394 (2022) 114939


www.elsevier.com/locate/cma

A 3D fully thermo–hydro–mechanical coupling model for saturated


poroelastic medium
Xin Cui, Louis Ngai Yuen Wong ∗
University of Hong Kong Shenzhen Institute of Research and Innovation (HKU-SIRI), Hong Kong, China
Department of Earth Sciences, the University of Hong Kong, Pokfulam, Hong Kong, China
Received 22 November 2021; received in revised form 22 February 2022; accepted 23 March 2022
Available online xxxx

Abstract
Thermo–hydro–mechanical (THM) coupling prevails in all sorts of underground activities, and numerical simulation is an
effective tool to understand THM coupling. The essence of performing THM coupling simulation lies in solving a system of
partial differential equations (PDEs) with displacements (U), pore pressure (P) and temperature (T) as the primary variables.
In previous studies, only the UPT and U-PT schemes are reported to solve the system PDEs. Another promising scheme UP-T
was rarely explored. In addition, although multiple solution schemes are theoretically possible, the performance thereof varies.
Few studies have ever compared the performance of different solution schemes. Moreover, the rock mass in hydrothermal fields
is highly permeable wherein heat advection dominates. Advection-dominated heat transfer may lead to numerical oscillation,
but how to stabilize the simulation in the 3D case was inadequately reported. In this study, we developed a new THM coupling
simulator HENGYI in the context of the finite element method. The UP-T solution scheme was verified to be robust enough
to simulate large engineering problems. The Streamline Upwind Petrov Galerkin method was generalized to 3D and applied
to HENGYI to address the advection-dominated heat transfer. The performance of the UPT, UP-T and U-PT schemes was
compared and the applicability of different solution schemes was found to be highly associated with the degree of coupling
between primary variables. Great details are provided regarding the development of HENGYI. Thus HENGYI not only bridges
the research gaps mentioned above but also provides a comprehensive understanding of THM coupling simulation techniques.
© 2022 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
Keywords: Thermo–hydro–mechanical (THM) coupling; Advective heat transfer; Streamline Upwind Petrov Galerkin Method (SUPG); Solution
scheme; HENGYI

1. Introduction
A number of engineering activities deep underground, such as exploitation of geothermal energy, disposal of
high-level nuclear wastes and steam-assisted gravity drainage for heavy oil recovery, are intimately associated with
the thermo–hydro–mechanical (THM) coupling in the saturated poroelastic medium [1–5]. Recently, the importance
of geothermal energy is particularly highlighted to march towards the goal of carbon neutrality in response to today’s
severe climate change. These sub-surface engineering activities necessitate an in-depth understanding of THM
coupling, yet THM coupling is a very challenging research topic due to its complexity. The multiphysics involved in
∗ Corresponding author at: Department of Earth Sciences, the University of Hong Kong, Pokfulam, Hong Kong, China.
E-mail address: lnywong@hku.hk (L.N.Y. Wong).

https://doi.org/10.1016/j.cma.2022.114939
0045-7825/© 2022 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http:
//creativecommons.org/licenses/by-nc-nd/4.0/).
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

THM coupling is difficult to be solved by analytical methods, therefore THM coupling research has been relying on
numerical simulations since its early stage. Aboustit et al. [6] first used variational principles to derive finite element
formulations to study thermo-elastic consolidation. In that paper, only 9 2D elements were utilized to discretize
a 7-meter high soil column. Despite the small scale of calculation, this seminal work provided the fundamental
methodology to perform the THM coupling study. Later, Noorishad et al. [7] developed a computer code ROCMAS
using a variational principle and Galerkin finite element method (FEM) to explore the potential application of
THM coupling in the rock mechanics field. Subsequently, Lewis et al. [8] extended the THM coupling spectrum
to the nonisothermal elastoplastic porous medium, which was also implemented by FEM. In 1992, to explore the
geological disposal of high-level nuclear wastes, an international collaboration project DECOVALEX was initiated
to advance the understanding of THM coupling [9]. Since then, studies on THM coupling boomed with a number of
DECOVALEX participants developing and improving their respective codes. Examples include FRACON developed
by Nguyen and Selvadurais [10] (FEM), UDEC improved by Ahola et al. [11] (discrete element method) and
THAMES developed by Sugita et al. [12] (FEM). These previous studies put substantial effort into the derivation
of governing equations, albeit to different sophistication, and the corresponding computer codes were mostly 2D.
Due to the presence of advective heat transfer, the energy conservation equation is nonlinear and the resulting
coefficient matrix is nonsymmetric. How to cope with the nonlinear and nonsymmetric issues were not adequately
reported by this time; in the limited number of studies that provided such details, the approaches adopted differed
a lot. Nguyen and Selvadurais [10] omitted advective heat transfer in the energy conservation equation to avoid
the nonlinear and nonsymmetric issues. Noorishad et al. [7] applied Crank–Nicolson approximation to linearize
the energy conservation equation, which might cost extra computational resources; a special technique reported by
Hsu and Nickell [13] was utilized to make the coefficient matrix symmetric. Lewis et al. [8] decoupled the energy
conservation equation from the stress equilibrium and mass conservation equations, and used a staggered iterative
method to address the nonlinearity; a partitioned solution procedure [14] was introduced to restore symmetry.
The development of 3D THM coupling simulators was typically represented by the linkage between TOUGH
family codes and various geomechanical solvers, such as TOUGH-FLAC3D [15], TOUGH-UDEC [16] and
TOUGH-RFPA [17]. TOUGH family codes are used to evaluate pore pressure and temperature, while the paired
geomechanical solvers are employed to calculate displacements. The Newton–Raphson iteration method was used
in TOUGH to linearize the nonlinear system equations [18]. In addition, to confront the challenges of large-
scale 3D simulations and system equations with a nonsymmetric coefficient matrix, Moridis et al. [19] enhanced
the solver package of TOUGH2, incorporating various iterative solvers to efficiently cope with sparse matrices.
Afterwards, the TOUGH-FLAC3D simulator was widely used to simulate CO2 sequestration [20–23], geothermal
exploitation [24–26], underground gas storage [27–29], injection-induced seismicity [26,30] and nuclear waste
disposal [31–33]. Meanwhile, the DECOVALEX project engaged several research teams to target 3D THM coupling
simulations [1,34], and a number of researchers also developed alternative THM coupling codes [35–37]. Another
surge of THM coupling study is catalyzed by the enhanced geothermal systems (EGS) because EGS holds a
substantial prospect to tap the giant hot dry rock energy. Effort on this aspect is exemplified by [38–44]. Upon this
time, a diversity of research and commercial codes (e.g. [45] for the latter) were developed to perform the THM
coupling simulation, and the numerical methods employed include FEM, finite volume method, finite difference
method, boundary element method, discrete element method, and the hybrid of these methods [1,33,39,42,44].
Assembling different well-accepted codes to develop a THM coupling simulator is a prevailing trend in the academic
field, and significant effort has been spent on the design of how to integrate those codes. As the THM coupling codes
become more and more complicated and hard to handle, another trend emerges that provides large sophisticated
open-source platforms to pool effort and enable object-oriented programs. Such platforms are represented by
OpenGeoSys [46], PFLOTRAN [47], MOOSE [48] and Open Porous Media initiative [49].
Although the knowledge regarding THM simulation techniques has been building up over the last two decades,
there remain some open questions. Firstly, solution schemes are still not adequately explored. The essence of
performing THM coupling simulations lies in solving a system of partial differential equations (PDEs) with
displacement (U), pore pressure (P) and temperature (T) as primary variables. Overall, two strategies are available
to solve the system PDEs: (a) monolithic and (b) staggered. The monolithic strategy solves the system PDEs
simultaneously, namely the UPT solution scheme. When the system PDEs are nonlinear, Newton–Raphson iteration
is mostly adopted to carry out linearization. The UPT solution scheme is typically represented by the OpenGeoSys
code [46]. In contrast, the staggered strategy divides the system PDEs into a number of subsets and relies on Picard
2
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

iteration to exchange data between subsets: when one subset is being solved, other subsets are held constant; after
having obtained the solution of the current subset, it is passed on to the next subset to advance the calculation.
In the context of THM coupling, two staggered solution schemes are possible: (a) U-PT and (b) UP-T, where
the hyphen segregates different subsets. The linkage between TOUGH family codes and various geomechanical
simulators typically adopts the U-PT solution scheme, while the UP-T solution scheme was rarely explored in the
3D case.
Secondly, most previous studies assumed the rock mass as low-permeable or impermeable material and thus
ignored the advective heat transfer therein. However, in hydrothermal fields, the rock mass is highly permeable,
especially after stimulations. In this case, advective heat transfer dominates. It has been well documented [50,51]
that advection-dominated heat transfer may lead to numerical instability and thus fails the conventional Galerkin
FEM. How to stabilize the simulation afflicted by advective heat transfer in the context of 3D THM coupling is
inadequately reported.
Thirdly, although different solution schemes are theoretically possible, the performance thereof varies. The 3D
THM coupling simulation is computationally expensive, and the solution scheme impacts the performance of a
simulator at a high level. Therefore, it is critical to choose a solution scheme that appropriately leverages Newton–
Raphson and/or Picard iteration to facilitate high performance. However, the performance comparison among the
UPT, UP-T and U-PT solution schemes is very rare, and factors influencing the applicability of these solution
schemes remain elusive.
Apart from the above three unexplored areas, we also identify another research gap. Integrating existing codes
is a prevailing trend in the THM coupling community. Much effort was spent on linking different existing codes
(mainly data exchange), while some critical solution techniques are left to the existing codes. In fact, developing a
simulator to carry out the 3D THM coupling simulation involves the derivation of governing equations, selection of
the numerical method(s), discretization under the prescribed numerical method framework, choosing the appropriate
element type, design of the solution scheme, determination of timestep, solvers for symmetric/nonsymmetric sparse
matrices, techniques to cope with advective heat transfer, and so on. After the development of the code, benchmark
examples should be provided to carefully verify the code before application. THM coupling features a wide spectrum
of knowledge as enumerated above. Previous studies only covered a fraction of these aspects instead of presenting
a full landscape of THM coupling simulation techniques.
In this study, we developed a new THM coupling simulator HENGYI using FEM to bridge the abovestated
research gaps. This paper has the following numerical novelties (the first three points) and one additional contribution
(the last point). (a) We explored the UP-T solution scheme in great detail and proved that the UP-T solution scheme
is robust enough to deal with large engineering problems. (b) We generalized the consistent Streamline Upwind
Petrov Galerkin (SUPG) method to the 3D case based on the 8-node element, and applied it to HENGYI to address
the advection-dominated heat transfer. (c) We compared the performance of UPT, UP-T and U-PT solution schemes
and revealed that the applicability of these solution schemes is closely related to the degree of coupling between
primary variables. (d) We presented as many details as possible regarding the development of HENGYI, from the
derivation of PDEs, to FEM discretization, to code design, to verification, to cross-comparison among different
solution schemes and to engineering applications. Therefore, HENGYI not only bridges the above research gaps
but also provides a comprehensive understanding of THM coupling simulation techniques.
The remainder of this paper is structured as follows. Section 2 details the development of HENGYI, with
Section 2.1 outlining its primary attributes, Sections 2.2 to 2.4 deriving the governing equations and Section 2.5
elaborating the solution procedure. Section 3 provides 4 verification examples. Since the UP-T solution scheme of
HENGYI consists of two major solvers, namely T SOLVER and HM SOLVER, the first two examples are used
to validate T SOLVER, with the first example focusing on heat conduction and the second on heat advection. The
third example is used to validate HM SOLVER, and the last example serves as a benchmark to test T SOLVER
and HM SOLVER working as a whole. Section 4 compares the performance of the UPT, UP-T and U-PT solution
schemes based on the 1D nonisothermal consolidation problem, which is followed by an analysis in terms of the
applicability of these solution schemes. Section 5 incorporates two typical engineering application cases to further
test the robustness of the U-PT solution scheme. The first case is associated with high-level nuclear waste disposal
using a hypothetical model proposed by DECOVALEX, and the second case is related to the exploitation of the
Yangbajing hydrothermal field in China. In particular, the effectiveness of the new 3D SUPG method is further
verified by the hydrothermal field case. Some conclusions are drawn in Section 6. In addition, since the UP-T
3
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

solution scheme is rarely reported, we tend to present as many details as possible regarding its development in the
main text. Nevertheless, the commonly used UPT and U-PT solution schemes are also available in HENGYI. For
completeness and conciseness, we present the UPT and U-PT solution schemes in Appendices A and B, respectively.

2. Model development
2.1. Outline

In this study, we consider single flow within a saturated isotropic poroelastic medium, and the small deformation
assumption applies to the solid phase. The flow rate is assumed to be sufficiently low thus local thermal equilibrium
is satisfied everywhere. The THM coupling model is governed by three partial differential equations (PDEs), namely
stress equilibrium, mass conservation and energy conservation. Displacements, pore pressure and temperature are
designated as the primary unknowns. We choose pure FEM to develop HENGYI because FEM is versatile to cope
with various problems and there are many precedents we can refer to. Another advantage of using pure FEM
rather than engaging multiple numerical methods is that there is no need to interpolate primary unknowns. The
weak forms of the PDEs were leveraged to derive the corresponding FEM formulations, and the consistent SUPG
was generalized to 3D to deal with advection-dominated heat transfer. In terms of solution schemes, the UPT,
UP-T and U-PT solution schemes are all available in HENGYI. The UP-T solution scheme is illustrated in great
detail in the main text, while the UPT and U-PT solution schemes are presented in Appendices A and B with
some conciseness. The UP-T solution scheme of HENGYI consists of two major solvers, namely T SOVLER used
to solve for temperature, and HM SOLVER used to solve for displacements and pore pressure. We adopted the
quadrilateral composite element to perform the calculation because it enables finer timesteps at the early stage
of the simulation. In addition, the estimation of the lower bound of timestep is also provided. The Conjugate
Gradient method using incomplete Cholesky factorization as the preconditioner is employed to solve the system
equations with the symmetric sparse coefficient matrix (HM SOLVER). In contrast, the Conjugate Gradient Squared
method using incomplete LU factorization as the preconditioner is adopted to solve the system equations with the
nonsymmetric sparse coefficient matrix (T SOLVER).

2.2. Thermoporoelastic mechanical model

The thermoporoelastic mechanical model is set up based on the stress equilibrium of a representative elemental
volume (REV) of the saturated poroelastic medium. For any candidate REV, the stress equilibrium under quasi-static
condition is given by
∇ ·σ +f =0 (1)
where σ is the second-order total stress tensor, and f is the body force vector per unit volume. In this paper, we will
switch back and forth between matrix and tensor designations. According to Terzaghi’s effective stress principle,
the total stress can be formulated by
σ = σ ′ − αpI (2)
where σ ′ is the effective stress, α is the Biot coefficient, p is the pore pressure, and I is an auxiliary matrix defined
as [1, 1, 1, 0, 0, 0]T . The Biot coefficient α is equal to 1.0 for soil material yet usually adopted as [52]
Ks
α =1− (3)
Kg
for rock material, where Ks and Kg are the bulk moduli of the rock matrix and mineral grains, respectively. The
volumetric thermal expansion of the solid phase can be evaluated by
ε T = −βv ∆T I (4)
where εT the second-order thermal expansivity tensor, βv is the volumetric thermal expansion coefficient and ∆T
is the temperature increment. The thermoelastic stress–strain relationship can be formulated as
σ ′ = Dε − βv K s ∆T I (5)
4
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

where D is the fourth-order elastic tensor. Under the small deformation assumption, the strain tensor ε can be
obtained by
1
ε = [(∇u) + (∇u)T ] (6)
2
where u is the displacement vector. According to Eqs. (2)–(6), Eq. (1) can be recast as
∇ · [ D∇u − βv K s ∆T I − αp I] + f = 0 (7)
which is the final PDE governing the thermoporoelastic mechanical model. Next, we will leverage its weak form
to derive the corresponding FEM approximation. The weak form of Eq. (7) can be written as

w T {∇ · [ D∇u − βv K s ∆T I − αp I] + f }dΩ = 0 (8)

where wT is the transpose of an arbitrary weight function vector, Ω is the volumetric region of interest which is
usually designated as one candidate FEM element. Isolate the first term of Eq. (8) and integrate by parts, we get
∫ ∫ ∫
w ∇ · ( D∇u)dΩ =
T
w n · ( D∇u)d S −
T
∇w T D∇udΩ (9)
Ω ∂Ω Ω
where n is the outward normal vector to the boundary of Ω (∂Ω ). The stress boundary condition σ b = n · ( D∇u)
is prescribed and thus can be directly applied to Eq. (9). In the framework of FEM, the primary unknowns, namely
displacements u, pore water pressure p and temperature T, are situated at the element nodes, and all these unknowns
in the interior of a candidate element can be interpolated by
u = N{u}, p = N{ p}, T = N{T } (10)
where {·} indicates the respective unknowns at the element nodes, and N and N are shape functions
⎡ ⎤
N1 0 0 N2 0 0 N3 0 0

⎦,
⎢ ⎥
N=⎢ ⎣0 N1 0 0 N2 0 0 N3 0 · · ·⎥ (11)
0 0 N1 0 0 N2 0 0 N3
[ ]
N = N1 N2 N3 · · · (12)
with Ni being readily found in any FEM textbook. From now on we change the primary unknowns to their
incremental form (∆u, ∆ p and ∆T ) in order to keep consistent with the mass and energy conservation equations
presented in Sections 2.3 and 2.4. Substitute Eqs. (9)–(10) into Eq. (8), one can get
∫ ∫ ∫
∇w T D∇ NdΩ {u(t−∆t) + ∆u} + βv w T I NdΩ {∆T } + αw T I NdΩ { p (t−∆t) + ∆ p}
Ω ∫ ∫ Ω Ω

= w∂Ω σ b ds +
T
w f dΩ
T
(13)
∂Ω Ω
(t−∆t) (t−∆t)
where u and p are the displacements and pore pressure of the last time step, respectively. If we adopt
the weight function being the same as the shape function (i.e. w = N), the so-called Galerkin FEM emerges and
Eq. (13) can be reformulated as
K uu {∆u} + K up {∆ p} + K uT {∆T } = F uu + F up + F ub + F u f = ∆F u (14)
where ∫ ∫ ∫
K uu = ∇ N T D∇ NdΩ , K up = α N T I NdΩ , K uT = βv N T I NdΩ ,
Ω Ω Ω
∫ ∫ ∫
F uu =− T
∇ N D∇ NdΩ {u (t−∆t)
}, F up = − α N I NdΩ { p (t−∆t) }, F ub =
T T
N ∂Ω σ b d S,
Ω Ω ∂Ω
⎤T (15)
∂/∂ x ∂/∂z ∂/∂ y

∫ 0 0 0
N T f dΩ , ∇ = ⎢ ∂/∂ y ∂/∂z ∂/∂ x ⎥
⎢ ⎥
Fu f = ⎣ 0 0 0 ⎦

0 0 ∂/∂z ∂/∂ y ∂/∂ x 0
5
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

2.3. Baroclinic flow model

The mass conservation of fluid in the Eulerian coordinate system reads



(φρ f ) + ∇ · (φρ f v f ) + δ(x − x 1 )ρ f q f s = 0 (16)
∂t
where φ is the porosity of the poroelastic medium, ρ f is the density of the fluid, v f is the flow velocity vector, x 1
is the location of the source, δ(·) is the Dirac function, and q f s is the flow rate of the source. The relative velocity
of the fluid (vr ) with respect to that of the solid phase (v s ) can be evaluated by
vr = v f − v s (17)
Substitute Eq. (17) into Eq. (16) yields

(φρ f ) + v s · ∇(φρ f ) + φρ f ∇ · v s + ∇ · (φρ f vr ) + δ(x − x 1 )ρ f q f s = 0 (18)
∂t
Introducing the material derivative
D ∂
(·) = (·) + v s · ∇(·) (19)
Dt ∂t
to convert Eq. (18) from the Eulerian coordinate system to the Lagrange coordinate system, one can get
D
(φρ f ) + φρ f ∇ · v s + ∇ · (φρ f vr ) + δ(x − x 1 )ρ f q f s = 0 (20)
Dt
Similarly, the mass conservation for the solid phase is given by

[(1 − φ)ρs ] + ∇ · [(1 − φ)ρs v s ] = 0 (21)
∂t
where ρs is the density of the solid phase. Expanding the second term above and applying the material derivative
yields
D
[(1 − φ)ρs ] + (1 − φ)ρs ∇ · v s = 0 (22)
Dt
Since ρs is much less sensitive to temperature and pressure as compared with ρ f , we assume ρs being constant
in this study. Therefore, Eq. (22) can be further manipulated as
1 Dφ
= ∇ · vs (23)
1 − φ Dt
Under the small deformation assumption, ∇ · v s can be related to the rate of volumetric strain
∂ε ∂
∇ · vs = = [∇ · (αu)] (24)
∂t ∂t
Substitution of Eqs. (23)–(24) into Eq. (20) leads to
Dρ f ∂
φ + ρf [∇ · (αu)] + ∇ · (φρ f vr ) + δ(x − x 1 )ρ f q f s = 0 (25)
Dt ∂t
For a baroclinic fluid, its density is a function of both temperature and pressure. We assume small changes in
density, thus the truncated Taylor expansion can be invoked to approximate fluid density with a reference value ρ0 ,
namely
∂ρ f ∂ρ f
ρ f = ρ0 + ( )0 ( p − p 0 ) − ( )0 (T − T0 ) (26)
∂p ∂T
The coefficients present in Eq. (26) can be defined as
(∂ρ f /∂ p)0 (∂ρ f /∂ T )0
βp = , βT = (27)
ρ0 ρ0
where β p and βT are the coefficients of fluid compressibility and volumetric thermal expansion, respectively.
Substitution of Eq. (27) into the first term of Eq. (25) yields
∂p ∂T
( ) [ ]
Dρ f Dp DT
φ = φρ0 β p − βT = φρ0 β p ( + v s · ∇ p) − βT ( + v s · ∇T ) (28)
Dt Dt Dt ∂t ∂t
6
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

In terms of the relative flow velocity vector vr , Darcy’s law correlates it with the gradient of fluid pressure, given
by
k
φvr = − (∇ p − ρ f g) (29)
µf
where k is the intrinsic permeability tensor, µ f is the dynamic viscosity of the fluid and g is the gravitational
acceleration vector. Substitution of Eqs. (28)–(29) into Eq. (25) and neglecting the terms v s · ∇ p and v s · ∇T lead
to
∂p ∂T ρf k ρf qfs αρ f ∂
( ) [ ]
φ βp − βT −∇ · (∇ p − ρ f g) + δ(x − x 1 ) =− (∇ · u) (30)
∂t ∂t µ f ρ0 ρ0 ρ0 ∂t
which is the PDE controlling the flow model. As with Section 2.2, we hereby derive the FEM approximation of
Eq. (30) based on its weak form. The weak form of Eq. (30) can be written as
∂p ∂T ρf k ρf qfs αρ f ∂
∫ { ( ) [ ] }
wT φ β p − βT −∇ · (∇ p − ρ f g) + δ(x − x 1 ) + (∇ · u) dΩ = 0 (31)
Ω ∂t ∂t µ f ρ0 ρ0 ρ0 ∂t
Isolate the second term on the left-hand side and integrate by parts, we can get
ρf k ρf k ρf g
∫ [ ] ∫ [ ] ∫
w ∇·
T
(∇ p − ρ f g) dΩ = w∂Ω n·
T
(∇ p − ρ f g) d S− ∇w T k(∇ p − ρ f g) dΩ
[ ]
Ω µ f ρ0 ∂Ω µ f ρ0 Ω µ f ρ0
(32)
where the term
ρf k
[ ]
φv b = n · (∇ p − ρ f g) (33)
µ f ρ0
indicates the prescribed flow rate boundary condition. For temporal discretization, we use a finite difference scheme
given by
∂X X t − X (t−∆t) (t−∆t) (t−∆t)
= , X = θ X t + (1 − θ )X = θ ∆X + X (34)
∂t ∆t
where ∆t is the timestep and θ is an integration parameter between 0 and 1. For the forward difference, θ = 0;
for the backward difference, θ = 1; for the central difference, θ = 0.5; and for the Galerkin difference, θ = 0.667.
θ = 0 leads to the explicit method while θ > 0 gives rise to the implicit method. θ is a key parameter for transient
problems because it profoundly influences the numerical stability thereof. θ ≥ 0.5 ensures unconditional numerical
stability [53], thus the timestep can be sufficiently large unless constrained by the modeling accuracy. Substitution
of Eqs. (32)–(34) into Eq. (31) and using the incremental form of the primary unknowns, we can get
ρf
∫ ( ) ∫ ∫
∆p ∆T
φw β p
T
− βT w ∂Ω φv b d S +
T
∇w T k ∇ θ ∆ p + p (t−∆t) − ρ f g dΩ
{ ( ) }
dΩ −
∆t ∆t µ ρ

∫ ∂Ω Ω f 0
ρf qfs ρ f α T ∇ · (∆u)

+ w δ(x − x 1 )
T
dΩ + w dΩ = 0 (35)
Ω ρ0 Ω ρ0 ∆t
Collecting terms, Eq. (35) becomes
αρ f T
∫ ∫ ∫ ∫
∆tθρ f
φβ p w T ∆ pdΩ − φβT w T ∆T dΩ + ∇w T k∇(∆ p)dΩ + w ∇ · (∆u)dΩ
Ω Ω Ω µ ρ
f 0 Ω ρ0
∆tρ 2f
∫ ∫ ∫
∆tρ f
= T
∆tw ∂Ω φv b d S + ∇w T k∇ p (t−∆t) dΩ + ∇w T kgdΩ
∂Ω Ω µ f ρ0 Ω µ f ρ0

∆tρ f q f s T
+ w δ(x − x 1 )dΩ (36)
Ω ρ0
With the weight function w being the same as the shape function N, Eq. (36) can be recast as
K pu {∆u} + (K pp1 + θ ∆t K pp2 ){∆ p} + K pT {∆T } = ∆F p = ∆t F pb + F pp2 + F pg + F pq
( )
(37)
7
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

where ∫
αρ f T

T

ρf T
K pu = N CdΩ , K pp1 = φβ p N NdΩ , K pp2 = ∇ N k∇ NdΩ ,
ρ µ ρ

∫ 0 Ω
∫Ω f 0
ρf

T T T
K pT = − φβT N NdΩ , F pb = N ∂Ω φv b d S, F pp2 = ∇ N k∇ NdΩ { p}(t−∆t) ,
Ω ∂Ω Ω µ f ρ0
ρ 2f ρf qfs T (38)
∫ ∫
T
F pg = ∇ N kgdΩ , F pq = N δ(x − x 1 )dΩ ,
Ω µ f ρ 0 Ω ρ0
[ ]T [ ]
∇ = ∂∂x , ∂∂y , ∂z

, C = ∂∂Nx1 , ∂∂Ny1 , ∂∂z
N1 ∂ N2 ∂ N2 ∂ N2
, ∂ x , ∂ y , ∂z , . . .

2.4. Convective heat transfer model

2.4.1. Governing equation


Convective heat transfer consists of both heat conduction and heat advection, and the former is governed by
Fourier’s law. We assume the flow velocity is sufficiently low such that the solid phase and liquid phase are always
in a local thermal equilibrium state. Consequently, the absorbed/dispersed heat due to convective heat transfer for
any REV can be formulated as
qhc = −∇ · (λm ∇T ) + ρ f c f v r ∇T (39)
where λm is the average thermal conductivity tensor of the poroelastic medium. λm can be evaluated by
λm = φλ f + (1 − φ)λs (40)
where λ f and λs are the thermal conductivity tensors of the fluid and solid, respectively. The heat increment due
to the thermal power caused by volumetric deformation can be expressed by [42,54]

qhu = βv K s T (∇ · u) (41)
∂t
for the solid, and
∂p
qhp = φβ p T (42)
∂t
for the fluid, respectively. The change in the thermal storage can be evaluated by
∂T
qhp = (ρc)m (43)
∂t
where
(ρc)m = φρ f c f + (1 − φ)ρs cs (44)
with c f /cs being the specific heat of the fluid/solid. Engaging the heat source qhs and assembling Eqs. ((39),
(41)–(43)), we can get the energy conservation equation as below
∂T ∂p ∂
(ρc)m − φβ p T − βv K s T (∇ · u) − ρ f c f v r ∇T + ∇ · (λm ∇T ) + δ(x − x ′1 )qhs = 0 (45)
∂t ∂t ∂t
where x ′1 is the location of the heat source.

2.4.2. Discretization with consistent SUPG method


The conventional Galerkin FEM, when applied to mechanical, seepage and heat conduction problems, gives
rise to symmetric coefficient matrices. In this case, the difference between the FEM result and the exact solution is
minimized, which leads to the so-called “best approximation” [50]. The wide and successful application of Galerkin
FEM to the abovementioned problems is largely due to this attribute. However, when dealing with advective heat
transfer, nonsymmetric coefficient matrices arise and Galerkin FEM suffers from spurious oscillation [50,51]. The
cause of spurious oscillation lies in that the upstream node physically takes more weight, yet it is equally treated
as the downstream node under the Galerkin FEM framework. The SUPG method corrects the shape function along
the streamline direction to give more weight to the upstream node (Fig. 1), and it uses the corrected shape function
8
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Fig. 1. The weight functions of SUPG in (a) continuous and (b) discontinuous cases. A heavier weight is put to the upstream node in the
SUPG context, while the Galerkin FEM treats the upstream and downstream nodes equally.

as the weight function to carry out the FEM approximation. Since the weight function applied to the weak form of
a PDE is arbitrary, such manipulation not only achieves an upwind scheme in the FEM context, but also is rigorous
in theory. There are a limited number of previous studies reporting the application of SUPG to 2D problems in the
rock mechanics field, mostly simulating the advective heat transfer in fractures [38,39,51]. To the best knowledge
of the authors, the generalization of SUPG to 3D cases has not been reported. In this section, we will fill such a
gap based on Eq. (45).
In the context of SUPG, the weak form of Eq. (45) can be written as

(w T + τ ∇w T vrT )

∂T ∂p ∂
[ ]
× (ρc)m − φβ p T − βv K s T (∇ · u) − ρ f c f v r ∇T + ∇ · (λm ∇T ) + δ(x − x 1 )qhs dΩ = 0 (46)

∂t ∂t ∂t

where (w T + τ ∇w T vrT ) is the weight function and the second term in the parentheses indicates the upwind
correction [50]. Isolating the heat conduction term and integrating by parts, we can get
∫ ∫ ∫
(w +τ ∇w vr )∇ ·(λm ∇T )dΩ =
T T T
(w +τ ∇w vr )n·(λm ∇T )dS− ∇(w T +τ ∇w T vrT )λm ∇T dΩ (47)
T T T
Ω ∂Ω Ω

where q b = n·(λm ∇T ) is the prescribed heat flux boundary condition. Regarding the linear quadrilateral element
adopted for T SOLVER in this study, Eq. (47) can be further simplified as
∫ ∫ ∫
(w + τ ∇w vr )∇ · (λm ∇T )dΩ =
T T T
(w + τ ∇w vr )n · (λm ∇T )dS −
T T T
∇w T λm ∇T dΩ (48)
Ω ∂Ω Ω

Readers may notice from Eq. (48) that SUPG does not influence the heat conduction part, if linear elements are
used. Implementing the finite difference method elaborated in Section 2.3 to perform the temporal discretization,
Eq. (46) can be reformulated as
∫ ∫
∆T [ ] ∆p
(ρc)m (w T + τ ∇w T vrT ) dΩ − φβ p (w T + τ ∇w T vrT ) θ ∆T + T (t−∆t) dΩ
Ω∫ ∆t Ω ∫ ∆t
[ ] ∇ · (∆u) [ ]
− βv K s (w T + τ ∇w T vrT ) θ ∆T + T (t−∆t) dΩ − ρ f c f (w T + τ ∇w T vrT )vr ∇ θ ∆T + T (t−∆t) d Ω
∫Ω ∆t
∫ [ ]Ω ∫
+ (w + τ ∇w vr )q b dS −
T T T
∇w λm ∇ θ ∆T + T
T (t−∆t)
dΩ + (w T + τ ∇w T vrT )δ(x − x ′1 )qhs d Ω = 0
∂Ω Ω Ω
(49)

Let w = N and collect terms, we get

θ K T u (∆u) + θ K T p1 (∆ p) + θ ∆t K T p2 (∆ p) + θ∆t K T T1 + K T T2 (∆ p) {∆T } = ∆F T


[ ]

= F T u + F T p1 + ∆t(F T p2 + F T b + F T T + F T q ) (50)
9
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

where ∫ ∫
T T T T
K T u (∆u) = − βv K s (N + τ ∇ N vrT )∇ · (∆u)Nd Ω , K T p1 (∆ p) = − θ φβ p (N + τ ∇ N vrT )∆ p Nd Ω
∫Ω ∫ Ω ,
T T T T
K T p2 (∆ p) = − ρ f c f (N + τ ∇ N vr )vr ∇ Nd Ω , K T T1 = − ∇ N λm ∇ Nd Ω
∫ Ω ∫ Ω ,
T T T T T T
K T T2 (∆ p) = (ρc)m (N + τ ∇ N vr )Nd Ω , F T u = βv K s (N + τ ∇ N vr )T (t−∆t)
∇ · (∆u)d Ω ,
∫ Ω Ω∫
T T T T
F T p1 = φβ p (N + τ ∇ N vrT )T (t−∆t) ∆ pd Ω , F T p2 = ρ f c f (N + τ ∇ N vrT )vr ∇T (t−∆t) d Ω ,
Ω∫ ∫ Ω
T T T
FT b = − (N + τ ∇ N vrT )q b dS, FT T = − ∇ N λm ∇T (t−∆t) d Ω ,
∫∂ Ω Ω
T T
F T q = − (N + τ ∇ N vrT )δ(x − x ′1 )qhs d Ω

(51)

The next critical step is to determine τ . The designation of τ is problem-specific and can significantly influence
the stability and accuracy of the solution. The Peclet number (Pe), flow velocity and mesh size can all contribute
to the determination of τ . In the 1D case, Pe is defined as
vr h e
Pe = (52)
2λm
where h e is the length of a linear element. Pe quantifies the ratio of advection to conduction, and a large Pe generally
indicates a more advection-dominated situation. In the 1D steady-state case, it has been verified that when τ being
used as
k̃ vr h e ξ
τ= 2
, k̃ = , ξ = coth(Pe) − 1/Pe (53)
|vr | 2
the numerical result is nodally exact as compared with the analytical solution [55]. Such a conclusion was
generalized to the 2D case by Brooks and Hughes [50] as
k̃ 1
τ= 2
, k̃ = (ξ vξ h ξ + ηvη h η ), (54)
|vr | 2
where
ξ = coth(Peξ ) − 1/Peξ , η = coth(Peη ) − 1/Peη
Peξ = vξ h ξ /2λm , Peη = vη h η /2λm (55)
vξ = e ξ · v r , vη = e η · v r
The distances h ξ & h η above and the direction unit vectors eξ & eη are shown in Fig. 2(a). The axes of ξ and
η are obtained by connecting the midpoints of the paired sides of the 2D quadrilateral element (Fig. 2(a)), and the
velocity vector is evaluated at the 4 points of Gaussian integration (A–D). Similarly, we can generalize Eqs. (54)
and (55) to the 3D case, which may be formulated as
k̃ 1
τ= 2
, k̃ = (ξ vξ h ξ + ηvη h η + ηvζ h ζ ), (56)
|vr | 2
with
ξ = coth(Peξ ) − 1/Peξ , η = coth(Peη ) − 1/Peη , ζ = coth(Peζ ) − 1/Peζ ,
Peξ = vξ h ξ /2λm , Peη = vη h η /2λm , Peζ = vζ h ζ /2λm (57)
vξ = e ξ · v r , vη = e η · v r , vζ = eζ · vr
The parameters of Eq. (57) are illustrated in Fig. 2(b). Note that the distances h ξ , h η and h ζ are obtained by
connecting the centroids of the respective pairs of faces. The velocity vector is evaluated at the 8 points of Gaussian
integration in our 3D model.
10
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Fig. 2. Schematic of parameters to evaluate directional Peclet numbers in (a) 2D (after [50]) and (b) 3D cases (this study) using linear
elements.

2.5. Solution procedure

2.5.1. Solution schemes


With Eqs. (14), (37) and (50) and the corresponding initial and boundary conditions, the 3D fully THM coupling
problem in the saturated poroelastic medium can be posed as
Governing Equations:
K uu {∆u} + K up ∆ p + K uT {∆T } = ∆F u
K pu {∆u} + (K pp1 + θ∆t K pp2 ){∆ p} + K pT {∆T } = ∆F p
θ K T u (∆u) + θ K T p1 (∆ p) + θ ∆t K T p2 (∆ p) + θ ∆t K T T1 + K T T2 (∆ p) {∆T } = ∆F T
[ ]

Initial conditions: (58)


T = T 0, p = p0
Boundary conditions:
Dirichlet type: u = u′ , p = p′ , T = T′
Neumann type: σ b = σ ′b , v b = v ′b , q b = q ′b
The energy conservation equation above features the nonlinear relationship of ∆T &∆u and ∆T &∆ p. If the UPT
scheme is applied to solve Eq. (58), the Newton–Raphson iteration method is involved to address the nonlinearity
(see Appendix A). Alternatively, if the staggered strategy is adopted, two solution schemes U-PT and U-PT emerge.
Picard iteration is required for both to exchange data between the subsets, while the U-PT scheme still needs
Newton–Raphson interaction nested in Picard iteration to cope with the nonlinear relationship between ∆T and
∆ p (see Appendix B). In contrast, both the subsets of the UP-T solutions scheme are linear, thus Newton–Raphson
iteration is not engaged. Under the UP-T solution scheme (Fig. 3), the ∆T of nth Picard iteration can be first
obtained from the energy conservation equation with the inputs of ∆u and ∆ p of (n − 1)th Picard iteration. The
resolved ∆T is then fed to the stress equilibrium and mass conservation equations to update ∆u and ∆ p, based on
which the (n + 1)th Picard iteration starts. Upon the end of one Picard iteration, the primary variable (represented
by X) is updated by
X = (1 − ω)X last + ωX (59)
before marching to the next iteration, where X last denotes the primary variable of the last iteration and ω is a weight
factor. Picard iteration terminates when ∆u, ∆ p and ∆T all meet the prescribed error tolerance, from which the
calculation goes to the next step. In such a way, ∆T is decoupled from ∆ p and ∆T , and the nonlinear issue is
naturally avoided.
11
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Fig. 3. (a) Physical interpretation of the coupling between displacements, pore pressure and temperature. (b) Schematic of Picard iteration
within one step under the UP-T solution scheme.

Fig. 4. Composite elements for (a) 2D and (b) 3D cases. The displacements are interpolated quadratically while the pore pressure and
temperature are interpolated linearly.

2.5.2. Composite element


Under the present UP-T solution scheme, displacements and pore pressure are solved simultaneously. Pore
pressure is equivalent to stress, with the latter being the first-order derivative of displacements. Therefore,
displacements are one order higher than pore pressure in terms of dimension. It was discovered in previous studies
that if the linear shape function is used to interpolate both displacements and pore pressure, numerical oscillation
may take place when a small timestep is applied [6,56]. To alleviate such a drawback, a 2D composite element was
recommended by Aboustit et al. [6] (Fig. 4(a)). The 2D composite element uses a quadratic interpolation to evaluate
displacements, while engaging pore pressure and temperature only at the corner nodes of the element to achieve
a linear interpolation. Since the composite element unifies the dimensions of displacements and pore pressure, it
enables a smaller timestep as compared with that of using linear interpolation for both displacements and pore
pressure. This is preferred in scenarios where the early stage of a simulation is the key of interest, yet the extra
computational cost is required. In this study, we generalize the 2D composite element to the 3D case by hybridizing
the 20-node quadratic element for displacements and the 8-node linear element for pore pressure and temperature
(Fig. 4(b)).

2.5.3. Determination of timestep


The timestep is critically important for transient problems because it significantly influences the stability,
efficiency and accuracy of the simulation. As mentioned in Section 2.3, θ ≥ 0.5 can guarantee unconditional
numerical stability, thus timesteps can be sufficiently large without incurring numerical oscillations. However, large
timesteps reduce simulation accuracy, especially at the early stage of the simulation. In contrast, small timesteps can
improve accuracy, but increase computational cost and may trigger numerical instability [6,56]. Generally, small
timesteps are preferable at the early stage of the simulation to ensure accuracy, and they can be gradually enlarged
afterwards to advance model efficiency. In this study, θ ≥ 0.5 is adopted thus the upper bound of the timestep is
not a key issue; attention is paid to determining the minimum timestep to avoid numerical oscillation. A criterion
12
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

to determine the minimum timestep was reported [53] based on the FEM approximation of the 1D consolidation
equation using linear elements. In the case of an incompressible fluid, such a criterion reads
(∆h)2 γ f
∆t ≥ , (60)
6θ km E
where ∆h is the element size, γ f is the specific weight of the fluid, km is the permeability and E is Young’s
modulus. Subsequently, a similar study [56] enriched the above criterion for different transient problems using
various elements. According to Cui et al. [56], for the 1D advection–conduction equation using linear elements, the
minimum timestep should satisfy
(∆h)2 (ρc)m
∆t ≥ (61)
3θ λm (2 + Pe)
and for the consolidation problem using the composite element, the corresponding criterion is
(∆h)2 γ f (1 + µ)(1 − 2µ)
∆t ≥ (62)
6θ km (1 − µ)E
The generalization of Eqs. (61) and (62) to multi-dimensional cases have not been reported to the best knowledge
of the authors. Nevertheless, we can rely on the maximum value of Eqs. (61) and (62) to roughly estimate the
minimum timestep and then adjust by trial and error.

2.5.4. Solve linear system equations


The linear system equations arising from stress equilibrium and mass conservation are associated with a
symmetric coefficient matrix, while the SUPG method used to tackle the heat advection in the energy conservation
equation leads to a nonsymmetric coefficient matrix. Direct solvers based on LU factorization are robust to solve
linear system equations with both symmetric and nonsymmetric coefficient matrices. Nevertheless, the degrees of
freedom (DOFs) in the 3D case is fairly large, rendering the direct solvers of too low efficiency to be used. Therefore,
iterative methods are necessary to be engaged to solve the large-scale sparse linear system equations. Solving sparse
linear system equations with iterative methods is a relatively mature field. The conjugate gradient (CG) method
is commonly used to solve the linear system equations with the symmetric coefficient matrix. In terms of the
nonsymmetric coefficient matrix case, a number of iterative methods are also available, such as the Generalized
Minimum Residual (GMRES) method [57], Bi-Conjugate Gradient method [58] and its variants (such as CGS [59],
Bi-CGSTAB [60], and BiCGStab(l) [61]). Readers are referred to Barret et al. [62] for a relatively comprehensive
review of the iterative methods for the solution of linear system equations. In addition, preconditioning is also a
key technique to accelerate convergence for iterative solvers. In this study, we used the CG method with incomplete
Cholesky factorization as the preconditioner [63] to solve for the displacements and pore pressure, and utilized the
CGS method with incomplete LU factorization as the preconditioner to solve for the temperature.

2.5.5. Model assembly


With the FEM approximation and corresponding solution strategy ready, we can now assemble all the components
stated above to solve the 3D THM coupling problem. HENGYI was developed in the Fortran language and
in a modular manner. The module structure and pseudo-codes of the UP-T solution scheme are shown in
Fig. 5. Algorithm 1 (MAIN) is the main program, which calls algorithms 2 (INPUT), 3 (HM SOLVER) and 4
(T SOLVER). INPUT is used to handle a large number of inputs, which are then passed on to HM SOLVER and
T SOLVER. The Picard iteration between HM SOLVER and T SOLVER is controlled by the main program, and
HM SOLVER/T SOLVER calls PCG/PCGS to solve the linear system equations.

3. Model verification
3.1. Geothermal heat pump

This example is used to verify T SOLVER in simulating heat conduction. A vertical cylindrical hole is drilled to
accommodate a heat pipe within which working fluid circulates to absorb/disperse heat from/to the ambient soil. The
diameter of the hole is a bit larger than that of the pipe, thus backfill material is utilized to fill the gap (Fig. 6(a)).
13
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Fig. 5. Module structure and pseudo-codes of the UP-T solution scheme in HENGYI.

Fig. 6. (a) Schematic of the geothermal heat pump model and (b) temperature distributions evaluated by T_SOLVER as compared with the
analytical solutions [64].

A heat flux boundary condition is applied on the surface of the pipe to simulate the heat exchange between the
working fluid and backfill. The analytical solution to such a problem was derived by Gu and O’neal [64], and
because the corresponding expressions are very lengthy, readers are referred to the original paper for details. The
14
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Table 1
Inputs of the geothermal heat pump model.
Pipe radius (ra ) 0.013 m
Backfill radius (rb ) 0.074 m
Far-field boundary radius (rc ) 11.800 m
Backfill density 1600 kg/m3
Backfill thermal conductivity 5.19 W/(m◦ C)
Backfill specific heat 3342 J/(kg◦ C)
Backfill initial temperature 0 ◦C
Soil density 1600 kg/m3
Soil specific heat 1114 J/(kg◦ C)
Soil thermal conductivity 1.73 W/(m◦ C)
Soil initial temperature 0 ◦C
Surface heat flux 162.8 W/m2
θ 0.875

Fig. 7. Problem statement of the rotating cosine hill.

model parameters adopted in this example are listed in Table 1, and Fig. 6(b) compares the results of T SOLVER
with the analytical solutions. Note the horizontal and vertical axes of Fig. 6(b) have been normalized. Since the
thermal properties of the backfill and host soil differ, the slope of the temperature manifests an abrupt change at
the backfill–soil interface. After one day, T SOLVER slightly underestimates the temperature within the backfill as
compared with the analytical solution; while one month/year later, the two results show good agreement with each
other.

3.2. Rotating cosine hill

This example is used to verify T SOLVER in simulating heat advection. The definition of the rotating cosine
hill problem is illustrated in Fig. 7 and it is a good benchmark to test advection algorithms [50,65]. The governing
equation and the corresponding initial and boundary conditions of such a problem are given by
∂T
+ vr · ∇T = 0, Ω ∈ [−1, 1] × [−1, 1], t ∈ [0, 2π ]
∂t
Initial condition:
0.5 cos(4πr ) + 1 r < 0.25
{ }

T (x, y) = where r = (x + 0.5)2 + y 2 (63)
0 else
Boundary condition:
T =0 at all boundaries
15
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Fig. 8. Cosine hill after one rotation simulated by (a) the conventional Galerkin FEM and (b) SUPG method.

The initial temperature field is zero except for the cosine hill centered at (−0.5, 0) with a height of 1.0 and
a diameter of 0.5. The velocity field is specified as vr = −yi + x j , which rotates around the origin in a
counterclockwise direction. This problem is purely advective and the exact solution should be a cosine hill rotating
around the origin, without changing its shape. We adopted a 50 (length) × 50 (width) × 1(height) 3D mesh to
simulate such a problem. An invariant timestep π/100 is used and θ is prescribed as 0.5. The results of the
conventional Galerkin FEM and SUPG method are compared in Fig. 8 after one complete rotation. The conventional
Galerkin FEM shows the trailing waves, which are signs of spurious oscillation. In contrast, the SUPG method can
effectively lay to rest the trailing waves.

3.3. Terzaghi 1D consolidation

This example is used to verify HM SOLVER. Terzaghi 1D consolidation is a very classic problem that can be
used as a benchmark to test HM coupling algorithms. As shown in Fig. 9, a saturated soil column of height H is
impermeable at the bottom but free to drain on the top. At t = 0 a constant load σ0 is exerted to the top surface
of the soil column. Pore pressure escalates as a result of σ0 , yet it dissipates gradually, leading to the settlement of
the soil column. The analytical solutions of the excess pore pressure and settlement at an arbitrary depth z within
the soil column are formulated by [66]
∞ ( )
∑ 2σ0 Mz
p= sin exp(−M 2 Tv )
m=0
M H
(64)
[ ∞
]
σ0 H ∑ 2 Mz 2
uz = (H − z) 1 − cos( ) exp(−M Tv )
Es H − z m=0 M 2 H
where
Cv t km E s E(1 − µ) 1
Tv = , Cv = , Es = , M= π (2m + 1), m = 0, 1, 2, . . . (65)
H 2 ρf g (1 + µ)(1 − 2µ) 2
with E and µ being Young’s modulus and Poisson’s ratio, respectively. Note that z = 0 is defined at the top surface
of the soil column and the positive direction of the z-axis points downward (Fig. 9(a)). We referred to the 2D
meshing strategy reported by [6] to discretize the soil column and performed the 3D HM coupling simulation. Such
a model will be used again in Section 3.4 to engage temperature and verify the T SOLVER and HM SOLVER
working as a whole. The dimensions of the soil column are 2 m × 2 m × 7 m, and it meshes into 9 elements
that are progressively finer near the top surface (see the inset of Fig. 9(a)). The soil column is impermeable at all
boundaries except for the top surface. The timesteps reported by [6] were as follows: 10 steps of ∆t = 0.01 over [0,
0.1], 10 steps of ∆t = 0.1 over [0.1, 1.1], 10 steps of ∆t = 10 over [1.1, 101.1], 10 steps of ∆t = 100 over [101.1,
1101.1] and 20 steps of ∆t = 1000 over [1101.1, 21101.1]. However, numerical oscillation took place at the early
stage in that study because Eq. (62) was violated with inputs ∆t = 0.01 s and ∆h = 0.2 m. In this study, we set
the initial timestep ∆t = 1.0 s and gradually enlarges it following ∆t = 1.0 × 1.1step , which has been verified to
be stable at the early stage of the HM coupling simulation. The inputs of this example are listed in Table 2. Note
16
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Table 2
Inputs of the isothermal and non-isothermal consolidation problems.
Elastic modulus (E) 6000 Pa
Poisson’s ratio (µ) 0.4
Permeability (km ) 4 × 10−6 m/s
a Thermal conductivity (λ ) 0.836 W/(m◦ C)
m
a Heat capacity ((ρc) ) 167.2 J/( m3◦ C)
m
a Volumetric thermal expansivity (β ) 9 × 10−7
v
θ 0.875
a Applicable to Section 3.4 only.

Fig. 9. Terzaghi 1D consolidation: (a) numerical results of excess pore pressure and settlement of the top surface as compared with the
analytical solutions, and (b) contours of excess pore pressure and vertical displacement at two different times.

that some parameters in Table 2 are only applicable to Section 3.4. Fig. 9(a) compares the results of HM SOLVER
with Eq. (64). The black solid lines indicate the analytical solutions, while the scattered dots are our numerical
results. The red circles display the excess pore pressure at z = 3.0 m, and the blue diamonds exhibit the settlement
at z = 0 m (the top surface). It is evident that the analytical and numerical results are in good agreement with each
other. Fig. 9(b) shows the contours of excess pore pressure and vertical displacement at two different times to give
readers an intuitive sense of the distribution of these two variables.

3.4. 1D nonisothermal consolidation

In Sections 3.1–3.3 we invoked analytical solutions to validate T SOLVER and HM SOLVER separately, and
in this section we will test the T SOLVER and HM SOLVER working as a whole. To the best knowledge of
the authors, no analytical solutions associated with THM coupling are available in the literature. Nevertheless,
a 1D nonisothermal consolidation problem has been leveraged by a number of researchers [6–8,37] to verify
their respective 2D THM coupling programs. Therefore, such a problem with the corresponding numerical results
can serve as a benchmark to validate the present 3D THM coupling model. The numerical setting of the 1D
nonisothermal consolidation problem is the same as Section 3.3, except that temperature is incorporated (see the
inset of Fig. 10(a)). The initial temperature of the soil column is 0 ◦ C, and along with the exertion of the top
surface traction, the top surface temperature is elevated to 50 ◦ C. The thermal properties of the soil column are
presented in Table 2. Fig. 10(a) compares the settlement of the top surface of the present THM coupling model
with those of previous studies. Under the HM coupling circumstance, the surface settlement builds up and levels
out as time elapses (Fig. 9(a)). In contrast, there is a rebound stage in the THM coupling case due to the thermal
expansion of the soil column (Fig. 10(a)). It is evident in Fig. 10(a) that all the numerical results show decent
agreement at both the early and late stages, but deviate a bit from each other near the maximum settlement. We
cannot plausibly explain the slight differences among these simulation results near the peak settlement, but our
result well matches that of Noorishad et al. [7]. Overall, the general agreement among these data sets would suffice
17
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Fig. 10. 1D nonisothermal consolidation: (a) surface settlement evaluated by the present 3D THM coupling model as compared with those of
previous 2D models, and (b) settlement history varying along with thermal conductivity where permeability is fixed to be 4.0 × 10−6 m/s.

Table 3
Three sets of initial guesses for the UPT solution scheme.
Expression of the initial guess Set no. γ
do i = 1, NNOD (Number of nodes) 1 0
X(i)=γ × i
end do 2 1 × 10−4
Note: X represents ∆u, ∆p and ∆T 3 4 × 10−4

to warrant the validity of the present THM coupling model. An upshot of this example is that the history of the
surface settlement was closely related to the ratio of thermal conductivity to permeability. Fig. 10(b) shows different
surface settlements under both isothermal and non-isothermal conditions. In the latter case, the permeability is fixed
but thermal conductivity varies. The thermal expansion leads to the uplift of the top surface while the dissipation
of pore pressure results in top surface settlement. Consequently, the settlement history is a net result of thermal
expansion and pore pressure dissipation. A larger ratio of thermal conductivity to permeability implies that thermal
expansion overwhelms the dissipation of pore pressure, causing a reduction in the peak settlement.

4. Comparison among different solution schemes


In this section, we will first compare the performance of the UPT, UP-T and U-PT solution schemes based on the
1D nonisothermal consolidation problem, and then elucidate what influences the applicability of different solution
schemes.

4.1. Performance assessment

The performance of the UPT, U-PT and UP-T solution schemes are evaluated from the perspective of accuracy,
efficiency and convergency. Under the UPT scheme, the Newton–Raphson iteration method relies on the initial guess
to update the coefficient matrix. We here test three sets of initial guesses (Table 3) to investigate their influence
on the performance of the UPT solution scheme. In contrast, the initial guess does not impact the UP-T solution
scheme because both the subsets are linear.
Fig. 11(a) compares the settlement of the top surface evaluated by the UP-T solution scheme and the UPT solution
scheme with the three sets of initial guesses. The three sets of initial guesses do not influence the simulation accuracy.
Across the UP-T and UPT solution schemes, the settlement curves exhibit only a slight difference. Fig. 11(b) shows
the history of the number of Picard/Newton–Raphson iterations under the UP-T/UPT solution scheme. As stated
above, Newton–Raphson iteration is sensitive to the initial guess. Although the initial guess does not influence
the simulation accuracy, it impacts the simulation efficiency. The number of Newton–Raphson iterations fluctuates
during the course of the simulation. In contrast, the UP-T solution scheme requires fewer Picard iterations to
18
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Fig. 11. (a) Settlement evaluated by the UP-T scheme, and the UPT scheme with three sets of initial guesses. (b) History of the number
of Picard/Newton–Raphson iterations under the UP-T/UPT scheme.

Fig. 12. (a) Under the U-PT solution scheme, Picard iteration is very sensitive to ω in Eq. (59) and diverges after two iterations. (b) In
contrast, Picard iteration under the UP-T solution scheme remains stable even changing ω a lot.

converge, and the number of Picard iterations seems to be more stable throughout the simulation. Fig. 11 indicates
that both the UPT and UP-T solution schemes are accurate, while the UP-T solution scheme tends to be more
efficient.
Unfortunately, the U-PT solution scheme fails even at the first step in our simulation. To investigate the reason, we
use the first-step output (u, p and T) evaluated by the UP-T solution scheme as the initial value and march towards
the second step with the U-PT solution scheme. Picard iteration usually relies on Eq. (59) to update the primary
variables. Fig. 12(a) shows that although the nested Newton–Raphson iteration converges, the Picard iteration is
extremely sensitive to ω (Eq. (59)). The inset of Fig. 12(a) shows that after the first Picard iteration, small differences
in ω lead to slight variations in excess pore pressure. However, such variations are drastically magnified after the
second Picard iteration, which results in the divergence of the Picard iteration. On the contrary, Fig. 12(b) shows
that the UP-T solution scheme is insensitive to ω and the calculation remains stable even changing ω a lot.

4.2. Degree of coupling vs applicability of solution schemes

Both the UP-T and U-PT solution schemes involve Picard iteration. What is the physical interpretation that the
U-PT solution scheme fails while the UP-T solution scheme well succeeds? We tend to believe it is due to the degree
of coupling between the subsets. As we stated in Section 1, the simulation of multiphysics boils down to solving a
system of PDEs, with one being coupled with the other(s). Suppose we have a set of system PDEs with A/B being
the candidate primary variable to establish PDE1/PDE2 (Table 4). For example, if A represents displacements and
19
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Table 4
Degree of coupling impacts the applicability of solution schemes.
Representative system PDEs Permutation Degree of coupling Applicability of solution scheme

small/no
1. δA ⇌ δB Weak (no) coupling Picard iteration (Yes); Newton–Raphson iteration (Yes)
small/no

small/no

P D E1(A, B, . . .) = 0 δA ⇌ δB

⎪ large
2.

P D E2(A, B, . . .) = 0 δA
large
⇌ δB
One-way coupling Picard iteration (Yes); Newton–Raphson iteration (Yes)
small/no


...

large
3. δA ⇌ δB Two-way coupling Picard iteration (May not); Newton–Raphson iteration (Yes)
large

B indicates pore pressure, then PDE1 and PDE2 are built by stress equilibrium and mass conservation equations,
respectively. The connotation of “coupling” indicates that if an increment δ A takes place on A, there must be an
associated increment δ B on B, and vice versa. According to the relative magnitude of δ A and δ B, three permutations
can be categorized, which are linked to the degree of coupling.
In the context of THM coupling, the coupling between temperature and displacements is typically one-way: a
small temperature increment can lead to significant thermal stress, while stress has a limited influence on temperature
evolution. In contrast, the principle of effective stress (Eq. (2)) and volumetric strain rate (Eq. (24)) makes the
coupling between pore pressure and displacements two-way. The failure of the U-PT solution scheme implies that
the two-way coupling is too tight, and even a small perturbation in one primary variable can lead to a significant
variation in the other. In this case, Picard iteration may not work and the two associated PDEs should better be
solved simultaneously. On the other hand, the success of the UP-T solution schemes indicates that if the coupling
is one-way or weaker, Picard iteration can reduce nonlinearity and improve simulation efficiency. To sum up, our
data indicate that the degree of coupling is a dominant factor that influences the applicability of solution schemes.

5. Engineering applications of the UP-T solution scheme


Section 4 demonstrates that the UP-T solution scheme takes an upper hand over the UPT and U-PT solution
schemes. In this section, we will use the UP-T solution scheme to simulate two large complex engineering cases
to further test its robustness.

5.1. Geological disposal of nuclear wastes

Geological disposal has long been considered as a sustainable way to manage high-level and long-lived nuclear
wastes. By designing a sophisticated system with engineered barriers, nuclear wastes can be protected and isolated
to prevent potential adverse influences on the biosphere. Although the inventory of nuclear wastes continuously
increases, as yet, the geological disposal of nuclear wastes has not been realized in any country, largely due to
insufficient confidence to ensure repository safety over an exceedingly long time. Nevertheless, research progress
on this topic in the last two decades has provided reliable technical solutions underpinned by sound scientific
investigations. One typical example is the remarkable international collaboration project DECOVALEX [9], which
was initiated in 1992 and is still ongoing by the time of writing. The core mission of DECOVALEX is to advance
the study of THM coupling (currently chemical reaction also involved, namely THMC coupling) in geomaterials
and shed light on the geological disposal of nuclear wastes. In this section, we will use HENGYI and adopt a
conceptual repository model proposed by DECOVALEX [34] to explore the general performance of the nuclear
waste repository.

5.1.1. Model conceptualization


Fig. 13 shows the conceptualization of the geometric model used in this section. Fig. 13(a) depicts the layout
of tunnels and deposition holes at the repository scale; the repository is composed of a large number of repetitive
tunnel units, one of which is portrayed in detail in Fig. 13(b). As shown in Fig. 13(b), many deposition holes are
20
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Fig. 13. Conceptualization of the geometric model used for the simulation of nuclear waste disposal. (a) Layout of the tunnels and deposition
holes at the repository scale, (b) arrangement of canisters, buffers and backfill at the tunnel scale and (c) final geometric model used for
the present THM coupling simulation. The dimensions and elevations are referred to [34].

drilled with equal spacing below the floor of the tunnel. The canister is surrounded by the buffer material (usually
bentonite) and the tunnel may be backfilled by rock spoil blended with clay after the emplacement of canisters.
Again, one tunnel unit demonstrated in Fig. 13(b) comprises a number of repetitive units, which leads to the final
geometric model utilized in this section (Fig. 13(c)). Fig. 13(c) illustrates the mesh of our model and the global
coordinate system is also defined there. A 1000-m-high rock column is considered, with the depth of the upper
boundary being 950 m and the tunnel situated in the middle of the rock column. The simulation is divided into two
stages as illustrated in Fig. 14. The first stage is to obtain the in-situ stress, pore pressure and temperature fields after
the excavation of tunnels and deposition holes, based on which the canister and buffer are emplaced to trigger the
second-stage THM coupling simulation. In the first stage (Fig. 14(a)), vertical stress σv = 26.09 MPa representing
the overburden is applied to the top surface of the model and the horizontal stress is designated as 1.5σv . The initial
temperature is 45 ◦ C throughout the model, and the initial pore pressure is equal to the hydrostatic pressure. The
lateral sides are considered to be adiabatic and impermeable. The boundaries of the tunnel and deposition hole are
free to deform, and the temperature and pore pressure therein are prescribed as 20 ◦ C and 0 MPa, respectively. With
the boundary and initial conditions stated above, an equilibrium state was reached in the first stage. Subsequently,
the lateral sides were fixed horizontally and the temperature boundary condition applied to the tunnel surface was
removed (Fig. 14(b)). The THM coupling simulation started immediately after the emplacement of the canister
and buffer. Some points/lines of interest are marked in Fig. 14(c) at/along which the simulation results will be
demonstrated in the subsequent sections. The heat power of the canister decays with time following the expression
shown in Fig. 15(a), and the stability of rock mass can be assessed by the Hoek–Brown failure criterion depicted
in Fig. 15(b). If we denote
σ1′ − σ3′ 2 σ′
F =( ) −m 3 −s (66)
σc σc
then F = 0 indicates the failure envelope shown in Fig. 15(b). In the case of F > 0, the stress state falls above
the Hoek–Brown failure envelope, suggestive of rock failure at the candidate location. Conversely, if F < 0 then
the rock mass is safe. The mechanical, hydraulic and thermal parameters of the host rock, buffer and canister are
summarized in Table 5.
It should be noted that the geometric model, simulations stages, boundary and initial conditions, heat power
of the canister and the inputs of Hoek–Brown criterion mentioned above were mostly referred to [34], but not
completely the same. Firstly, the research teams reported by [34] adopted only one quarter of the geometric model
shown in Fig. 13(c) given its symmetry. As we have relatively abundant computational resources, the whole model
shown in Fig. 13(c) was retained. Secondly, Millard et al. [34] emplaced not only buffer and canister in the second
simulation stage, but also engaged the backfill. We did not consider the backfill because there is a rising trend that
many countries start to regulate the retrievability of the canisters [67]. For example, current U.S. law stipulates “up
to 50 years after the waste emplacement operations are initiated” (in 10 CFR 60.111). As reported by previous
21
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Fig. 14. Boundary and initial conditions used for two simulation stages: (a) before and (b) after waste emplacement. (c) Some points and
lines of interest are marked to demonstrate the simulation results. (After [34]).

Fig. 15. (a) Expression by curve fitting to indicate the decay of heat power of a canister along with time, and (b) Hoek–Brown failure
criterion to assess the stability of rock mass. In (b), σ1′ and σ3′ denote the minimum and maximum effective stresses, σc is the uniaxial
compressive stress, and m and s are two empirical parameters. (After [34]).

studies, the peak temperature within the buffer tends to take place around 20–30 years after the waste emplacement
[1,34,68,69]. A large volume of backfill substantially influences the near-field performance, leading to potential
uncertainties if the waste retrievability policy is to be complied with. Lastly, the timestep scheme in the second
simulation stage was not provided by [34]. Even if it was, since we used a different mesh, the timestep scheme
might be not well suitable for our simulation. In the present simulation, we used 0.33 years for the first 3 steps and
enlarged the timestep afterwards by a factor of 1.1 up to 63 steps to reach a duration of ∼1000 years.

5.1.2. Before emplacement of waste


Fig. 16 shows the general performance of the repository before the emplacement of waste. To demonstrate the
inner distribution of the contours, one quarter of the lower part of the model is cut off. According to the Sx′ x
contour, the lateral sides of the tunnel manifest the smallest compressive stress, which is vulnerable to spalling.
Tips of the deposition hole in the x-axis direction also show relatively small compressive stress. The disturbance of
′ ′
Syy is confined to the periphery of the deposition hole, and tensile Szz occurs at the floors of both the tunnel and
22
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Table 5
Material parameters of the nuclear waste disposal model.
Material property Host rock Buffer Canister
Solid density (ρs ) (kg/m3 ) 2746 1600 7800
Physical
Fluid density (ρ f ) (kg/m3 ) 1000 1000 1000
Porosity (φ) (%) 0.379 33 10−5
Elastic modulus (E) (GPa) 61 0.018 200
Mechanical
Poisson’s ratio (µ) 0.303 0.350 0.300
Fluid compressibility (β p ) (Pa−1 ) 4.5 × 10−10 4.5 × 10−10 4.5 × 10−10
Hydraulic Intrinsic permeability (k) (m2 ) 6.6 × 10−17 2.1 × 10−21 10−24
Fluid dynamic viscosity (µ f ) (Pa s) 0.001 0.001 0.001
Solid specific heat (cs ) (J/(kg◦ C)) 900 800 460
Solid heat conduction coefficient (λs ) (W/(m◦ C)) 2.71 1.20 53
Thermal Solid volumetric thermal expansion coefficient (βv ) (◦ C−1 ) 2.46 × 10−5 3.00 × 10−5 4.92 × 10−6
Fluid specific heat (c f ) (J/(kg◦ C)) 4200 4200 4200
Fluid heat conduction coefficient (λ f ) (W/(m◦ C)) 0.6 0.6 0.6
Fluid volumetric thermal expansion coefficient (βT ) (◦ C−1 ) 3.00 × 10−4 3.00 × 10−4 3.00 × 10−4

deposition hole. To quantitatively assess the stability of rock mass after excavation, Eq. (66) was used to evaluate
F for each element. According to the F contour, the rock mass after excavation is largely stable, except for the
floors of the tunnel and deposition hole. In terms of pore pressure and temperature, the distributions thereof are
very much similar, with small values adjacent to the tunnel and deposition hole and a gradual enlargement towards
the upper and lower boundaries of the model.

5.1.3. After emplacement of waste


Fig. 17(a) compares the temperature at point B4 (refer to Fig. 14 for position) of the present simulation with
those of two research teams (JNC and SKI) reported by [34]. The peak temperature of our simulation is higher and
comes a bit earlier than those of the other two due to the absence of backfill. Yet all the three peak temperatures are
smaller than 100 ◦ C (boiling point of water), ensuring the safety of the buffer. Fig. 17(b) illustrates the variation of
temperature on the xOz cross-section (see Fig. 13(c) for definition) over 100 years. Because the canister is located
below the floor of the tunnel and the empty tunnel retards the upward heat transfer, the temperature increase is
mainly confined to the lower part of the model. Fig. 18(a) shows the variation of temperatures at points B1, B2, B4
and B5 over 1000 years. Among them, the temperature at B4 is the highest, therefore the interface between buffer
and canister is the critical location to ensure buffer safety. Fig. 18(a) displays the temperature distribution along ray
r (see Fig. 14(c) for definition) after 1, 10, 25 and 100 years; after 100 years, the temperature difference among
canister, buffer and host rock is fairly minor.
Fig. 19 depicts the contours of pore pressure and streamlines over 100 years. Since the tunnel surfaces are set as
zero-pore-pressure boundaries in this simulation and the permeability of the host rock is much larger than that of
the buffer, the pore pressure and streamlines in the host rock manifest limited changes over 100 years. In contrast,
the pore pressure variation in the buffer is considerable due to its low permeability, and for the same reason, the
streamlines in the host rock bypass the buffer to reach the floor of the tunnel (Fig. 19(b)). Fig. 20(a) shows the pore
pressures at B2, B5 and B6 over 1000 years. As stated above, the variation of pore pressure is very limited in the
host rock as evidenced by the almost constant pore pressure at B6. In contrast, the pore pressure in the buffer first
builds up, and then dissipates as the heat of the canister is gradually depleted. Interestingly, Fig. 20(b) implies that
the gradient of pore pressure in the buffer along the radial direction points outward at the early stage (see the lines
of 10 and 25 years), but flips after 150 years.
Fig. 21 shows the total stresses 25 years after the emplacement of waste. Because the lateral sides of the host rock
are fixed horizontally upon the emplacement of the canister and buffer, compressive stresses increase in the host rock
as a result of thermal expansion. In addition, Fig. 22(a) indicates that the total stresses (Sx x ) at B2 and B5 also rise
continuously at the early stage and reach their respective peaks after ∼10 years. Fig. 22(b) shows the evolution of
the effective stress (Sx′ x and Szz

) at B2 and B5. Interestingly, our simulation result suggests tensile effective stresses
within the buffer, which constitutes a potential threat to the buffer. We interpret the tensile effective stresses being
23
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Fig. 16. General performance of the reservoir before the emplacement of waste. The model is rotated 90◦ clockwise and one quarter of the
lower part is cut off to show the inner distribution of the contours.

caused by a larger thermal expansion coefficient (3 × 10−4 ) of the fluid than that of the buffer material (3 × 10−5 ).

Such speculation can be further underpinned by greater Szz than Sx′ x at both B2 and B5 (Fig. 22(b)) because the
vertical extent of the buffer exceeds its lateral extent, which facilitates the build-up of pore pressure. Note the
assumption of the fully saturated medium in this study may overestimate the pore pressure since the resaturation
process is omitted. Nevertheless, Rutqvist et al. [1] reported that it took around 3–10 years for the buffer to be fully
saturated, while Fig. 17(a) indicates the peak temperature arrives 20–30 years after the emplacement of the canister
and buffer. Therefore the resaturation well precedes the peak temperature and the potential occurrence of tensile
effective stresses persists. Although the magnitude of the tensile effective stresses suggested by Fig. 22(b) is small,

24
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Fig. 17. (a) Temperature at B4 of this study as compared with those of two research teams reported by [34], and (b) temperature contours
on the xOz cross-section after 1, 10, 25 and 100 years.

Fig. 18. (a) Temperatures at B1, B2, B4 and B5 over 1000 years, and (b) temperature distribution along ray r after different years.

Fig. 19. (a) Contours of pore water pressure and (b) streamlines over 100 years. Background of (b) is the pore pressure contour and refer
to (a) for legend.

Fig. 20. (a) Pore pressures at B2, B5 and B6 over 1000 years and (b) distribution of pore pressure along r ′ after different years.

25
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Fig. 21. Total stresses: 25 years after the emplacement of waste.

Fig. 22. (a) Total stresses (Sx x ) and (b) effective stresses (Sx′ x and Szz
′ ) at B2 and B4 over 1000 years.

the buffer is fairly weak to resist tensile stress. To the best knowledge of the authors, such an issue has not been
reported in the literature, thus we advise a thorough future investigation into this.

5.2. Yangbajing geothermal exploitation

The example in Section 5.1 well demonstrates the capability of the UPT solution scheme in simulating 3D
THM coupling, yet the heat transfer therein was conduction-dominated. In this section, we will provide a more
practical application example based on Yangbajing geothermal field in China to highlight the advection-dominated
heat transfer.

5.2.1. Background
Due to the collision and subduction of the continental plate of China with the surrounding Indian Ocean plate,
Philippine sea plate and Pacific plate, China is fairly rich in geothermal energy, especially in south Tibet, west
Yunnan and north Hainan [2,70]. The Yangbajing geothermal field is located 94 km northwest of Lhasa, the
capital city of Tibet, and is the first high-temperature advective hydrothermal field in China [2,70]. The Yangbajing
geothermal field is believed to be a typical non-volcanic high-temperature geothermal system, which is heated
26
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Fig. 23. (a) Faults and well locations of the Yangbajing geothermal field (After [70]). 1-road, 2-observed fault, 3-inferred fault, 4-exploration
well, 5-production well and 6-deep well. (b) Pore pressure (p) and temperature (T) logs of well ZK4001 with the lithology profile (After [73]).

by a magma chamber 5–15 km underneath [70–73]. The aquifer of the Yangbajing geothermal field is mainly
recharged by atmospheric precipitation and melting snow [70–72]. Fig. 23(a) shows the faults and well locations
at the Yanbajing geothermal field. As indicated by the lithology profile of well ZK4001 (Fig. 23(b)), there are
mainly two reservoirs (one shallow and the other deep) down to 1500 m depth. The shallow reservoir has been
exploited for more than 30 years to generate power with an installed capacity of 27.30 MW. However, the bottom
hole temperature, pressure and flow rate at production wells dropped significantly, supporting only 16 MW as of
2016 [70]. Since the shallow reservoir is gradually depleted, the deep reservoir is targeted to sustain and expand the
power plant. The deep reservoir is located 950–1350 m below the surface and consists of highly fractured granite
(Fig. 23(b)). According to the drill hole logs of well ZK4001, the natural intrinsic permeability of the deep reservoir
is approximately 1–25 mD (1mD = 1 × 10−15 m2 ), which requires only low-level reservoir stimulation [70,72,73].
Considering the relatively shallow depth and remarkable permeability of the deep reservoir, it holds significant
prospects to be tapped. Fig. 23(b) also illustrates the temperature and pore pressure logs of well ZK4001. The
pore pressure at the upper bond of the deep reservoir is ∼8.1 MPa and increases linearly downward. In contrast,
the temperature of the deep reservoir is fairly uniform at 248 ◦ C. The deep reservoir is situated between a cap
rock layer (biotite granite) and a base rock layer (granite mylonite); both the cap and base rock layers are far less
permeable than the deep reservoir.

5.2.2. Model concepturalization


Different strategies were proposed to tap the Yangbajing deep reservoir, such as the horizontal-well design
suggested by [2] and the classic five-spot well configuration employed by [72]. These previous studies aimed at
HT coupling using the TOUGH2 simulator, and to the best knowledge of the authors, no THM coupling simulation
associated with the Yangbajing deep reservoir have been reported. In this section, we will leverage the UP-T solution
scheme and the five-spot well configuration (Fig. 24) to fill such a gap. Due to symmetry, only one quarter of the
five-spot configuration is considered (Fig. 24(a)). The highly fractured reservoir is assumed to be a continuous
saturated poroelastic medium, and its average permeability increases tenfold after stimulation (∼250 mD). The
injection well is situated at the bottom of the reservoir while the production well is located at the top. The purpose
of such a layout is to facilitate a longer trajectory for the streamlines (Fig. 24(b)), which can enhance the lifetime of
the reservoir [2,72]. The last 40 m of both the injection and production wells are perforated to better inject/withdraw
water [72].
As with the model in Section 5.1, two simulation stages are performed herein with the first stage providing
the in-situ stress, pore pressure and temperature fields before injection, and the second stage simulating a 30-year
prolonged injection. The mesh and dimensions of the present THM coupling model are indicated in Fig. 25(a), and
27
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Fig. 24. (a) Conceptual five-spot configuration to tap the deep reservoir. (b) Demonstration of well design on the injector–producer profile
(After [72]).

Fig. 25. (a) Mesh and dimensions of the THM coupling model of the deep reservoir of Yangbajing geothermal field. Boundary and initial
conditions for (b) the first simulation stage to provide the in-situ stress, pore pressure and temperature fields before injection and (c) the
second stage to simulate a 30-year prolonged injection.

the initial and boundary conditions of the two simulation stages are displayed in Figs. 25(b) and 25(c), respectively.
The model measures 363.6 m in both x- and y-directions (Fig. 25(a)), thus the distance between the injector and
producer is 500 m. In the z-direction, apart from the 400-m reservoir, a cap rock layer (100 m) and a base rock
layer (100 m) are incorporated to reduce the influence of upper and lower boundaries on heat transfer [2]. The
mesh required for advection-dominated problems is generally finer than that of conduction-dominated problems.
Therefore in this model, we not only use a relatively fine mesh but further refine the mesh near the injector and
producer (Fig. 25(a)).
In the first-stage simulation, the vertical stress σv is set as 25.0 MPa to consider the overburden, and the lateral
stresses are prescribed as 1.5σv . In the second-stage simulation, the fluid is injected at 60 ◦ C and at a rate of
18.5 L/s which is equally distributed along the perforated portion of the injector, while the pore pressure along the
perforated portion of the producer is fixed as 4.0 MPa. The material parameters of the present model are listed in
Table 6. The timestep scheme of the second-stage simulation is designed as follows: 11.5 days for the first step and
then enhancing the timestep by a factor of 1.15 up to 25 steps (∼6.7 years for the first 25 steps); the 26th step is
designated as 0.3 years, after which the timestep is fixed to be 1.0 year and the simulation continues for another
24 steps. In total, a 30-year injection is simulated.

28
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Table 6
Material parameters of the Yangbajing geothermal model.
Material property Cap & base rock rock Reservoir
Solid density (ρs ) (kg/m3 ) 2746 2650
Physical
Fluid density (ρ f ) (kg/m3 ) 1000 1000
Porosity (φ) (%) 0.379 10
Elastic modulus (E) (GPa) 61 30
Mechanical Poisson’s ratio (µ) 0.303 0.3
Fluid compressibility (β p ) (Pa−1 ) 4.5 × 10−10 4.5 × 10−10
Intrinsic permeability (k) (m2 ) 6.6 × 10−17 250 × 10−15
Hydraulic
Fluid dynamic viscosity (µ f ) (Pa · s) 0.001 0.001
Solid specific heat (cs ) (J/(kg◦ C)) 900 1000
Solid heat conduction coefficient (λs ) (W/(m◦ C)) 2.71 2.50
Thermal Solid volumetric thermal expansion coefficient (βv ) (◦ C−1 ) 2.70 × 10−5 2.70 × 10−5
Fluid specific heat (c f ) (J/(kg◦ C)) 4200 4200
Fluid heat conduction coefficient (λ f ) (W/(m◦ C)) 0.6 0.6
Fluid volumetric thermal expansion coefficient (βT ) (◦ C−1 ) 3.00 × 10−4 3.00 × 10−4

5.2.3. Effectiveness of the 3D consistent SUPG


To further verify the effectiveness of the 3D consistent SUPG method, we compare the temperature evolution
with/without SUPG stabilization. Peclet number is associated with both mesh size and flow velocity. In this model,
the flow velocity varies over the reservoir and we use unstructured mesh, thus a unique Peclet number is not
available to quantify the degree of heat advection. Nevertheless, we can use an alternative approach to evaluate the
effectiveness of SUPG stabilization. In Eq. (47), the term τ ∇w T vrT indicates the SUPG contribution to the weighting
function. Instead of applying this term all at once, we can use κτ ∇w T vrT and gradually increase κ from 0 to 1 to
observe the SUPG influence. The temperature along a monitoring line CD (see Fig. 24(b)) is used to demonstrate
the effectiveness of SUPG stabilization. Five sets of κ(0.00, 0.05, 0.15, 0.50 and 1.00) are tested and temperatures
after 11.5 and 233 days of injection are shown in Fig. 26(a) and (b) respectively. Calculation fails when κ=0.00,
thus we only present the results of the remaining four sets. Fig. 26(a) shows violent numerical oscillation near the
injector when κ=0.05 and 0.15, which implies that κ is too small to stabilize the simulation. In Fig. 26(b), numerical
oscillation builds up when κ=0.05 and it is so large that off the chart. In contrast, the simulation seems to gradually
turn to stable when κ=0.15 because only a minor oscillation is present in Fig. 26(b). In both Figs. 26(a) and 26(b),
κ>0.50 can well lay to rest the oscillation. Fig. 26 proves that the 3D consistent SUPG method developed by
this study is effective to stabilize the numerical oscillation incurred by advection-dominated heat transfer, which
otherwise is impossible to simulate the Yangbajing hydrothermal field under the conventional FEM scope.

5.2.4. Reservoir performance over 30-year injection


In this section, we will present the reservoir performance over a 30-year injection period with SUPG stabilization
κ=1.00. Fig. 27 displays the evolutions of temperature, pore pressure and vertical displacement (might be associated
with surface settlement) on the injector–producer profile over the 30-year prolonged injection. As the injection
proceeds, the host rock near the injector (bottom left) gradually cools down and the thermal front migrates towards
the producer (top right). After ∼22.5 years the thermal front reaches the producer, and the production temperature
begins to drop at an increasing gradient (Fig. 28(a)). Since the permeability of the base rock is much smaller than
that of the reservoir, the flow is mostly confined within the reservoir. Therefore, heat advection is less significant in
the base rock, and the low temperature percolates downward very slowly. The thermal front is smooth and flat in the
first 15 years, implying that the low-temperature part expands evenly at the early injection stage. As the streamlines
converge at the producer, the low-temperature fluid is carried to the producer too. Consequently, the temperature
contours are sharpened near the producer towards the late injection stage.
The pore pressure reaches a steady state very fast, with ∼19.5 MPa at the injector, and shows almost no
change in the subsequent injection. In contrast, the vertical displacement manifests substantial variation during the
30-year injection. After 1 year, the base rock deforms upward slightly, while the host rock above the injector moves
downward to a larger extent. This is because the host rock contracts upon cooling and pulls the ambient rock mass
29
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Fig. 26. Temperature distribution along CD (see Fig. 24(b)) with different degrees of SUPG stabilization after (a) 11.5 and (b) 233 days of
injection.

towards it. As time elapses, the settlement region within the reservoir expands and migrates towards the caprock.
Fig. 25(b) shows the evolutions of vertical displacement at A and B over 30 years. Readers can refer to the inset
of Figs. 28(b) and 24(b) for the locations of A and B. Interestingly, the vertical displacements at A and B both
vary with time in a perfectly linear manner. The inset of Fig. 24(b) indicates the distribution of settlement on
the top surface of our model after 30 years; the settlement centers around the injector and radiates towards the
producer. Given previous HT coupling simulations [2,70,72] and the present THM coupling simulation associated
with the Yangbajing deep reservoir, the temperature evolutions revealed by all these studies are comparable, but
we used a higher intrinsic permeability for the reservoir to obtain the same injection pressure. A possible reason
for the discrepancy may lie in the coupling between displacements and pore pressure, which was not considered
in the previous studies mentioned above. In other words, our study suggests that HT coupling may underestimate
the injection pressure. Since injection pressure plays a key role in engineering operation and it also significantly
influences the stability of existing faults, we advise more future efforts to dig deeper into this.

6. Conclusions
THM coupling is a critical and vibrant research topic associated with various sorts of underground activities. In
this study, we develop a new THM coupling simulator HENGYI and bridge a series of research gaps as follows.
(1) UP-T is another competent solution scheme to perform THM coupling simulations. Our study supports
that it is superior to the UPT and U-PT counterparts. The UP-T solution scheme naturally avoids nonlinearity
by decomposing the system PDEs into two linear subsets, thus Newton–Raphson iteration is not involved. In
addition, the low sensitivity of temperature to displacements and pore pressure (one-way coupling) facilitates a fast
convergence of Picard iteration. Two large engineering application examples prove that the UP-T solution scheme
is robust enough to simulate large complex THM coupling problems.
(2) We generalize the consistent Streamline Upwind Petrov Galerkin (SUPG) method to the 3D case based on
the 8-node linear element to stabilize the numerical oscillation incurred by advection-dominated heat transfer. The
proposed SUPG is rigorous in theory as we derived it from modifying the weight function in the weak form of the
associated PDE. The new SUPG method embedded in HENGYI allows us to simulate the heat extraction of the
Yangbajing hydrothermal field, which otherwise is impossible under the conventional FEM scope.
(3) The degree of coupling significantly influences the applicability of solution schemes. In the context of
multiphysics, one prime variable can be in weak(no)/one-way/two-way coupling with another. The monolithic
scheme with Newton–Raphson iteration is a generic strategy to solve system PDEs. However, reasonably leveraging
Picard iteration can reduce nonlinearity and improve code efficiency. The key idea of applying Picard iteration is
trying to avoid the two-way coupling, because the two-way coupling is so tight that may lead to the divergence of
Picard iteration.
30
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Fig. 27. Evolutions of temperature, pore pressure and vertical displacement on the injector-producer profile over 30-year prolonged injection.

31
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

Fig. 28. (a) Temperature at the producer and (b) vertical displacements at A and B (see the present inset and Fig. 22(b) for locations) over
30 years. The inset in (b) shows the vertical displacement contour of the top surface of the model after 30 years.

(4) HENGYI does not rely on any existing codes, and this enables us to present great details regarding the
development of HENGYI. Our motivation is trying to expose the key rationale and provide a comprehensive
understanding of THM coupling simulation techniques. This may benefit a wide range of THM coupling audiences,
from beginners to veterans.
Nomenclature

Nomenclature
σ Total stress α Biot coefficient

σ Effective stress K s Bulk modulus of the rock matrix
f Body force K g Bulk modulus of mineral grains
u Displacement βv Volumetric thermal expansion coefficient
ε Strain φ Porosity
εT Thermal expansivity ρ f Fluid density
D Elastic matrix ρs Solid density
σ b Stress boundary condition q f s Flow source
v f Flow velocity β p Compressibility coefficient of the fluid
v s Solid velocity βT Volumetric thermal expansion coefficient
vr Relative velocity of the fluid
x 1 Location of the fluid source µ f Dynamic viscosity of the fluid

x 1 Location of the heat source c f Specific heat of the fluid
k Intrinsic permeability cs Specific heat of the solid
g Gravitational acceleration qhs Heat source
λ f Thermal conductivity of the fluid ∆t Timestep
λs Thermal conductivity of the solid θ Integration parameter
λm Average thermal conductivity E Young’s modulus
w Weight function µ Poisson’s ratio
n Outward normal vector km Permeability
I Auxiliary matrix δ Dirac function
N Shape function Pe Peclet number
N Shape function ∆h Mesh size
p Pore pressure Ω Volumetric region of interest
T Temperature ∂Ω Boundary of Ω
∆ p Pore pressure increment ω Weight factor of Picard iteration
∆T Temperature increment
32
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

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.

Acknowledgments
We thank Dr. Tianjiao Li at Dalian University of Technology, China for discussing THM coupling details with
us. We acknowledge the support from National Natural Science Foundation of China under Grant No. 41877217,
General Research Fund of the Research Grants Council (Hong Kong) under Grant No. 17200721, and Natural
Science Foundation of Guangdong Province, China under Grant No. 2019A1515010999. The Open Access Charge
of this paper is supported by the first grant mentioned above (No. 41877217). The second author acknowledges the
Postgraduate Scholarship from the University of Hong Kong.

Appendix A. UPT Solution scheme in HENGYI


Suppose the stress equilibrium, mass and energy conservation equations in Eq. (58) can be denoted by
F̂ u (∆u, ∆ p, ∆T ) = 0, F̂ p (∆u, ∆ p, ∆T ) = 0 and F̂ T (∆u, ∆ p, ∆T ) = 0, respectively. With the Newton–Raphson
method, Eq. (58) can be linearized as
⎡ ∂ F̂ ∂ F̂ u ∂ F̂ u ⎤
u
⎢ ∂(∆u) ∂(∆ p) ∂(∆T ) ⎥
⎡ ⎤ ⎡ ⎤
F̂ u (∆u, ∆ p, ∆T ) ∆u N R
⎥ ⎢ ∂ F̂ p ∂ F̂ p ∂ F̂ p ⎥
⎢ ⎥
⎢ ⎢ ⎥
⎢ F̂ p (∆u, ∆ p, ∆T ) ⎥ + ⎢ ⎢∆ p N R ⎥ = 0 (A.1)
⎦ ⎢ ∂(∆u) ∂(∆ p) ∂(∆T ) ⎥

⎣ ⎣ ⎦
⎣ ∂ F̂ T ∂ F̂ T ∂ F̂ T ⎦
⎢ ⎥
F̂ T (∆u, ∆ p, ∆T ) ∆TN R
∂(∆u) ∂(∆ p) ∂(∆T ) J aco
The Jacobian matrix in (A.1) can be evaluated by
⎡ ∂ F̂ ∂ F̂ u ∂ F̂ u ⎤
u
⎢ ∂(∆u) ∂(∆ p) ∂(∆T ) ⎥
⎡ ⎤
K̂ uu K̂ up K̂ uT
⎢ ∂ F̂ p ∂ F̂ p ∂ F̂ p ⎥
⎢ ⎥
⎣ K̂ pu K̂ pp1 + θ ∆t K̂ pp2
⎢ ⎥
⎢ ∂(∆u) ∂(∆ p) ∂(∆T ) ⎥
⎢ ⎥ =⎢ K̂ pT ⎥
⎦ (A.2)
⎣ ∂ F̂ T ∂ F̂ T ∂ F̂ T K̂ T u K̂ T p1 + θ ∆t K̂ T p2 θ∆t K̂ T T 1 + K̂ T T 2
⎢ ⎥

∂(∆u) ∂(∆ p) ∂(∆T ) J aco
with the entries
∫ being ∫ ∫
K̂ uu = ∇ N D∇ NdΩ , K̂ up =
T
α N I NdΩ ,
T
K̂ uT = βv N T I NdΩ ,
∫Ω ∫Ω ∫Ω
αρ f T T ρf T
K̂ pu = N CdΩ , K̂ pp1 = φβ p N NdΩ , K̂ pp2 = ∇ N k∇ NdΩ
Ω ρ 0 Ω Ω µ ρ
f 0
∫ ∫
T T T T
K̂ pT = − φβT N NdΩ , K̂ T u = − βv K s (N T̂ )(N + τ ∇ N vr )CdΩ
Ω Ω

T T (A.3)
K̂ T p1 = − θ φβ p (N T̂ )(N + τ ∇ N vrT )NdΩ ,

ρ 2f c f
∫ ∫
T T T
K̂ T p2 = − (N + τ ∇ N vrT )(∇ T
N T̂ ) k∇ NdΩ − ρ f c f τ ∇ N k∇ Nvr (∇ N T̂ )dΩ
Ω µ f ρ0 Ω
∫ ∫
T T T
K̂ T T1 = − ∇ N λm ∇ NdΩ , K̂ T T2 = (ρc)m (N + τ ∇ N vrT )NdΩ
Ω Ω
Collecting terms, Eq. (A.1) can be recast as
⎡ ⎤⎡ ⎤ ⎡ ⎤
K̂ uu K̂ up K̂ uT ∆u N R − F̂ u
⎢ K̂ pu K̂ pp1 + θ ∆t K̂ pp2 K̂ pT
⎢ ⎥⎢ ⎥ ⎢ ⎥
⎥ ⎢∆ p N R ⎥ = ⎢− F̂ p ⎥ (A.4)
⎣ ⎦⎣ ⎦ ⎣ ⎦
K̂ T u K̂ T p1 + θ ∆t K̂ T p2 θ∆t K̂ T T 1 + K̂ T T 2 ∆TN R − F̂ T
33
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

where the components of the right-hand side vector are


F̂ u = K̂ uu ∆u + K̂ up ∆ p + K̂ uT ∆T − F ub − F u f
F̂ p = K̂ pu ∆u + K̂ pp1 ∆ p + θ ∆t K̂ pp2 T + K̂ pT ∆T − F pb − F pg − F pq (A.5)
F̂ T = K̂ T u ∆u + K̂ T p1 ∆ p + θ ∆t K̂ T p2 T + θ∆t K̂ T T 1 T + K̂ T T 2 ∆T − F T b − F T q
Refer to Eqs. (15), (38) and (51) for expressions of F ub ∼ F T q in Eq. (A.5). The pseudo-codes of the UPT solution
scheme in HENGYI is illustrated in Algorithm A.

Appendix B. U-PT Solution scheme in HENGYI

Under the U-PT solution scheme, Eq. (58) can be recast as


F̂ u + K̂ uu ∆u = 0
⎡ ∂ F̂ p ∂ F̂ p ⎤
[ ] [ ]
F̂ p ⎢ ∂(∆ p) ∂(∆T ) ⎥ ∆ pN R (B.1)
+ ⎣ ∂ F̂ ∂ F̂ T ⎦ =0
F̂ T T ∆TN R
∂(∆ p) ∂(∆T ) J aco

The Jacobian matrix in (B.1) can be evaluated by


⎡ ∂ F̂ p ∂ F̂ p ⎤
K̂ pp1 + θ ∆t K̂ pp2
[ ]
⎢ ∂(∆ p) ∂(∆T ) ⎥ K̂ pT
⎣ ∂ F̂ T ∂ F̂ T ⎦ = (B.2)
K̂ T p1 + θ ∆t K̂ T p2 θ∆t K̂ T T 1 + K̂ T T 2
∂(∆ p) ∂(∆T ) J aco
where the entries can be found in Eq. (A.3). Collecting terms, Eq. (B.1) can be reformulated as
K̂ uu ∆û = − F̂ u
K̂ pp1 + θ ∆t K̂ pp2
[ ][ ] [ ]
K̂ pT ∆ pN R − F̂ p (B.3)
=
K̂ T p1 + θ ∆t K̂ T p2 θ ∆t K̂ T T 1 + K̂ T T 2 ∆TN R − F̂ T
34
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

with the right-hand side vector being


F̂ u = K̂ up ∆ p + K̂ uT ∆T − F ub − F u f
F̂ p = K̂ pu ∆u + K̂ pp1 ∆ p + θ ∆t K̂ pp2 p + K̂ pT ∆T − F pb − F pg − F pq (B.4)
F̂ T = K̂ T u ∆u + K̂ T p1 ∆ p + θ ∆t K̂ T p2 p + θ∆t K̂ T T 1 T + K̂ T T 2 ∆T − F T b − F T q
Refer to Eq. (A.3) for the expressions of K up ∼ K T T 2 in Eq. (B.4). The pseudo-codes of the U-PT solution scheme
in HENGYI is illustrated in Algorithm B.

References
[1] J. Rutqvist, M. Chijimatsu, L. Jing, A. Millard, T.S. Nguyen, A. Rejeb, et al., A numerical study of THM effects on the near-field
safety of a hypothetical nuclear waste repository—BMT1 of the DECOVALEX III project. Part 3: Effects of THM coupling in sparsely
fractured rocks, Int. J. Rock Mech. Min. Sci. 42 (5–6) (2005) 745–755.
[2] Y. Zeng, J. Zhan, N. Wu, Y. Luo, W. Cai, Numerical investigation of electricity generation potential from fractured granite reservoir
by water circulating through three horizontal wells at Yangbajing geothermal field, Appl. Therm. Eng. 104 (2016) 1–15.
[3] D. Baghernezhad, M. Siavashi, A. Nakhaee, Optimal scenario design of steam-assisted gravity drainage to enhance oil recovery with
temperature and rate control, Energy 166 (2019) 610–623.
[4] T.Y. Guo, L.N.Y. Wong, Z. Wu, Microcracking behavior transition in thermally treated granite under mode I loading, Eng. Geol. 282
(2021) 105992.
[5] D. Baghernezhad, M. Siavashi, A. Nakhaee, Optimal scenario design of steam-assisted gravity drainage to enhance oil recovery with
temperature and rate control, Energy 166 (2019) 610–623.
[6] B.L. Aboustit, S.H. Advani, J.K. Lee, Variational principles and finite element simulations for thermo-elastic consolidation, Int. J.
Numer. Anal. Methods Geomech. 9 (1) (1985) 49–69.
35
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

[7] J. Noorishad, C.F. Tsang, P.A. Witherspoon, Coupled thermal-hydraulic-mechanical phenomena in saturated fractured porous rocks:
Numerical approach, J. Geophys. Res. Solid Earth 89 (B12) (1984) 10365–10373.
[8] R.W. Lewis, C.E. Majorana, B.A. Schrefler, A coupled finite element model for the consolidation of nonisothermal elastoplastic porous
media, Transp. Porous Media 1 (2) (1986) 155–178.
[9] J.T. Birkholzer, C.F. Tsang, A.E. Bond, J.A. Hudson, L. Jing, O. Stephansson, 25 Years of DECOVALEX-scientific advances and
lessons learned from an international research collaboration in coupled subsurface processes, Int. J. Rock Mech. Min. Sci. 122 (2019)
103995.
[10] T.S. Nguyen, A.P.S. Selvadurai, July). Coupled thermal-mechanical-hydrological behaviour of sparsely fractured rock: implications for
nuclear fuel waste disposal, Int. J. Rock Mech. Min. Sci. Geomech. Abstracts 32 (5) (1995) 465–479, Pergamon.
[11] M.P. Ahola, A. Thoraval, Distinct element models for the coupled T − H − M processes: Theory and implementation, in: O. Stephansson,
L. Jing, C.-F. Tsang (Eds.), Developments in Geotechnical Engineering, vol. 79, Elsevier, 1996, pp. 181–211.
[12] Y. Sugita, M. Chijimatsu, A. Ito, H. Kurikaml, A. Kobayashi, Y. Ohnishi, THM simulation of the full-scale in-situ engineered barrier
system experiment in grimsel test site in switzerland, in: O. Stephansson (Ed.), in: Elsevier Geo-Engineering Book Series, vol. 2,
Elsevier, (ISSN: 1571-9960) ISBN: 9780080445250, 2004, pp. 119–124.
[13] M.B. Hsu, R.E. Nickell, Coupled convective and conductive heat transfer by finite element method, in: J.T. Oden, O.C. Zienkiewicz,
R.H. Gallagher, C. Taylor (Eds.), Finite Element Methods in Flow Problems, UAH Press, Huntsville, Ala, 1974, pp. 427–450.
[14] B.A. Schrefler, A partitioned solution procedure for geothermal reservoir analysis, Commun. Appl. Numer. Methods 1 (2) (1985) 53–56.
[15] J. Rutqvist, Y.S. Wu, C.F. Tsang, G. Bodvarsson, A modeling approach for analysis of coupled multiphase fluid flow, heat transfer,
and deformation in fractured porous rock, Int. J. Rock Mech. Min. Sci. 39 (4) (2002) 429–442.
[16] J. Lee, K.I. Kim, K.B. Min, J. Rutqvist, TOUGH-UDEC: a simulator for coupled multiphase fluid flows, heat transfers and discontinuous
deformations in fractured porous media, Comput. Geosci. 126 (2019) 120–130.
[17] T. Li, J. Rutqvist, M. Hu, TOUGH-RFPA: Coupled thermal-hydraulic-mechanical rock failure process analysis with application to deep
geothermal wells, Int. J. Rock Mech. Min. Sci. 142 (2021) 104726.
[18] K. Pruess, C.M. Oldenburg, G.J. Moridis, TOUGH2 User’s Guide Version 2, No. LBNL-43134, Lawrence Berkeley National
Lab.(LBNL), Berkeley, CA (United States), 1999.
[19] G.J. Moridis, K. Pruess, T2SOLV: An enhanced package of solvers for the TOUGH2 family of reservoir simulation codes, Geothermics
27 (4) (1998) 415–444.
[20] J. Rutqvist, J. Birkholzer, F. Cappa, C.F. Tsang, Estimating maximum sustainable injection pressure during geological sequestration of
CO2 using coupled fluid flow and geomechanical fault-slip analysis, Energy Convers. Manage. 48 (6) (2007) 1798–1807.
[21] J. Rutqvist, D.W. Vasco, L. Myer, Coupled reservoir-geomechanical analysis of CO2 injection and ground deformations at In Salah,
Algeria, Int. J. Greenh. Gas Control 4 (2) (2010) 225–230.
[22] A.P. Rinaldi, V. Vilarrasa, J. Rutqvist, F. Cappa, Fault reactivation during CO2 sequestration: Effects of well orientation on seismicity
and leakage, Greenh. Gases: Sci. Technol. 5 (5) (2015) 645–656.
[23] A.P. Rinaldi, J. Rutqvist, S. Finsterle, H.H. Liu, Inverse modeling of ground surface uplift and pressure with iTOUGH-PEST and
TOUGH-FLAC: The case of CO2 injection at In Salah, Algeria, Comput. Geosci. 108 (2017) 98–109.
[24] J. Rutqvist, C.M. Oldenburg, Analysis of injection-induced micro-earthquakes in a geothermal steam reservoir, the geysers geothermal
field, california, in: The 42nd US Rock Mechanics Symposium, USRMS, OnePetro, 2008.
[25] P. Jeanne, J. Rutqvist, P.F. Dobson, M. Walters, C. Hartline, J. Garcia, The impacts of mechanical stress transfers caused by
hydromechanical and thermal processes on fault stability during hydraulic stimulation in a deep geothermal reservoir, Int. J. Rock
Mech. Min. Sci. 72 (2014) 149–163.
[26] A.P. Rinaldi, J. Rutqvist, E.L. Sonnenthal, T.T. Cladouhos, Coupled THM modeling of hydroshearing stimulation in tight fractured
volcanic rock, Transp. Porous Media 108 (1) (2015) 131–150.
[27] J. Rutqvist, H.M. Kim, D.W. Ryu, J.H. Synn, W.K. Song, Modeling of coupled thermodynamic and geomechanical performance of
underground compressed air energy storage in lined rock caverns, Int. J. Rock Mech. Min. Sci. 52 (2012) 71–81.
[28] J. Kim, G.J. Moridis, Development of the T+ M coupled flow–geomechanical simulator to describe fracture propagation and coupled
flow–thermal–geomechanical processes in tight/shale gas systems, Comput. Geosci. 60 (2013) 184–198.
[29] R. Walsh, O. Nasir, N. Calder, S. Sterling, J. Avis, Combining TOUGH2 and FLAC3D to solve problems in underground gas storage,
Transp. Porous Media 123 (3) (2018) 501–519.
[30] J. Rutqvist, F. Cappa, A.P. Rinaldi, M. Godano, Modeling of induced seismicity and ground vibrations associated with geologic CO2
storage, and assessing their effects on surface structures and human perception, Int. J. Greenh. Gas Control 24 (2014) 64–77.
[31] L. Zheng, J. Rutqvist, J.T. Birkholzer, H.H. Liu, On the impact of temperatures up to 200 C in clay repositories with bentonite engineer
barrier systems: A study with coupled thermal, hydrological, chemical, and mechanical modeling, Eng. Geol. 197 (2015) 278–295.
[32] J. Rutqvist, Thermal management associated with geologic disposal of large spent nuclear fuel canisters in tunnels with thermally
engineered backfill, Tunn. Undergr. Space Technol. 102 (2020) 103454.
[33] C. Lee, J. Lee, G.Y. Kim, Numerical analysis of coupled hydro-mechanical and thermo-hydro-mechanical behaviour in buffer materials
at a geological repository for nuclear waste: Simulation of EB experiment at mont terri URL and FEBEX at grimsel test site using
Barcelona basic model, Int. J. Rock Mech. Min. Sci. 139 (2021) 104663.
[34] A. Millard, A. Rejeb, M. Chijimatsu, L. Jing, J. De Jonge, M. Kohlmeier, et al., Numerical study of the THM effects on the near-field
safety of a hypothetical nuclear waste repository—BMT1 of the DECOVALEX III project. Part 2: effects of THM coupling in continuous
and homogeneous rocks, Int. J. Rock Mech. Min. Sci. 42 (5–6) (2005) 731–744.
[35] Y. Chen, C. Zhou, L. Jing, Modeling coupled THM processes of geological porous media with multiphase flow: theory and validation
against laboratory and field scale experiments, Comput. Geotech. 36 (8) (2009) 1308–1329.
36
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

[36] F. Tong, L. Jing, R.W. Zimmerman, A fully coupled thermo-hydro-mechanical model for simulating multiphase flow, deformation and
heat transfer in buffer material and rock masses, Int. J. Rock Mech. Min. Sci. 47 (2) (2010) 205–217.
[37] W. Cui, D.M. Potts, L. Zdravković, K.A. Gawecka, D.M. Taborda, An alternative coupled thermo-hydro-mechanical finite element
formulation for fully saturated soils, Comput. Geotech. 94 (2018) 22–30.
[38] B. Gee, R. Gracie, Comparison of fully-coupled and sequential solution methodologies for enhanced geothermal systems, Comput.
Methods Appl. Mech. Engrg. 373 (2021) 113554.
[39] A. Ghassemi, X. Zhou, A three-dimensional thermo-poroelastic model for fracture response to injection/extraction in enhanced
geothermal systems, Geothermics 40 (1) (2011) 39–49.
[40] B. Guo, P. Fu, Y. Hao, C.A. Peters, C.R. Carrigan, Thermal drawdown-induced flow channeling in a single fracture in EGS, Geothermics
61 (2016) 46–62.
[41] S.N. Pandey, V. Vishal, Sensitivity analysis of coupled processes and parameters on the performance of enhanced geothermal systems,
Sci. Rep. 7 (1) (2017) 1–14.
[42] S. Salimzadeh, A. Paluszny, H.M. Nick, R.W. Zimmerman, A three-dimensional coupled thermo-hydro-mechanical model for deformable
fractured geothermal systems, Geothermics 71 (2018) 212–224.
[43] Y. Wang, T. Li, Y. Chen, G. Ma, A three-dimensional thermo-hydro-mechanical coupled model for enhanced geothermal systems (EGS)
embedded with discrete fracture networks, Comput. Methods Appl. Mech. Engrg. 356 (2019) 465–489.
[44] X. Cui, L.N.Y. Wong, A 3D thermo-hydro-mechanical coupling model for enhanced geothermal systems, Int. J. Rock Mech. Min. Sci.
143 (2021) 104744.
[45] COMSOL Multiphysics, COMSOL Multiphysics User’s Guide, 2012.
[46] O. Kolditz, S. Bauer, L. Bilke, N. Böttcher, J.O. Delfs, T. Fischer, et al., Opengeosys: an open-source initiative for numerical simulation
of thermo-hydro-mechanical/chemical (THM/C) processes in porous media, Environ. Earth Sci. 67 (2) (2012) 589–599.
[47] P.C. Lichtner, G.E. Hammond, C. Lu, S. Karra, G. Bisht, B. Andre, R.T. Mills, J. Kumar, J.M. Frederick, PFLOTRAN User Manual,
Technical Report, Sandia National Laboratories, 2019.
[48] A. Wilkins, C.P. Green, J. Ennis-King, An open-source multiphysics simulation code for coupled problems in porous media, Comput.
Geosci. 154 (2021) 104820.
[49] Open Porous Media, Open porous media, 2019, http://www.opm-project.org.
[50] A.N. Brooks, T.J. Hughes, Streamline upwind/Petrov–Galerkin formulations for convection dominated flows with particular emphasis
on the incompressible Navier–Stokes equations, Comput. Methods Appl. Mech. Engrg. 32 (1–3) (1982) 199–259.
[51] M. Cacace, A.B. Jacquey, Flexible parallel implicit modelling of coupled thermal–hydraulic–mechanical processes in fractured rocks,
Solid Earth 8 (5) (2017) 921–941.
[52] R.W. Zimmerman, Coupling in poroelasticity and thermoelasticity, Int. J. Rock Mech. Min. Sci. 37 (1–2) (2000) 79–87.
[53] P.A. Vermeer, A. Verruijt, An accuracy condition for consolidation by finite elements, Int. J. Numer. Anal. Methods Geomech. 5 (1)
(1981) 1–14.
[54] M.A. Biot, Thermoelasticity and irreversible thermodynamics, J. Appl. Phys. 27 (3) (1956) 240–253.
[55] T.J. Hughes, A multidimentional upwind scheme with no crosswind diffusion, in: Finite Element Methods for Convection Dominated
Flows, AMD, 1979, p. 34.
[56] W. Cui, K.A. Gawecka, D.M.G. Taborda, D.M. Potts, L. Zdravković, Time-step constraints in transient coupled finite element analysis,
Internat. J. Numer. Methods Engrg. 106 (12) (2016) 953–971.
[57] Y. Saad, M.H. Schultz, GMRES: A generalized minimal residual algorithm for solving nonsymmetric linear systems, SIAM J. Sci.
Stat. Comput. 7 (3) (1986) 856–869.
[58] R. Fletcher, Conjugate gradient methods for indefinite systems, in: Numerical Analysis, Springer, Berlin, Heidelberg, 1976, pp. 73–89.
[59] P Sonneveld, CGS, a fast lanczos-type solver for nonsymmetric linear systems, SIAM J. Sci. Stat. Comput. 10 (1) (1989) 36–52.
[60] H.A. Van der Vorst, Bi-CGSTAB: A fast and smoothly converging variant of bi-CG for the solution of nonsymmetric linear systems,
SIAM J. Sci. Stat. Comput. 13 (2) (1992) 631–644.
[61] G.L. Sleijpen, D.R. Fokkema, Bicgstab (L) for linear equations involving unsymmetric matrices with complex spectrum, Electron.
Trans. Numer. Anal. 1 (1993) 11–32.
[62] R. Barrett, M. Berry, T.F. Chan, J. Demmel, J. Donato, J. Dongarra, et al., Templates for the Solution of Linear Systems: Building
Blocks for Iterative Methods, Society for Industrial and Applied Mathematics, 1994.
[63] M.T. Jones, P.E. Plassmann, An improved incomplete cholesky factorization, ACM Trans. Math. Softw. 21 (1) (1995) 5–17.
[64] Y. Gu, D.L. O’neal, An analytical solution to transient heat conduction in a composite region with a cylindrical heat source, J. Solar
Energy Eng. 117 (3) (1995) 242–248.
[65] J. Donea, S. Giuliani, H. Laval, Accurate explicit finite element schemes for convective-conductive heat transfer problems, in: Finite
Element Methods for Convection Dominated Flows, 14, 1979, pp. 9–166.
[66] B.M. Das, Advanced Soil Mechanics, fifth ed., CRC Press, 2019.
[67] R.A. Muller, S. Finsterle, J. Grimsich, R. Baltzer, E.A. Muller, J.W. Rector, et al., Disposal of high-level nuclear waste in deep
horizontal drillholes, Energies 12 (11) (2019) 2052.
[68] J. Sundberg, P.E. Back, R. Christiansson, H. Hökmark, M. Ländell, J. Wrafter, Modelling of thermal rock mass properties at the
potential sites of a Swedish nuclear waste repository, Int. J. Rock Mech. Min. Sci. 46 (6) (2009) 1042–1054.
[69] X. Zhou, A. Zhou, D.A. Sun, Three-dimensional thermal analysis of the repository for high-level radioactive nuclear waste, Int. J.
Energy Res. 44 (10) (2020) 8208–8220.
[70] Y.C. Zeng, N.Y. Wu, Z. Su, J. Hu, Numerical simulation of electricity generation potential from fractured granite reservoir through a
single horizontal well at yangbajing geothermal field, Energy 65 (2014) 472–487.
37
X. Cui and L.N.Y. Wong Computer Methods in Applied Mechanics and Engineering 394 (2022) 114939

[71] Z. Feng, Y. Zhao, A. Zhou, N. Zhang, Development program of hot dry rock geothermal resource in the yangbajing basin of China,
Renew. Energy 39 (1) (2012) 490–495.
[72] Y. Yuan, T. Xu, Z. Jiang, B. Feng, Prospects of power generation from the deep fractured geothermal reservoir using a novel vertical
well system in the yangbajing geothermal field, China, Energy Rep. 7 (2021) 4733–4746.
[73] Q. Guo, Y. Wang, W. Liu, Major hydrogeochemical processes in the two reservoirs of the yangbajing geothermal field, Tibet, China,
J. Volcanol. Geotherm. Res. 166 (3–4) (2007) 255–268.

38

View publication stats

You might also like