Numerical Simulation Beatrice

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

J182607 DOI: 10.

2118/182607-PA Date: 8-June-18 Stage: Page: 1 Total Pages: 18

Direct Numerical Simulation of Flow


on Pore-Scale Images Using
the Phase-Field Method
Florian Frank and Chen Liu, Rice University; Faruk O. Alpak, Shell International Exploration and Production
and Rice University; Steffen Berg, Shell Global Solutions International and Imperial College London; and
Beatrice Riviere, Rice University

Summary
Advances in pore-scale imaging, increasing availability of computational resources, and developments in numerical algorithms have
started rendering direct pore-scale numerical simulations of multiphase flow on pore structures feasible. In this paper, we describe a
two-phase-flow simulator that solves mass- and momentum-balance equations valid at the pore scale (i.e., at scales where the Darcy ve-
locity homogenization starts to break down). The simulator is one of the key components of a molecule-to-reservoir truly multiscale
modeling work flow.
A Helmholtz free-energy-driven, thermodynamically based diffuse-interface/phase-field method is used for the effective simulation
of numerous advecting interfaces, while honoring the interfacial tension (IFT). The advective Cahn-Hilliard (CH) (mass-balance,
energy dissipation) and Navier-Stokes (NS) (momentum-balance, incompressibility) equations are coupled to each other within the
phase-field framework. Wettability on rock/fluid interfaces is accounted for by means of an energy-penalty-based wetting (contact-
angle) boundary condition. Individual balance equations are discretized by use of a flexible discontinuous Galerkin (DG) method.
The discretization of the mass-balance equation is semi-implicit in time using a convex/concave splitting of the energy term. The
momentum-balance equation is split from the incompressibility constraint by a projection method and linearized with a Picard splitting.
Mass- and momentum-balance equations are coupled to each other by means of operator splitting, and are solved sequentially.
We discuss the mathematical model and its DG discretization, and briefly introduce nonlinear and linear solution strategies.
Numerical-validation tests show optimal convergence rates for the DG discretization, indicating the correctness of the numerical
scheme and its implementation. Physical-validation tests demonstrate the consistency of the phase distribution and velocity fields simu-
lated within our framework. Finally, two-phase-flow simulations on two real pore-scale images demonstrate the usefulness of the pore-
scale simulator. The direct pore-scale numerical-simulation methodology rigorously considers the flow physics by directly acting on
pore-scale images of rocks without remeshing. The proposed method is accurate, numerically robust, and exhibits the potential for tack-
ling realistic problems.

Introduction
Until recently, direct observation of fluid flow at the pore scale was not possible inside a natural rock sample. Flow experiments were
possible only on 2D micromodels etched in glass or in polyester resin where the pore structure was mimicked by idealized pore bodies
and connecting channels (Lenormand et al. 1983; Lenormand and Zarcone 1988; Karadimitriou and Hassanizadeh 2012). This well-
defined geometry also was accessible to direct computation. The simplest approach was to compute the hydraulic conductivities in indi-
vidual channels and determine the total conductivity of the network, resembling an electrical circuit, by applying Kirchhoff’s laws on
hydraulic conductivities. Early versions of pore-network models were the bundle-of-tubes family of models (Purcell 1949). The first
interconnected networks were the work of Fatt (1956a, b, and c). One of the earliest micromodels is discussed in Chatenever and
Calhoun et al. (1952). The use of micromodels gained widespread acceptance with the work of Lenormand et al. (1983) and Lenormand
and Zarcone (1988). Later, this approach was extended into 3D rocks and referred to as “pore-network modeling” (Blunt 2001; Blunt
et al. 2002, 2013). Even though the earliest pore-network models date approximately 2 decades back, the current state-of-the-art pore-
network models are, to some extent, predictive, and the development is actively driven, including commercial providers (Ryanazov
et al. 2009; Joekar-Niasar et al. 2010; Joekar-Niasar and Hassanizadeh 2011). Despite their success, pore-network models provide an
indirect representation of the rock. This is because the actual pore structure is abstracted into an artificial “equivalent” network that
may lack much of the complexity of the rock. In addition to the purely geometrical abstraction, there is also a conceptual abstraction. A
pore-network model only delivers what has been considered in terms of physical mechanisms. In most cases, pore-network models do
not follow directly the first-principle physics but are rather based on individual mechanisms that are believed to be important. A direct
validation of these assumptions is often not possible.
Microcomputed-tomography (mCT)-scanning technology started in the late 1980s and early 1990s (Flannery et al. 1987; Spanne
et al. 1994). Driven by major developments in the field of mCT scanning (Wildenschild et al. 2002; Berg et al. 2013; Cnudde and Boone
2013; Wildenschild and Sheppard 2013; Bultreys et al. 2016) and by increased availability of computational resources and capabilities,
recent research has been focusing on numerical modeling of flow at the pore scale on mCT images (Blunt et al. 2013). The basic idea of
this approach is to make use of 3D images at pore-scale resolution or better (e.g., from mCT scanning) and to simulate fluid flow at the
pore-scale geometry, typically by use of computational-fluid-dynamics (CFD) based techniques.
Direct-numerical-simulation (DNS) approaches have been becoming relatively more popular because no a priori choice between the
level of rigor and captured phenomena is made. In DNS, capillary and viscous forces act at the same time. Thus, depending on the
choice of flow parameters, both capillary- and viscous-dominated flows can be captured rigorously. DNS enables the description of a
wide range of flow regimes and simulation of a wide range of pore-scale dynamics such as cooperative and/or nonlocal displacement
processes. Berg et al. (2016) states that such nonlocal processes have been observed during drainage in 2D micromodels (Armstrong

Copyright V
C 2018 Society of Petroleum Engineers

This paper (SPE 182607) was accepted for presentation at the SPE Reservoir Simulation Symposium, Montgomery, Texas, USA, 20–22 February 2017, and revised for publication. Original
manuscript received for review 24 July 2017. Revised manuscript received for review 19 February 2018. Paper peer approved 21 February 2018.

2018 SPE Journal 1

ID: jaganm Time: 17:21 I Path: S:/J###/Vol00000/180034/Comp/APPFile/SA-J###180034


J182607 DOI: 10.2118/182607-PA Date: 8-June-18 Stage: Page: 2 Total Pages: 18

and Berg 2013) and during imbibition (Berg et al. 2013) with fast mCT (e.g., Berg et al. 2014; Andrew et al. 2015). However, from a
computational perspective, DNS techniques are notably more expensive than pore-network models. For single-phase flow, this method
is relatively well-established and, to some extent, commercially available in software packages. For two-phase flow, the underlying
transport phenomena are substantially more complex.
For two-phase flow, both DNS and pore-network modeling approaches have their own merits: When pore-scale simulation is per-
formed with the purpose of deriving flow parameters relevant for the Darcy scale, it is implicitly assumed that the flow regime is cov-
ered within its relevant characteristics and relevant details. Although direct simulation is quite robust and accurate at the fine-scale side,
it does not currently reach the length scales multiple times larger than internal characteristic length scales (e.g., the cluster length). In
terms of saturation, we do not reach a representative elementary volume in a few millimeter-large samples for two-phase flow with
DNS. On the other hand, pore-network modeling is more approximate in terms of fine-scale physics. It can, however, reach larger
length scales to capture the flow regimes. Our vision for DNS is to use it for developing a conceptual understanding and ensuing rules
to bridge the gap between pore and Darcy scales, and supply this information into more-approximate approaches, such as pore-network
models, and thereby render them more physically consistent for Darcy-scale modeling.
Several approaches have been developed for simulating flow at molecular and pore scales, including the lattice gas method and the
lattice Boltzmann method (LBM) (e.g., Chen and Doolen 1998; Porter et al. 2009; Chen et al. 2013; Boek et al. 2017), Monte Carlo
models (e.g., Li and Sultan 2015), molecular dynamics (e.g., Moseler and Landman 2000; Nie et al. 2004; Sedghi et al. 2014), smoothed
particle hydrodynamics (SPH) (e.g., Kunz et al. 2016; Sivanesapillai et al. 2016), dissipative particle dynamics (DPD) (e.g.,
Hoogerbrugge and Koelman 1992; Pagonabarraga and Frenkel 2001; Li et al. 2011), and Eulerian CFD. The latter family of techniques
includes the front-tracking method (e.g., Glimm et al. 1999; Unverdi and Tryggvason 1992); the volume of fluid method (e.g., Hirt and
Nichols 1981; Rider and Koth 1998; Ubbink and Issa 1999; Darwish and Moukalled 2006; Raeini et al. 2012, 2014a, b; Ferrari and
Lunati 2014); the level-set method (e.g., Sethian 1999; Adalsteinsson and Sethian 1999; Enright et al. 2002; Osher and Fedkiw 2003;
Prodanovic and Bryant 2006; Nourgaliev and Theofanous 2007; Prodanovic et al. 2010; Abu-Al-Saud et al. 2017); the phase-field
method (e.g., Antanovskii 1995; Jacqmin 1996; Anderson et al. 1998; Abels et al. 2017; Jacqmin 1999; Badalassi et al. 2003; Yue et al.
2004; Ding et al. 2007; Kim 2012; Chen and Shen 2016). A detailed survey of techniques for simulating pore-scale flow can be found
in Meakin and Tartakovsky (2009). There are many ways to categorize the various numerical techniques to simulate two-phase flow at
the pore scale. If one only focuses on relatively more-practical methods for direct two-phase pore-scale simulation in actual porous
media, LBM (e.g., Boek et al. 2017), volume of fluid method (e.g., Ferrari and Lunati 2014; Raeini et al. 2014b), and level-set method
(e.g., Prodanovic and Bryant 2006) stand out as popular techniques. More recently, the phase-field method and its direct hydrodynamics
multicomponent extensions started to emerge as a practically applicable and thermodynamically consistent family of techniques for
simulating multiphase flow in real porous media (e.g., Koroteev et al. 2014; Alpak et al. 2017).
In general, for multiphase-flow modeling, one needs to combine a method to compute mass and momentum balance and a method to
represent the moving liquid/liquid interfaces. Concerning the momentum balance, the Eulerian CFD family of techniques and the LBM
technique are relatively more commonly used for simulating fluid-flow phenomena at the pore scale. When it comes to representing the
moving liquid/liquid interfaces, there is a range of methods—from interface-tracking methods to diffuse-interface methods—that has
specific pros and cons, and it is generally advantageous to use a physically based method that represents curvature and capillarity very
well. Some methods, such as the LBM method, combine a Lagrangian momentum balance with a diffuse-interface method by default,
which combines the advantage of a diffuse-interface method and a numerically efficient momentum balance, but the disadvantage of—
depending on exact formulation—requiring many adjustable parameters that need to be calibrated to the physical system.
The main equations constituting immiscible two-phase flow in the diffuse-interface framework are the advective CH equation
describing the mass conservation and interface propagation subject to an energy-minimization concept, and an extension of the NS
equations describing the momentum balance in the presence of two immiscible fluids. CH and NS equations are coupled to each other
within a thermodynamically based diffuse-interface modeling framework. As such, the considered mathematical model is a member of
the phase-field branch of the Eulerian CFD family of models.
In this paper, the CH and the NS equations are spatially discretized using the DG method (Riviere 2008) for increased accuracy in local
conservation and flexibility in terms of the order of the approximation. Several DG schemes are reported for the solution of the CH equa-
tion (e.g., Wells et al. 2006; Feng and Karakashian 2007; Kay et al. 2009; Aristotelous et al. 2013; Frank et al. 2018a) and variants of the
NS equations (e.g., Girault et al. 2005; Montlaur et al. 2009; Persson et al. 2009; Luo et al. 2012) separately. Typically, coupled CH/NS
equations are solved with finite-difference/finite-volume schemes (e.g., Badalassi et al. 2003; Ding et al. 2007; Ding and Spelt 2007;
Alpak et al. 2016, 2017). Several references document the use of continuous finite-element methods and mixed finite-element methods
(e.g., Boyer et al. 2010; Bao et al. 2012), respectively. Recently, Pigeonneau and Saramito (2016) reported a DG-based finite-element
method for the coupled solution of CH and Stokes equations (neglecting fluid inertia), and applied it to 2D open-domain problems.
The DG framework for pore-scale two-phase-flow simulation described in this paper facilitates the simultaneous coupled solution of
CH and NS equations for flow in complex 3D pore-scale images as well as on 3D open domains. The unconditionally stable temporal dis-
cretization of the CH equation is semi-implicit in time using a classic convex/concave splitting protocol (Elliott and Stuart 1993; Eyre
1998). The momentum-balance equation is split from the incompressibility constraint by a projection method, and linearized with a Picard
splitting. Mass- and momentum-balance equations are coupled to each other by means of operator splitting, and are solved sequentially.
The outline of this paper is as follows: We first introduce and discuss the mathematical model including the CH and the NS equa-
tions. In the numerical-solution section that follows, we describe the meshing approach, and the temporal and spatial discretizations
separately for mass- and momentum-conservation equations. We then report numerical-validation tests that indicate optimal conver-
gence rates for the DG discretization of the CH equation. We next focus on physical problems, showing the accuracy of the momen-
tum-balance component of the CH/NS solver on a Poiseuille flow problem by solely testing the NS solver. Next, we investigate a
simple, yet intuitive, thermodynamic-system evolution problem, which validates the energy-minimization attribute of the CH/NS
solver. Two sets of numerical tests are discussed in the following section. The first set of numerical simulations features a symmetric
crossing-ducts geometry, which is designed to test the fundamental physical characteristics of two-phase flow. The second set features
the DNS of two-phase flow on two real pore images. The paper is closed with a section on the summary of the work and conclusions.

Mathematical Model
We consider the isothermal flow of two immiscible fluid phases in a pore-scale medium. By making the continuum assumption for the
involved fluid phases, the mathematical model of pore-scale flow is described by a mass-conservation and a momentum-conservation
equation. In addition to the isothermal-flow assumption, we assume that the heat generated from the friction of flowing-fluid phases is
negligible; both assumptions, in turn, allow the elimination of an otherwise necessary energy-conservation equation. On the other

2 2018 SPE Journal

ID: jaganm Time: 17:21 I Path: S:/J###/Vol00000/180034/Comp/APPFile/SA-J###180034


J182607 DOI: 10.2118/182607-PA Date: 8-June-18 Stage: Page: 3 Total Pages: 18

hand, we use a thermodynamically-based energy-minimizing model in conjunction with a diffuse-interface model to describe the
propagating-fluid interfaces.

Advective CH Equation. The CH model characterizes phase segregation, which is the alignment of a thermodynamic system into spa-
tial domains predominated by one of the two components, in the presence of a mass constraint and the dissipation of energy (Cahn and
Hilliard 1958, 1959, 1971; Cahn 1959, 1961). The main unknown of the model is the order parameter c ¼ cðt; xÞ, which can be either a
volume fraction or a mass fraction of one of the two components c1 , c2 {i.e., c ¼ c1 , c2 ¼ 1  c1 , where cðt; xÞ 2 ½0; 1 (i.e.,
c1 þ c2 ¼ 1)}, or the difference between fractions {i.e., c ¼ c1  c2 ¼ 2c1  1, where cðt; xÞ 2 ½1; 1}. The latter convention is used
throughout this paper.
The total Helmholtz free energy F for an incompressible binary isotropic mixture in a domain X  R3 is given by (Carlson et al.
2011; Carlson 2012):
ð ð
a
FðcÞ :¼ bWðcÞ þ j$cj2 dx þ ðrBs  rAs ÞvðcÞ þ rAs dsx ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð1Þ
X 2 @X

where bWðcÞ is called “chemical energy density or homogeneous free-energy density” corresponding to the Helmholtz free energy of a
unit volume of homogeneous material of composition c. The volume term in Eq. 1 was proposed by Cahn and Hilliard (1958), who con-
sidered incompressible binary alloys at constant temperature. The idea behind these choices is that the bulk free-energy density function
WðcÞ has two minima that correspond to each bulk component [i.e., c 2 f1; 1g], provided that the (constant) system temperature is
below the critical temperature of the mixture (i.e., where a homogeneous mixture becomes energetically unfavorable). A popular form
for WðcÞ is given next, and the definition of vðcÞ depends on this form (Carlson 2012):

1
WðcÞ :¼ ðc  1Þ2 ðc þ 1Þ2 ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð2aÞ
4
1
vðcÞ :¼ ðc3  3c þ 2Þ: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð2bÞ
4

The second component of the volume term is called “gradient energy density or interfacial energy density,” where the integrand
a
j$cj2 (with 0 < a  1) is the first term of an expansion. It represents the increase in free energy caused by the presence of a composi-
2
tion gradient across the interface between two bulk components. The integrand in the surface term in Eq. 1 models a surface-energy
density, where rAs is the surface tension at Phase A ðc ¼ 1Þ and a solid surface (e.g., rock wall). Analogously, rBs denotes the surface
tension at Phase B ðc ¼ 1Þ and a solid surface. The function vðcÞ is a blending between rAs and rBs . The angle between the (center
of) the diffuse interface and a solid surface (wall/rock) is called “contact angle” (also called the “wetting angle”) h 2 ½0; p. We take
Phase A with c ¼ 1 as the reference phase in our contact-angle convention. Thus, the contact angle in our numerical scheme is defined
as the angle at the three-phase contact line between the interface and the solid surface toward Phase A. The contact angle reflects the
macroscopic wetting characteristics of solid surfaces. The relationship between the contact angle h and the IFT r between two fluid
phases is given by Young’s equation:
rBs  rAs
cosðhÞ ¼ : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð3Þ
r

The factors a and b can be expressed in terms of the IFT r and the interface parameter e as follows:

3 3 r
a :¼ pffiffiffi re; b :¼ pffiffiffi : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð4Þ
2 2 2 2e
Occasionally, in the literature, e is referred to as the interface width, which is misleading since the interface width is approximately 4e
(Wodo and Ganapathysubramanian 2011). The above-described way of prescribing a contact angle is called the energy-based approach
in the literature (e.g., Jacqmin 2000; Ding and Spelt 2007; Carlson 2012). A geometric approach is derived by using the tangent of the
interface and the gradient of the order parameter in Ding and Spelt (2007). The geometric contact-angle approach is extended to handle
nonequilibrium flows in Alpak et al. (2016). Although the previous references use a finite-volume discretization scheme, the geometric
contact-angle modeling approach did not operate reliably in our DG implementation.
Let X denote a spatial domain such that X  R3 , and let J :¼ ð0; tf Þ denote the considered time interval with end (final) time
tf 2 Rþ . The CH equation is designed such that its solution c is a minimizer of F as given in Eq. 1. The chemical potential arises within
this context. It equals the variational/functional derivative of the volume term in the total Helmholtz free energy with respect to c:
ð 
d a 2
l :¼ bWðcÞ þ j$cj dx ¼ bW0 ðcÞ  aDc: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð5Þ
dc X 2
The stationary state then satisfies the condition

l ¼ constant: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð6Þ

In addition, the first variation of F with respect to c leads to a boundary condition on solid surfaces, @Xwall , to enforce a stationary state:

a$c  n ¼ ðrAs  rBs Þv0 ðcÞ ¼ rcosðhÞv0 ðcÞ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð7Þ

where n denotes the unit normal on the boundary @X exterior to the domain X, whereas Young’s equation (Eq. 3) was used to
describe the IFT. The assumption here is that the solution c field satisfies the equilibrium contact angle on solid surfaces at every time
point t 2 J.

2018 SPE Journal 3

ID: jaganm Time: 17:21 I Path: S:/J###/Vol00000/180034/Comp/APPFile/SA-J###180034


J182607 DOI: 10.2118/182607-PA Date: 8-June-18 Stage: Page: 4 Total Pages: 18

A chemical-potential gradient implies a mass flux j :¼ M$l, where the phenomenological coefficient M  0 is called mobility.
Conservation is described by @t c þ $  ðvcÞ  $  j ¼ 0, where v is a solenoidal velocity field. Combining these equations with Eqs. 5 and
7 and adding appropriate boundary and initial conditions yields the following system of partial-differential equations:

@t c  $  ðM$lÞ þ $  ðvcÞ ¼ 0; where l ¼ bW0 ðcÞ  aDc; in J  X; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð8aÞ

a$c  n ¼ rcosðhÞv0 ðcÞ; on J  ð@Xout [ @Xwall Þ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð8bÞ

c ¼ cin ; on J  @Xin ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð8cÞ

M$l  n ¼ 0; on J  @X; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð8dÞ

c ¼ c0 ; on f0g  X: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð8eÞ

The coupling of the CH equation to the NS equation is through the velocity field v. Conditions Eq. 8b and Eq. 8c are Neumann on the
outflow boundary @Xout and the solid (rock) boundary @Xwall and Dirichlet on the inflow boundary @Xin , respectively. The outflow
boundary @Xout consists of points x 2 @X such that vðxÞ  n > 0, and we set @Xin :¼ @Xnð@Xout [ @Xwall Þ. Further details of the CH
equation (e.g., properties of the continuous solution, and others) are documented in Frank et al. (2018a). A novel correction technique is
developed for energy-based wetting-boundary conditions on jagged/rough surfaces (typical of mCT images) in Frank et al. (2018b) for
accurate two-phase-flow simulations with the phase-field method. A special solver for the DG discretization of the CH equation is pre-
sented in Thiele et al. (2017).

Incompressible NS Equations. We introduce the momentum-balance equation for incompressible fluids. This equation is coupled to a
mass-balance system described by Eq. 8 for two-component immiscible and incompressible binary mixtures.
Let us consider an incompressible fluid mixture occupying a domain X  R3 . Let the unknown velocity v : J  X ! R3 satisfy the
momentum-balance equation with an incompressibility constraint:

@t ðqvÞ þ $  ðqv vÞ  2$  ½ls eðvÞ ¼ $  Tc þ f ; in J  X; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð9aÞ

$v¼0 in J  X: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð9bÞ
1
In Eq. 9a, the term eðvÞ denotes the deformation tensor, which is defined as eðvÞ :¼ ½$v þ ð$vÞT . The shear viscosity ls is nonnega-
2
tive. For single-phase Newtonian fluids, ls is constant. For two immiscible fluid phases separated by diffuse interfaces (whereby mixing
is restricted only to the immediate vicinity of the interfaces), we use the following mixing rule for the density:
1þc 1c
q¼ qA þ qB ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð10Þ
2 2
where qA and qB denote the mass densities of Phase A and Phase B, respectively. The vector-valued force density f stems from the
external forces acting on the fluid mixture—for instance, f ¼ qg, when gravity is considered. The term Tc denotes the capillary stress
tensor, the divergence of which can be expressed as

$  Tc ¼ $p þ l$c; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð11Þ

where p denotes the pressure. Badalassi et al. (2003) propose the coupling term l$c, whereas other authors use the coupling term –c$l
in the presence of a solenoidal velocity field (e.g., Ding and Spelt 2007). The equivalence between these two terms is shown (e.g., Ding
and Spelt 2007). We use a projection algorithm to linearize and decouple the incompressible NS equations (Guermond et al. 2006),
which is described next.

Numerical Solution
We describe the main features of the numerical-solution technique implemented for CH and NS equations, and their coupling for direct
two-phase simulation on pore-scale image volumes.

Meshing for lCT Images. The mCT can create cross sections of small rock samples on micrometer to millimeter scale by means of
X-rays, which subsequently can be used to construct a model by means of 3D image processing and segmentation (e.g., Georgiadis
et al. 2013). In our case, this model consists of a set of binary voxels, each of which is associated either with the pore space of the rock
sample or with its solid rock matrix. We directly simulate mass and momentum balance on the voxels that form the pore space. Because
every voxel is a regular cube of the same size, the voxel types (solid/rock and pore/fluid) are efficiently stored as a 3D Boolean array.
No further grid storage is required as the voxels form a Cartesian grid. Simulating directly on the image data in the form of a voxel sub-
set is common in numerical pore-scale modeling (e.g., Koroteev et al. 2014; Boek et al. 2017; Alpak et al. 2017) because it simplifies
the simulation work flow and eliminates mesh-generation-related uncertainties. In some cases, the pore space is remeshed with an
unstructured grid to obtain a smooth-surface approximation (e.g., Bogdanov et al. 2012).
We denote the voxels that are associated with the pore-space as “elements”; Ek , k 2 f0; …; Nel  1g [i.e., eh :¼ fEk gk ] is a nonover-
lapping partition of X. We denote by h the edge length of one cubic element. The computational domain Xh may consist of several
nonconnected subdomains. Faces that are a subset of the boundary, @Xh , of the domain are grouped into two sets: Cwall h containing faces
that lie between an element and a solid voxel and faces on @½0; 13 \ @Xh that are associated with an impermeable solid surface (wall)
3
boundary, and Cext int
h containing faces on @½0; 1 \ @Xh that are associated with an exterior fluid domain (Fig. 1). Let Ch denote the set
of interior faces (i.e., faces connecting two fluid elements). The resulting meshing eh has the structure of a pseudo-Cartesian grid, which
is exploited in the implementation.

4 2018 SPE Journal

ID: jaganm Time: 17:21 I Path: S:/J###/Vol00000/180034/Comp/APPFile/SA-J###180034


J182607 DOI: 10.2118/182607-PA Date: 8-June-18 Stage: Page: 5 Total Pages: 18

Γwall
h
Γint
h

Γext
h

Fig. 1—Examples of computational domains Xh  [0, 1]3 illustrated in gray: 1003 -voxel l. CT image of Berea sandstone (left), 1003 -
voxel sphere pack (middle), 23 -voxel example containing four elements illustrating the face sets Cint ext wall
h , Ch , and Ch (right). In the lat-
ter case, faces on ›[0, 1]3 \ ›Xh can be declared to be part of Cint ext
h or of Ch .

Discretization of the Advective CH Equation. The dimensionless version of the model problem in Eq. 8 follows:
1
@t c  $  ðM $lÞ þ $  ðvcÞ ¼ 0; in J  X; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð12aÞ
Pe
l ¼ W0 ðcÞ  Cn2 Dc; in J  X; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð12bÞ

where Pe is the Péclet number and Cn the Cahn number. Let xc denote the characteristic length and vc the characteristic velocity. The
definitions of Pe and Cn are given by
pffiffiffi
2 2vc xc e e
Pe ¼ ; Cn ¼ : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð13Þ
3Mr xc
We describe briefly the discretization of the next CH equation.
Temporal Discretization. Let J be decomposed into Nst intervals of length s. For temporal discretization of the CH equation, a pop-
ular unconditionally stable method that relies on the convex/concave splitting of WðcÞ is adopted in this work. The splitting has the fol-
lowing form:

WðcÞ ¼ Wþ ðcÞ þ W ðcÞ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð14Þ

where the convex part Wþ is treated time implicitly and the concave part W is treated time explicitly (Elliott and Stuart 1993; Eyre
1998). If Wþ is nonlinear, the convex/concave splitting technique results in a nonlinear system. After discretization in time, the CH
problem follows:

Cn2 Dcn  ln þ W0þ ðcn Þ ¼ W0 ðcn1 Þ ; in X; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð15aÞ


s
cn  $  ðM $ln Þ þ s$  ½vðtn Þcn  ¼ cn1 ; in X; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð15bÞ
Pe
subjected to respective initial and boundary conditions (Eqs. 8b through Eq. 8e).
Spatial Discretization. The differential operators are discretized in space using the interior-penalty DG method. The discrete-order
parameter ch and the chemical potential lh belong to the space of discontinuous polynomials of degree r, denoted by Pr ðeh Þ.
The fully discrete problem for the CH equation is defined as follows: For n ¼ 1; …; Nst , find cnh ; lnh 2 Pr ðeh Þ such that
8wh 2 Pr ðeh Þ,

Cn2 adiff ð1; cnh ; wh Þ þ Cn2 bdiff ðcnh ; wh Þ  ðlnh ; wh Þ þ ½W0þ ðcnh Þ; wh  ¼ Cn2 ddiff ðwh Þ  ½W0 ðcn1
h Þ; wh ; . . . . . . . . . . . . . . . . . . ð16aÞ
s diff
ðcnh ; wh Þ þ a ðM; lnh ; wh Þ þ saadv;n ðunh ; cnh ; wh Þ ¼ ðcn1
h ; wh Þ þ sb
adv n
ðuh ; wh Þ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð16bÞ
Pe
ðc0h ; wh Þ ¼ ðc0 ; wh Þ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð16cÞ

where ð; Þ denotes the L2 ðXÞ inner product. System Eq. 16 is linear in lnh and nonlinear in cnh . In the following equations, $ should be
interpreted as the broken gradient; that is, for all E in eh , ð$h wÞjE ¼ $ðwjE Þ while suppressing the subscript h. The form adiff is the clas-
sical symmetric interior-penalty Galerkin (SIPG) bilinear form for the operator $  ðz$cÞ,
ð ð ð
r
adiff ðz; c; wÞ :¼ z$c  $w  Re2Cint ðfz$c  ne g½½w þ fz$w  n e g½½cÞ þ R e2Cint ½½c½½w; . . . . . . . . . . . . . . . . . . . . . ð17aÞ
h
e h h
e
X

where r > 0 denotes the penalty parameter. With r chosen sufficiently large, coercivity of the bilinear form adiff ðz; ; Þ is ensured
(Riviere 2008). Averages and jumps of a discontinuous function w across faces are denoted by fwg and ½½w.

2018 SPE Journal 5

ID: jaganm Time: 17:21 I Path: S:/J###/Vol00000/180034/Comp/APPFile/SA-J###180034


J182607 DOI: 10.2118/182607-PA Date: 8-June-18 Stage: Page: 6 Total Pages: 18

The boundary conditions (Eqs. 8b through 8d) are discretized by the forms
ð ð
diff r
b ðc; wÞ :¼  ð$c  nw þ $w  ncÞ þ cin w; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð17bÞ
@Xinh
h @Xinh
ð pffiffiffi ð ð
2 2 1 r
ddiff ðwÞ :¼  $w  ncin  cosðhÞ v0 ðcÞ þ cin w: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð17cÞ
@Xin
h
3 Cn @X out
h
h @X in
h

Throughout this paper, we use r :¼ 2r . The forms for the advection operator $  ðcvÞ at tn 2 J subject to the boundary condition (Eq.
8c) are given by
ð X ð ð
"
aadv ðv; c; wÞ :¼  cv  $w þ e2C int c fv  n e g½½w þ cv  nw; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð18aÞ
X h e @Xout
h
ð
badv ðv; wÞ :¼  cin v  nw: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð18bÞ
@Xin
h

The discretization of the CH equation yields a nonlinear system with saddle-point structure, which is transformed by a Schur-
complement reduction. The resulting nonlinear system is solved with an inexact Newton method. Further details of the numerical solu-
tion of the CH equation (e.g., approximation spaces, matrix formulation, and linearization) can be found in Thiele et al. (2017) and
Frank et al. (2018a). The discrete velocity unh is defined in Eq. 23e.
Discretization of the Incompressible NS Equations. The dimensionless version of Eq. 9 with inflow, wall, and outflow boundary
conditions, and initial condition for the velocity, follows:
1 1 3 1 1
@t ðqtÞ þ $  ðqt tÞ  $  ½$t þ ð$tÞT  ¼  $p þ pffiffiffi l$c þ 2 f ; in J  X; . . . . . . . . . . . . . ð19aÞ
Re Re Ca 2 2 Re Ca Cn Fr
$  v ¼ 0; in J  X; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð19bÞ

v ¼ vD ; on J  @Xin ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð19cÞ

v ¼ 0; on J  @Xwall ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð19dÞ
 
1
$v þ ð$vÞT  pI  n ¼ 0 ; on J  @Xout ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð19eÞ
Ca
v ¼ v0 ; on f0g  X; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð19fÞ

where we use the same symbols for dimensionless quantities. Note that we use a Dirichlet boundary condition on inflow boundary @Xin and
a Neumann-type boundary (also called open boundary) condition on outflow boundary @Xout . In Eq. 19a, Re denotes the Reynolds number,
Ca the capillary number, and Fr the Froude number: Re ¼ qc vc xc =ls , Ca ¼ ls vc =r, and Fr2 ¼ qc v2c =xc fc . We recall that the Cahn number
Cn is defined in Eq. 13. Additional characteristic variables are the characteristic mass density qc and characteristic force density fc .
Temporal Discretization. Recall that s denotes the timestep length. We apply the incremental pressure-correction scheme in rota-
tional form, choosing first order in time (Guermond et al. 2006) to discretize the NS equation in time. This yields a sequence of two
“linear” equations—a momentum-balance equation without incompressibility constraint yielding a nonsolenoidal velocity, and an aux-
iliary Poisson-type problem—and explicit update rules for pressure and solenoidal velocity. First, we solve the incompressible velocity
equation as follows:
s s
qn vn þ s yðqn ; vn ; vn1 Þ  $  ½$vn þ ð$vn ÞT  ¼ qn1 vn1  $ðpn1 þ un1 Þ
Re Re Ca
3 1 s
þ s pffiffiffi ln $cn þ 2 f n ; in X;                                               ð20aÞ
2 2 Re Ca Cn Fr
vn ¼ vD ðtn Þ; on @Xin ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð20bÞ

vn ¼ 0; on @Xwall ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð20cÞ
 
n T 1 n1
n
$v þ ð$v Þ  p I  n ¼ 0; on @Xout : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð20dÞ
Ca
Then, for the second step, we solve the auxiliary equation for an auxiliary unknown un :
 
1 n Re Ca
$  $u ¼ $  vn ; in X; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð21aÞ
qn s
$un  n ¼ 0; on @Xin [ @Xwall ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð21bÞ

un ¼ 0; on @Xout : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð21cÞ

Finally, the pressure pn and velocity un at time level n are updated by

pn ¼ pn1 þ un  2Ca$  vn ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð22aÞ


s 1
u n ¼ vn  $un : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð22bÞ
Re Ca qn

6 2018 SPE Journal

ID: jaganm Time: 17:21 I Path: S:/J###/Vol00000/180034/Comp/APPFile/SA-J###180034


J182607 DOI: 10.2118/182607-PA Date: 8-June-18 Stage: Page: 7 Total Pages: 18

Note that both vn and un are approximations of the velocity v and that vn satisfies the Dirichlet boundary condition while possibly
violating an incompressibility condition, whereas un is solenoidal but satisfies the Dirichlet condition only in the normal direction. The
term yðqn ; vn ; vn1 Þ in Eq. 20a is a linearization of the convection

term $  ðqv vÞ— for instance, the skew-symmetric approximation
given by yðqn ; vn ; vn1 Þ ¼ qn ðvn1  $Þvn þ 12 ð$  vn1 Þvn . The new pressure pn is obtained by correcting the old pressure with the
“pressure update” un and the divergence of the velocity (cf. Eq. 22a).
Spatial Discretization. The SIPG variant of the DG method is used for the spatial discretization of the NS equation. The partition of
the domain into elements and related notation are discussed previously. Then, the fully discrete scheme of the incompressible velocity
equation (Eq. 20) reads as follows: Find vnh 2 Pr ðeh Þ3 such that for any hh 2 Pr ðeh Þ3 ,
s s s
ðqh vnh ; hh Þ þ saC ðqh ; vn1 n
ae ðvnh ; hh Þ ¼ ðqh vn1
h ; vh ; hh Þ þ h ; hh Þ þ bP ðqh ; hh Þ þ bu ðun1
h ; hh Þ
Re Re Ca Re Ca
3 1 s s n
þ s pffiffiffi bjðlh ; Ch ; hh Þ þ 2 ð f n ; hh Þ þ sbnC ðqh ; hh Þ þ b ðhh Þ;                         ð23aÞ
2 2 Re CaCn Fr Re e
ðv0h ; hh Þ ¼ ðv0 ; hh Þ: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð23bÞ

For the elliptic problem (Eq. 21), we define the fully discrete problem by: Find unh 2 Pr ðeh Þ, such that for any wh 2 Pr ðeh Þ,
  ð
1 Re Ca X
adiff n ; unh ; wh ¼  $  vnh wh : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð23cÞ
q s E2eh
E

And for the post-processing step, the discrete pressure pnh and velocity unh are updated by
X ð
n n1 n
ðph ; wh Þ ¼ ðph ; wh Þ þ ðuh ; wh Þ  2Ca E2e
$  vnh wh ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð23dÞ
h
E

s Xð 1
ðunh ; hh Þ ¼ ðvnh ; hh Þ  $unh hh : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð23eÞ
Re Ca E2eh qn
E

In Eq. 23, the trilinear form aC ð; ; Þ and the linear form bC ðÞ stem from the convection term $  ðqv vÞ; the bilinear form ae ð; Þ
and the linear form be ðÞ stem from the divergence-strain term $  ½$v þ ð$vÞT ; the linear form bP ðÞ stems from pressure term –$p;
the linear form bu ðÞ stems from the term –$u; and the trilinear form bI ð; ; Þ stems from the interface term l$c. The linear, bilinear,
and trilinear forms in Eq. 23 can be found in Riviere et al. (2008). The discrete-order parameters ch 2 Pp ðeh Þ and lh 2 Pp ðeh Þ are
obtained from the solution of the CH equation for each timestep using the same order of approximation as NS.

Coupled Solution of the CH-NS System. Eqs. 16 and 23 describe the coupled CH and NS systems without iterating for a given time-
step. To increase accuracy and reduce a splitting error, the CH and NS systems may be solved in an iteratively coupled fashion using a
fixed-point iteration scheme. Here, the CH system is solved first using the velocity field from the preceding timestep. Then, the NS sys-
tem is solved with the order parameter from the CH solution for a new velocity field. This process is repeated until both iterates con-
verge at a given timestep. The Picard-split velocity contributions in aC ð; ; Þ are updated in every iteration; thus, the NS velocity
converges to the fully implicit solution.

Convergence Results
The numerical convergence of the CH module is studied, and the details related to the NS module will be the subject of a separate publi-
cation. Estimated convergence rates are reported in Tables 1 and 2 for the CH module with r ¼ 1 and r ¼ 2, respectively, where p
denotes the polynomial order of the DG approximation. The following manufactured solution is used in this study for the CH module:

cðt; x; y; zÞ ¼ et sinð2pxÞ sinð2pyÞ sinð2pzÞ; t 2 J; ðx; y; zÞ 2 ½0; 13 : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð24Þ

For this study, the velocity is given and is independent of time:


2 p p 3
sin x þ sin y þ cosðzÞ
6 6 3 7
6 7
6 sin x þ p cosðyÞcos z þ p 7 3
vðt; x; y; zÞ ¼ 6 7; t 2 J; ðx; y; zÞ 2 ½0; 1 : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð25Þ
6 3 6 7
4 p p 5
cosðxÞcos y þ sin z þ
6 3

Number of Elements Number of Timesteps L2-Error Convergence Rate


–1
1×1×1 1 3.364×10 –
–2
2×2×2 4 8.350×10 2.010
–2
4×4×4 16 6.055×10 0.464
–2
8×8×8 64 2.506×10 1.273
–3
16×16×16 256 8.028×10 1.642
–3
32×32×32 1,024 2.197×10 1.870

Table 1—Convergence data for the advective CH component of the CH-NS algorithm ðr ¼ 1; r ¼ 2Þ:

2018 SPE Journal 7

ID: jaganm Time: 17:22 I Path: S:/J###/Vol00000/180034/Comp/APPFile/SA-J###180034


J182607 DOI: 10.2118/182607-PA Date: 8-June-18 Stage: Page: 8 Total Pages: 18

Number of Elements Number of Timesteps L2-Error Convergence Rate


–2
1×1×1 1 6.257×10 –
–2
2×2×2 8 5.397×10 0.213
–2
4×4×4 64 3.699×10 0.545
–3
8×8×8 512 5.093×10 2.861

Table 2—Convergence data for the advective CH component of the CH-NS algorithm ðr ¼ 2; r ¼ 4Þ:

Numerical Tests
The simulator that encapsulates the numerical implementation of CH-NS equations is referred to as the pore-scale multiphase-flow sim-
ulator (PMFS). PMFS was validated using an extensive number of fundamental physical-validation tests, some of which are described
next. Subsequently, PMFS is applied to the direct simulation of two-phase flow on two pore-scale images with the purpose of demon-
strating its utility. A linear approximation (r ¼ 1) is used for the numerical solution of the primary solution variables c and v, and the
chemical potential l. Here, the mobility was set equal to one.

Single-Phase Poiseuille Flow. The momentum-conservation component of PMFS is first validated using steady-state single-phase
incompressible flow in a cylindrical geometry. The Poiseuille flow is a simple test case generally used to verify the correct implementa-
tion of viscous forces and no-slip boundaries at the wall (Kunz et al. 2016). For this test, we use only the NS module; the details will be
the subject of a separate publication. The numerical-test problem is set up in the dimensionless space/time domain. Note that the cylin-
drical flow domain (25,984 fluid elements) is embedded into a cubic impermeable-solid domain discretized initially with a 323232
mesh. A Dirichlet velocity boundary condition is used at the inlet, and a Neumann-type boundary condition is used at the outlet faces of
the cylindrical domain. A no-slip boundary condition is enforced on the solid surfaces delineating the boundaries of the cylindrical flow
domain. An accurate velocity profile is obtained for the initial coarse-mesh simulation. The numerically obtained velocity profile
approaches the analytical solution when the mesh (for the entire domain) is refined two-fold in each direction to a 646464 mesh,
which gives rise to 206,592 fluid elements in the cylindrical domain. The maximum error in the numerical velocity profile reduces from

2.5% to
1.9% as the result of mesh refinement (Fig. 2).

1.25
Analytical solution
PMFS solution (32×32×32)
PMFS solution (64×64×64)
1

0.75
Velocity

0.5

0.25

0
0 0.2 0.4 0.6 0.8 1
r

Fig. 2—Analytical and numerical solutions for the dimensionless velocity profile. Numerical solutions are reported for a coarse
mesh and a fine mesh.

Relaxation to Capillary Equilibrium Problems. Two fundamental problems in which systems relax from nonequilibrium initial con-
dition to a capillary equilibrium situation are investigated to qualitatively and quantitatively validate the CH-NS implementation inside
PMFS: capillary bridge and merging droplets. They are closely related to each other, and thus are described in a common section. We
briefly describe these problems and document the results next.
Capillary Bridge. One of the simplest examples that demonstrates the thermodynamically stable coexistence of two liquid phases in
contact with solid walls is the capillary-bridge problem (Myshkis et al. 1987). The capillary-bridge problem not only illustrates the energy-
minimizing attribute of the CH-NS formulation, but also facilitates the validation of the wetting boundary-condition implementation.
An initially rectangular prism-shaped fluid (Phase A) is embedded into a background fluid (Phase B) (Fig. 3) with both fluids touch-
ing solid walls with their top and bottom surfaces. Initially, the system is at rest from the viewpoint of advection (i.e., v ¼ 0Þ: The dy-
namics of the system is fully governed by the thermodynamic nonequilibrium that seeks the minimization of Helmholtz free-energy
density. A wetting-boundary condition is enforced at solid-wall surfaces to honor prescribed equilibrium-contact angles. The prescribed
equilibrium-contact angle is 90 on the top surface and 60 on the bottom surface (computed with respect to Phase A). The mesh size is
202010. The system evolves to reach an equilibrium configuration with a constant dimensionless timestep size of 104: Evolution
of the order-parameter field c and velocity field v for the capillary-bridge problem is shown in the frames of Fig. 3. Eventually, the sys-
tem reaches equilibrium capillary-bridge configuration after approximately 512 timesteps. Upon equilibrium, the enforced contact
angles are honored accurately.

8 2018 SPE Journal

ID: jaganm Time: 17:22 I Path: S:/J###/Vol00000/180034/Comp/APPFile/SA-J###180034


J182607 DOI: 10.2118/182607-PA Date: 8-June-18 Stage: Page: 9 Total Pages: 18

(4 timesteps) (32 timesteps) (512 timesteps)

Fig. 3—Evolution of the order-parameter field for the capillary-bridge problem. Phase A (red) is surrounded by Phase B (transparent/
blue). Top frames show 3D views of the order-parameter field c with superimposed velocity vectors v. Bottom frames show 2D
views of the order-parameter field with superimposed velocity vectors. The first bottom frame shows the 2D horizontal cross sec-
tion across the center of the domain. The second and third bottom frames show 2D vertical cross sections across the domain. The
zoom factors for the velocity vectors are 2X, 5X, and 15X from left to right.

Merging Droplets. In the second scenario, we study the evolution of two merging droplets. The initial and problem-configuration
data were adapted from Ceniceros et al. (2010) for two droplets with an already developed smooth interface:
    
r  jjx0  xjj r  jjx1  xjj
c0 ðxÞ ¼ max 1; tanh pffiffiffi ; tanh pffiffiffi ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð26Þ
2Cn 2Cn
where x0 ¼ ð0:35; 0:35; 0:35ÞT and x1 ¼ ð0:65; 0:65; 0:65ÞT are the centers of the initial droplets with radius r ¼ 1=4. The mesh size is
606060. The initial velocity field equals to zero. Therefore, the dynamics of the system is dominated by thermodynamic nonequili-
brium. The system evolves to reach an equilibrium configuration with a constant dimensionless timestep size of 104: Fig. 4 shows the
numerical approximations of the order-parameter field c and the velocity field v as a function of time. As the system evolves, the two
droplets merge to form one single drop in the stationary case because the most thermodynamically favorable (energy-minimizing) con-
figuration is a single droplet.

Piston-Type Motion in Two Crossing Ducts. Piston-type displacements in duct-like geometries have been described by Lenormand
et al. (1983) and Lenormand and Zarcone (1988) as elementary processes for two-phase flow at the pore-scale. For drainage, where non-
wetting phase displaces wetting phase, the flow can take place only when the pressure at the entrance of a capillary equals or exceeds a
threshold pressure determined by the capillary geometry and IFT between wetting and nonwetting fluids (Lenormand et al. 1983),
whereas in the case of imbibition (wetting fluid displaces nonwetting fluid), the process starts spontaneously after the wetting fluid is
put in contact with the capillary entrance.
To describe the two crossing ducts, we construct a domain consisting of 808080 cubic cells and prescribe an equilibrium contact
angle of 45 for Case 1, 135 for Case 2, and 180 for Case 3 on solid surfaces. In other words, Phase A is the wetting-fluid phase in
Case 1, and Phase B is the wetting-fluid phase in Case 2 and in Case 3.
In the frames of Fig. 5, we present the simulation results corresponding to Case 1. Initially, both ducts of the model are filled with
the nonwetting fluid (Phase B). In the simulation, two out of four ducts were connected to reservoirs containing Phase A at the same ref-
erence pressure as in Phase B. The wetting phase (Phase A) invades the model, causing piston-like displacement of the nonwetting
phase. The contact angle enforced at the line where the fluid/fluid interface touches the solid-wall surface (contact line) is approxi-
mately 45 . It is important to note that an equilibrium contact angle is used to model surface wetting in a dynamic-displacement pro-
cess, assuming that the contact-angle equilibrium is reached instantaneously at the contact line (i.e., the equilibration time scale is faster
than the dynamic timestep size). The investigated phase configuration enables the temporal formation of a single interface inside the
wider area of the crossing ducts (Figs. 5c and 5d). The interface subsequently touches the corner and splits again into two separate inter-
faces (Fig. 5e). This mechanism is consistent with the observations of Lenormand et al. (1983) and Armstrong et al. (2016).
Next, we investigate the case with nonwetting fluid displacing the wetting fluid (Case 2 and Case 3). The model is initially filled with
the wetting fluid (Phase B). The nonwetting fluid (Phase A) was injected through two ducts, while the other two were connected to the
reservoirs containing Phase B at reference pressure. For Case 2, the dynamics of the displacement process are presented in the frames of
Fig. 6. The contact angle enforced at the contact line is approximately 135 . As in Case 1, the investigated phase configuration enables
the temporal formation of a single interface inside the wider area of the crossing ducts (Figs. 6c and 6d). Subsequently, the interface
touches the corner and splits into two separate interfaces. For Case 3, the dynamics of the displacement process is presented in the frames
of Fig. 7. They resemble Case 2, except that we now observe the residual phase phenomenon in this simulation (Figs. 7e and 7f).

2018 SPE Journal 9

ID: jaganm Time: 17:22 I Path: S:/J###/Vol00000/180034/Comp/APPFile/SA-J###180034


J182607 DOI: 10.2118/182607-PA Date: 8-June-18 Stage: Page: 10 Total Pages: 18

Snapshot 1 Snapshot 2 Snapshot 3

Snapshot 4 Snapshot 5 Snapshot 6

Fig. 4—3D views of the evolution of the order-parameter field for the merging-droplets problem. Interface (c 5 0) is shown in green,
while phases are transparent. The velocity is visualized as black glyphs.

(a) (b) (c)

(d) (e) (f)

Fig. 5—Imbibition process in crossing ducts: Wetting phase (Phase A) displaces nonwetting phase (Phase B). Interface is shown
in green, wetting phase (Phase A) is red, and nonwetting phase (Phase B) is blue. The equilibrium contact angle imposed on solid
surfaces is 458.

10 2018 SPE Journal

ID: jaganm Time: 17:22 I Path: S:/J###/Vol00000/180034/Comp/APPFile/SA-J###180034


J182607 DOI: 10.2118/182607-PA Date: 8-June-18 Stage: Page: 11 Total Pages: 18

(a) (b) (c)

(d) (e) (f)

Fig. 6—Drainage process in crossing ducts: Nonwetting phase (Phase A) displaces wetting phase (Phase B). Interface is shown in
green, nonwetting phase (Phase A) is red, and wetting phase (Phase B) is blue. The equilibrium contact angle imposed on solid
surfaces is 1358.

(a) (b) (c)

(d) (e) (f)

Fig. 7—Drainage process in crossing ducts: Nonwetting phase (Phase A) displaces wetting phase (Phase B). Interface is shown in
green, nonwetting phase (Phase A) is red, and wetting phase (Phase B) is blue. The equilibrium contact angle imposed on solid
surfaces is 1808.

Flow Simulation on lCT Images. This set of numerical tests features immiscible two-phase two-component flow through post-
segmentation pore-image domains that are generated by means of the mCT imaging of a sphere pack and a Berea-sandstone sample
(Ändra et al. 2013a, b). A classical approach for approximating a porous medium is to use spherical grains packed randomly. Depend-
ing on the material constituting the solid spheres, sphere packs lend themselves to imaging more naturally than complex rocks. Thus,
one of our application examples features a sphere-pack image. However, Berea sandstone is one of the benchmark rock types typically
used for pore-scale modeling. It is important to note that this section focuses on the application of PMFS rather than validating a certain
physical attribute of it. The objective here is to demonstrate the applicability of PMFS to complex pore-scale images of realistic/real
porous materials.
The sample-domain size used for the sphere pack is 100100100 with 232,308 active mesh blocks in the image volume. The sam-
ple-domain size used for Berea sandstone is 150150150 with 798,190 active mesh blocks in the image volume. To induce a flow
field v into the pore space, we attach inflow and outflow buffers, each with a size of 10YY and 15YY voxels at the opposing-sides
surfaces of the porous medium aligned with the main flow direction (Y ¼ 100 for the sphere pack and Y ¼ 150 for the Berea sandstone).
The buffers allow the flow field to develop in a more physically consistent fashion and, thereby, permit the imposition of physically re-
alistic boundary conditions at the inlet and outlet.

2018 SPE Journal 11

ID: jaganm Time: 17:23 I Path: S:/J###/Vol00000/180034/Comp/APPFile/SA-J###180034


J182607 DOI: 10.2118/182607-PA Date: 8-June-18 Stage: Page: 12 Total Pages: 18

At initial time, the domain is saturated by a single phase (i.e., c0 ¼ 1). Then, the second component is injected through the inlet
buffer. The sphere-pack pore-scale image volume and a snapshot of the velocity field are shown in Fig. 8. The evolution of the order-
parameter field is documented in the frames of Fig. 9 for the sphere-pack model. Fig. 10 shows a snapshot of the order-parameter and
velocity fields for the Berea-sandstone model. The evolution of the order-parameter field is documented in the frames of Fig. 11 for the
Berea-sandstone model. Illustrations in Figs. 8 through 11 provide first-order evidence that PMFS can handle complicated geometries
from segmented 3D images without giving rise to distorted/nonconvergent flow fields.

Pore-scale image volume Velocity field

Fig. 8—Sphere-pack pore-scale image volume (left-hand side) and snapshot of the computed velocity field (right-hand side).

Snapshot 1 Snapshot 2 Snapshot 3

Snapshot 4 Snapshot 5 Snapshot 6

Fig. 9—Evolution of the order-parameter field for the sphere-pack model. Phase A (c > 0) is visualized red, Phase B (c < 0) is blue,
and the center of the diffuse interface (c 5 0) is green.

12 2018 SPE Journal

ID: jaganm Time: 17:23 I Path: S:/J###/Vol00000/180034/Comp/APPFile/SA-J###180034


J182607 DOI: 10.2118/182607-PA Date: 8-June-18 Stage: Page: 13 Total Pages: 18

Pore-scale image volume Velocity field

Fig. 10—Berea-sandstone image volume (left-hand side) and a snapshot of the computed velocity field (right-hand side). The
increase in velocity through the narrow throat in the vicinity (right frame) is clearly visible.

Snapshot 1 Snapshot 2 Snapshot 3

Snapshot 4 Snapshot 5 Snapshot 6

Fig. 11—Evolution of the order-parameter field for the Berea-sandstone model. Phase A (c > 0) is visualized red, Phase B (c < 0) is
blue, and the center of the diffuse interface (c 5 0) is green.

PMFS captures the trapping of the displaced phase during a viscocapillary displacement process, as shown in Fig. 12. Having stated
that, the simulated image is too small to provide a physically reasonable value for the trapped-phase saturation. Because the physical
model underlying PMFS is based on a full viscocapillary coupling, there is no theoretical blocker for PMFS not to capture trapping at
imbibition. However, one of the challenges for achieving the objective of simulating a complete drainage and imbibition cycle is devel-
oping a method for the accurate incorporation of the wetting-boundary condition on models operating on mCT-image data, which we
recently addressed in Frank et al. (2018b). A complete drainage and imbibition cycle simulation will be the subject of a future publica-
tion after integration of this method into PMFS.

2018 SPE Journal 13

ID: jaganm Time: 17:24 I Path: S:/J###/Vol00000/180034/Comp/APPFile/SA-J###180034


J182607 DOI: 10.2118/182607-PA Date: 8-June-18 Stage: Page: 14 Total Pages: 18

1
Component A
Component B

0.8

0.6
Saturation

0.4

0.2

0
0 0.4 0.8 1.2 1.6 2
Time

Fig. 12—Evolution of the average saturation of Phase A and Phase B in a 100 3 100 3 100 subset of the Berea-sandstone image data.

Another major challenge is being able to simulate larger models (e.g., sizes 1,0003 to 2,0003 voxels), extending the capability to cm-
scale volumes for computing a reasonable estimate of residual saturation and other effective properties. All simulations in this paper
were run shared-memory parallel. Run times vary between a few hours (for small cases) and a few days (e.g., the sphere pack and
Berea-sandstone cases) for the numerical examples shown in this paper. Significant acceleration is necessary to reach more-realistic
problem sizes and to achieve multirealization simulation studies. Therefore, current work is focusing on distributed parallel implemen-
tation of PMFS within the TRILINOS framework (Heroux et al. 2005) to handle significantly larger pore images. Another challenge
being worked on is the simulation of low-capillary-number (more capillary forces dominated) flows. As shown in Eq. 19a, as the capil-
lary number becomes smaller, the pressure-gradient term in the right-hand side of the equation approaches an increasingly larger num-
ber. In the zero-capillary-number limit, this term approaches infinity. Currently, the lowest-capillary-number flows simulated with
PMFS are on the order of 1.0103. Activities are ongoing to reduce this limit to approximately 1.0106 through specially designed
solver techniques that handle singularities better. Also, we are focusing on further validation of PMFS featuring detailed investigations
of snap-off phenomena and Haines jumps (e.g., Armstrong and Berg 2013; Armstrong et al. 2015, 2016).

Conclusions
A Helmholtz free-energy-driven, thermodynamically based diffuse-interface method is used for the effective simulation of numerous
advecting interfaces, while honoring the IFT. The advective CH (mass balance, energy dissipation) and NS (momentum balance) equa-
tions are coupled to each other within the phase-field framework. Wettability on rock/fluid interfaces is accounted for by means of an
energy-penalty-based wetting-boundary condition. Individual balance equations are discretized with a flexible DG method. The discreti-
zation of either equation is semi-implicit in time, where, for the mass-balance equation, a convex/concave splitting and, for the momen-
tum-balance equation, a linear-projection method are used.
In this paper, we discuss the mathematical model and its DG discretization, and briefly describe the nonlinear- and linear-solution
strategies. Numerical-validation tests show optimal convergence rates for the DG discretization, signifying the correctness of the nu-
merical scheme. Physical-validation tests demonstrate the consistency of the order-parameter and velocity fields simulated with PMFS.
Finally, two-phase-flow simulations on real pore-scale images are shown to demonstrate the usefulness of the pore-scale simulator.
The proposed method is accurate, numerically robust, and exhibits the potential for simulating realistic problems. It models rigorously
the flow physics by directly acting on complex pore-scale images. Our investigations show that PMFS has the potential to serve as an accu-
rate tool for the investigation of the dynamics of pore-scale displacement phenomena. More importantly, the demonstrated accuracy of the
DG discretization within PMFS renders it an ideal benchmarking tool for more-approximate finite-volume and LBM-based simulators.
PMFS is also designed to provide fundamental physical insights and conceptual understanding about multiphase flow at the pore scale and,
thereby, to enable upscaling through more-approximate techniques to larger spatially averaged flow-simulation scales (e.g., Darcy scale).
Future studies will focus on understanding the physical operational limits of PMFS, such as capillary number and mobility ratio.
The Helmholtz free-energy minimization-based nature of the phase-field model lends itself naturally to energy-based wetting-
boundary conditions that can be imposed on pore-scale images that directly stem from mCT imaging (Frank et al. 2018b), with the ulti-
mate objective of modeling mixed wettability more accurately.
Although there are three-phase immiscible extensions of the CH equation available in the literature, a more interesting direction is
the handling of miscibility and multiphase multicomponent flow at the pore scale. The thermodynamics-based nature of the phase-field
model naturally allows the incorporation of an equation-of-state-based phase-behavior framework to the simulation environment. Such
a multicomponent extension of the CH equation will permit the simulation of miscible multicomponent flow at the pore scale to develop
a conceptual understanding of miscible enhanced-oil-recovery schemes.

Acknowledgment
The authors would like to thank Shell International Exploration and Production for permission to publish this paper.

References
Abels, H., Garcke, H., Grün, G. et al. 2017. Diffuse Interface Models for Incompressible Two-Phase Flows With Different Densities. In Transport Proc-
esses at Fluidic Interfaces, D. Bothe and A. Reusken (eds), 203–229. Advances in Mathematical Fluid Mechanics Series: Birkhäuser, Cham. https://
doi.org/10.1007/978-3-319-56602-3_8.

14 2018 SPE Journal

ID: jaganm Time: 17:24 I Path: S:/J###/Vol00000/180034/Comp/APPFile/SA-J###180034


J182607 DOI: 10.2118/182607-PA Date: 8-June-18 Stage: Page: 15 Total Pages: 18

Abu-Al-Saud, M. O., Riaz A., and Tchelepi, H. A. 2017. Multiscale Level-Set Method for Accurate Modeling of Immiscible Two-Phase Flow With De-
posited Thin Films on Solid Surfaces. Journal of Computational Physics 333 (Supplement C): 297–320. https://doi.org/10.1016/j.jcp.2016.12.038.
Adalsteinsson, D. and Sethian, J. A. 1999. The Fast Construction of Extension Velocities in Level Set Methods. Journal of Computational Physics 148
(1): 2–22. https://doi.org/10.1006/jcph.1998.6090.
Alpak, F. O., Riviere, B., and Frank, F. 2016. A Phase-Field Method for the Direct Simulation of Two-Phase Flows in Pore-Scale Media Using a Non-
Equilibrium Wetting Boundary Condition. Computational Geosciences 20 (5): 881–908. https://doi.org/10.1007/s10596-015-9551-2.
Alpak, F. O., Samardžić, A., and Frank, F. 2017. A Distributed Parallel Direct Simulator for Pore-Scale Two-Phase Flow on Digital Rock Images Using
a Finite-Volume-Based Implementation of the Phase-Field Method. Journal of Petroleum Science and Engineering. In Press. https://doi.org/10.1016/
j.petrol.2017.11.022.
Anderson, D. M., McFadden, G. B., and Wheeler, A. A. 1998. Diffuse-Interface Methods in Fluid Mechanics. Annual Review of Fluid Mechanics 30:
139–165. https://doi.org/10.1146/annurev.fluid.30.1.139.
Andrä, H., Combaret, N., Dvorkin, J. et al. 2013a. Digital Rock Physics Benchmarks—Part I: Imaging and Segmentation. Computers & Geosciences 50:
25–32. https://doi.org/10.1016/j.cageo.2012.09.005.
Andrä, H., Combaret, N., Dvorkin, J. et al. 2013b. Digital Rock Physics Benchmarks—Part II: Computing Effective Properties. Computers & Geoscien-
ces 50: 33–43. https://doi.org/10.1016/j.cageo.2012.09.008.
Andrew, M., Menke, H., and Blunt, M. J. 2015. The Imaging of Dynamic Multiphase Fluid Flow Using Synchrotron-Based X-Ray Microtomography at
Reservoir Conditions. Transport in Porous Media 110 (1): 1–24. https://doi.org/10.1007/s11242-015-0553-2.
Antanovskii, L. K. 1995. A Phase Field Model of Capillarity. Physics of Fluids 7: 747–753. https://doi.org/10.1063/1.868598.
Aristotelous, A. C., Karakashian, O. A., and Wise, S. M. 2013. A Mixed Discontinuous Galerkin, Convex Splitting Scheme for a Modified Cahn-Hilliard
Equation and an Efficient Nonlinear Multigrid Solver. Discrete and Continuous Dynamical Systems - B 18 (9): 2211–2238. https://doi.org/10.3934/
dcdsb.2013.18.2211.
Armstrong, R. T. and Berg, S. 2013. Interfacial Velocities and Capillary Pressure Gradients During Haines Jumps. Physical Review E 88 (4): 043010-1-
9. https://doi.org/10.1103/PhysRevE.88.043010.
Armstrong, R. T., Evseev, N., Koroteev, D. et al. 2015. Modeling the Velocity Field During Haines Jumps in Porous Media. Advances in Water Resour-
ces 77: 57–68. https://doi.org/10.1016/j.advwatres.2015.01.008.
Armstrong, R. T., Berg, S., Dinariev, O. et al. 2016. Modeling of Pore-Scale Two-Phase Phenomena Using Density Functional Hydrodynamics. Trans-
port in Porous Media 112 (3): 577–607. https://doi.org/10.1007/s11242-016-0660-8.
Badalassi, V. E., Ceniceros, H. D., and Banerjee, S. 2003. Computation of Multiphase Systems With Phase Field Models. Journal of Computational
Physics 190 (2): 371–397. https://doi.org/10.1016/S0021-9991(03)00280-8.
Bao, K., Shi, Y., Sun, S. et al. 2012. A Finite Element Method for the Numerical Solution of the Coupled Cahn–Hilliard and Navier-Stokes System for
Moving Contact Line Problems. Journal of Computational Physics 231 (24): 8083–8099. https://doi.org/10.1016/j.jcp.2012.07.027.
Berg, S., Ott, H., Klapp, S. A. et al. 2013. Real-Time 3D Imaging of Haines Jumps in Porous Media Flow. Proc. of the National Academy of Sciences
110 (10): 3755–3759. https://doi.org/10.1073/pnas.1221373110.
Berg, S., Armstrong, R., Ott, H. et al. 2014. Multiphase Flow in Porous Rock Imaged Under Flow Conditions With Fast X-Ray Computed Microtomog-
raphy. Petrophysics 55 (4): 304–312. SPWLA-2014-v55n4a3.
Berg, S., Rücker, M., Ott, H. et al. 2016. Connected Pathway Relative Permeability From Pore-Scale Imaging of Imbibition. Advances in Water Resour-
ces 90: 24–35. https://doi.org/10.1016/j.advwatres.2016.01.010.
Blunt, M. J. 2001. Flow in Porous Media—Pore-Network Models and Multiphase Flow. Current Opinion in Colloid & Interface Science 6 (3): 197–207.
https://doi.org/10.1016/S1359-0294(01)00084-X.
Blunt, M. J., Jackson, M. D., Piri, M. et al. 2002. Detailed Physics, Predictive Capabilities and Macroscopic Consequences for Pore-Network Models of
Multiphase Flow. Advances in Water Resources 25 (8–12): 1069–1089. https://doi.org/10.1016/S0309-1708(02)00049-0.
Blunt, M. J., Bijeljic, B., Dong, H. et al. 2013. Pore-Scale Imaging and Modelling. Advances in Water Resources 51: 197–216. https://doi.org/10.1016/
j.advwatres.2012.03.003.
Boek, E. S., Zacharoudiou, I., Gray, F. et al. 2017. Multiphase-Flow and Reactive-Transport Validation Studies at the Pore Scale by Use of Lattice Boltz-
mann Computer Simulations. SPE J. 22 (3): 940–949. SPE-170941-PA. https://doi.org/10.2118/170941-PA.
Bogdanov, I. I., Kpahou, J., and Guerton, F. 2012. Pore-Scale Single and Two-Phase Transport in Real Porous Medium. Presented at the ECMOR XIII–13th
European Conference on the Mathematics of Oil Recovery, Biarritz, France, 10–13 September. https://doi.org/10.3997/2214-4609.20143250.
Boyer, F., Lapuerta, C., Minjeaud, S. et al. 2010. Cahn-Hilliard/Navier-Stokes Model for the Simulation of Three-Phase Flows. Transport in Porous
Media 82 (3): 463–483. https://doi.org/10.1007/s11242-009-9408-z.
Bultreys, T., De Boever, W., and Cnudde, V. 2016. Imaging and Image-Based Fluid Transport Modeling at the Pore Scale in Geological Materials:
A Practical Introduction to the Current State-of-the-Art. Earth-Science Reviews 155: 93–128. https://doi.org/10.1016/j.earscirev.2016.02.001.
Cahn, J. W. and Hilliard, J. E. 1958. Free Energy of a Nonuniform System. I. Interfacial Free Energy. Journal of Chemical Physics 28: 258–267. https://
doi.org/10.1063/1.1744102.
Cahn, J. W. 1959. Free Energy of a Nonuniform System. II. Thermodynamic Basis. Journal of Chemical Physics 30: 1121–1124. https://doi.org/
10.1063/1.1730145.
Cahn, J. W. and Hilliard, J. E. 1959. Free Energy of a Nonuniform System. III. Nucleation in a Two-Component Incompressible Fluid. Journal of Chem-
ical Physics 31: 688–699. https://doi.org/10.1063/1.1730447.
Cahn, J. W. 1961. On Spinodal Decomposition. Acta Metallurgica 9 (9): 795–801. https://doi.org/10.1016/0001-6160(61)90182-1.
Cahn, J. W. and Hilliard, J. E. 1971. Spinodal Decomposition: A Reprise. Acta Metallurgica 19 (2): 151–161. https://doi.org/10.1016/0001-
6160(71)90127-1.
Carlson, A., Do-Quang, M., and Amberg, G. 2011. Dissipation in Rapid Dynamic Wetting. Journal of Fluid Mechanics 682: 213–240. https://doi.org/
10.1017/jfm.2011.211.
Carlson, A. 2012. Capillarity and Dynamic Wetting. Technical Report. Stockholm, Sweden: Royal Institute of Technology.
Ceniceros, H. D., Nos, R. L., and Roma, A. M. 2010. Three-Dimensional, Fully Adaptive Simulations of Phase-Field Fluid Models. Journal of Computa-
tional Physics 229 (17): 6135–6155. https://doi.org/10.1016/j.jcp.2010.04.045.
Chatenever, A. and Calhoun, J. C. 1952. Visual Examinations of Fluid Behavior in Porous Media—Part I. J Pet Technol 4 (6): 149–156. SPE-135-G.
https://doi.org/10.2118/135-G.
Chen, S. and Doolen, G. D. 1998. Lattice Boltzmann Method for Fluid Flows. Annual Review of Fluid Mechanics 30: 329–364. https://doi.org/10.1146/
annurev.fluid.30.1.329.
Chen, L., Kang, Q., Robinson, B. A. et al. 2013. Pore-Scale Modeling of Multiphase Reactive Transport With Phase Transitions and Dissolution-Precipi-
tation Processes in Closed Systems. Physical Review E 87: 043306-1-16. https://doi.org/10.1103/PhysRevE.87.043306.

2018 SPE Journal 15

ID: jaganm Time: 17:25 I Path: S:/J###/Vol00000/180034/Comp/APPFile/SA-J###180034


J182607 DOI: 10.2118/182607-PA Date: 8-June-18 Stage: Page: 16 Total Pages: 18

Chen, Y. and Shen, J. 2016. Efficient, Adaptive Energy Stable Schemes for the Incompressible Cahn–Hilliard Navier–Stokes Phase-Field Models. Jour-
nal of Computational Physics 308 (issue C): 40–56. https://doi.org/10.1016/j.jcp.2015.12.006.
Cnudde, V. and Boone, M. N. 2013. High-Resolution X-Ray Computed Tomography in Geosciences: A Review of the Current Technology and Applica-
tions. Earth–Science Reviews 123: 1–17. https://doi.org/10.1016/j.earscirev.2013.04.003.
Darwish, M. and Moukalled, F. 2006. Convective Schemes for Capturing Interfaces of Free-Surface Flows on Unstructured Grids. Numerical Heat
Transfer, Part B: Fundamentals 49: 19–42.
Ding, H. and Spelt, P. D. M. 2007. Wetting Condition in Diffuse Interface Simulations of Contact Line Motion. Physical Review E 75 (4): 046708-1–8.
https://doi.org/10.1103/PhysRevE.75.046708.
Ding, H., Spelt, P. D. M., and Shu, C. 2007. Diffuse Interface Model for Incompressible Two-Phase Flows With Large Density Ratios. Journal of Com-
putational Physics 226 (2): 2078–2095. https://doi.org/10.1016/j.jcp.2007.06.028.
Elliott, C. M. and Stuart, A. M. 1993. The Global Dynamics of Discrete Semilinear Parabolic Equations. SIAM Journal on Numerical Analysis 30 (6):
1622–1663. https://doi.org/10.1137/0730084.
Enright, D., Fedkiw, R., Ferziger, J. et al. 2002. A Hybrid Particle Level Set Method for Improved Interface Capturing. Journal of Computational Physics
183 (1): 83–116. https://doi.org/10.1006/jcph.2002.7166.
Eyre, D. J. 1998. Unconditionally Gradient Stable Time Marching the Cahn–Hilliard Equation. In Symposia BB–Computational & Mathematical Models
of Microstructural Evolution, Vol. 529 of MRS Proc. https://doi.org/10.1557/PROC-529-39.
Fatt, I. 1956a. The Network Model of Porous Media I. Capillary Pressure Characteristics. Petroleum Trans., AIME 207: 144–159. SPE-574-G. https://
doi.org/10.2118/574-G.
Fatt, I. 1956b. The Network Model of Porous Media II. Dynamic Properties of a Single-Size Tube Network. Petroleum Trans., AIME 207: 160–163.
SPE-574-G. https://doi.org/10.2118/574-G.
Fatt, I. 1956c. The Network Model of Porous Media III. Dynamic Properties of Networks With Tube Radius Distribution. Petroleum Trans., AIME 207:
164–181. SPE-574-G. https://doi.org/10.2118/574-G.
Feng, X and Karakashian, O. A. 2007. Fully Discrete Dynamic Mesh Discontinuous Galerkin Methods for the Cahn-Hilliard Equation of Phase Transi-
tion. Mathematics of Computation 76 (259): 1093–1117. https://doi.org/10.1090/S0025-5718-07-01985-0.
Ferrari, A. and Lunati, I. 2014. Inertial Effects During Irreversible Meniscus Reconfiguration in Angular Pores. Advances in Water Resources 74: 1–13.
https://doi.org/10.1016/j.advwatres.2014.07.009.
Flannery, B. P., Deckman, H. W., Roberge, W. G. et al. 1987. Three-Dimensional X-ray Microtomography. Science 237 (4821): 1439–1444. https://
doi.org/10.1126/science.237.4821.1439.
Frank, F., Liu, C., Alpak, F. O. et al. 2018a. A Finite Volume/Discontinuous Galerkin Method for the Advective Cahn-Hilliard Equation With Degener-
ate Mobility on Porous Domains Stemming From Micro-CT Imaging. Computational Geosciences. 22 (2): 543–563. https://doi.org/10.1007/s10596-
017-9709-1.
Frank, F., Liu, C., Scanziani, A. et al. 2018b. An Energy-Based Equilibrium Contact Angle Boundary Condition on Jagged Surfaces for Phase-Field
Methods. Journal of Colloid and Interface Science 523: 282–291. https://doi.org/10.1016/j.jcis.2018.02.075.
Georgiadis, A., Berg, S., Makurat, A. et al. 2013. Pore-Scale Micro-Computed-Tomography Imaging: Nonwetting-Phase Cluster-Size Distribution Dur-
ing Drainage and Imbibition. Physical Review E 88: 033002. https://doi.org/10.1103/PhysRevE.88.033002.
Girault, V., Riviere, B., and Wheeler, M. F. 2005. A Discontinuous Galerkin Method With Nonoverlapping Domain Decomposition for the Stokes and
Navier-Stokes Problems. Mathematics of Computation 74 (2005): 53–84. https://doi.org/10.1090/S0025-5718-04-01652-7.
Glimm, J., Grove, J. W., Li, X.-L et al. 1999. Simple Front Tracking. In Contemporary Mathematics, ed. G.-Q. Chen and E. DiBenedetto, Vol. 238, pp.
133–149. Providence, Rhode Island, USA: American Mathematical Society.
Guermond, J. L., Minev, P., and Shen, J. 2006. An Overview of Projection Methods for Incompressible Flows. Computer Methods in Applied Mechanics
and Engineering 195 (44–47): 6011–6045. https://doi.org/10.1016/j.cma.2005.10.010.
Heroux, M. A., Bartlett, R. A., Howle, V. E. et al. 2005. An Overview of the Trilinos Project. ACM Trans. on Mathematical Software 31 (3): 397–423.
https://doi.org/10.1145/1089014.1089021.
Hirt, C. W. and Nichols, B. D. 1981. Volume of Fluid (VOF) Method for the Dynamics of Free Boundaries. Journal of Computational Physics 39 (1):
201–225. https://doi.org/10.1016/0021-9991(81)90145-5.
Hoogerbrugge, P. J. and Koelman, K. M. V. A. 1992. Simulating Microscopic Hydrodynamic Phenomena With Dissipative Particle Dynamics, Europhy-
sics Letters 19 (3): 155–160.
Jacqmin, D. 1996. An Energy Approach to the Continuum Surface Tension Method. AIAA96-0858. In Proc. of the 34th Aerospace Sciences Meeting
and Exhibit, American Institute of Aeronautics and Astronautics. Reno, Nevada, USA.
Jacqmin, D. 1999. Calculation of Two-Phase Navier-Stokes Flows Using Phase-Field Modeling. Journal Computational Physics 155 (1): 96–127.
https://doi.org/10.1006/jcph.1999.6332.
Jacqmin, D. 2000. Contact Line Dynamics of a Diffuse Fluid Interface. Journal of Fluid Mechanics 402: 5788. https://doi.org/10.1017/
S0022112099006874.
Joekar-Niasar, V., Hassanizadeh, S. M., and Dahle, H. 2010. Dynamic Pore-Network Modeling of Drainage in Two-Phase Flow. Journal of Fluid
Mechanics 655: 38–71.
Joekar-Niasar, V. and Hassanizadeh, S. M. 2011. Effect of Fluids Properties on Non-Equilibrium Capillarity Effects: Dynamic Pore-Network Modeling.
International Journal of Multiphase Flow 37 (2): 198–214. https://doi.org/10.1016/j.ijmultiphaseflow.2010.09.007.
Karadimitriou, N. K. and Hassanizadeh, S. M. 2012. A Review of Micromodels and Their Use in Two-Phase Flow Studies. Vadose Zone Journal 11 (3).
https://doi.org/10.2136/vzj2011.0072.
Kay, D., Styles, V., and Suli, E. 2009. Discontinuous Galerkin Finite Element Approximation of the Cahn–Hilliard Equation With Convection. SIAM
Journal on Numerical Analysis 47 (4): 2660–2685. https://doi.org/10.1137/080726768.
Kim, J. 2012. Phase-Field Models for Multi-Component Fluid Flows. Communications in Computational Physics 12 (3): 613–661. https://doi.org/
10.4208/cicp.301110.040811a.
Koroteev, D., Dinariev, O., Evseev, N. et al. 2014. Direct Hydrodynamic Simulation of Multiphase Flow in Porous Rock. Petrophysics 55 (4): 294–303.
SPWLA-2014-v55n4a2.
Kunz, P., Zarikos, I. M., Karadimitriou, N. K. et al. 2016. Study of Multi-Phase Flow in Porous Media: Comparison of SPH Simulations With Micro-
Model Experiments. Transport in Porous Media 114 (2): 581–600.
Lenormand, R., Zarcone, C., and Sarr, A. 1983. Mechanisms of the Displacement of One Fluid by Another in a Network of Capillary Ducts. Journal of
Fluid Mechanics 135: 337–353. https://doi.org/10.1017/S0022112083003110.
Lenormand, R. and Zarcone, C. 1988. Physics of Blob Displacement in a Two-Dimensional Porous Medium. SPE Form Eval 3: 271–275. SPE-14882-
PA. https://doi.org/10.2118/14882-PA.

16 2018 SPE Journal

ID: jaganm Time: 17:25 I Path: S:/J###/Vol00000/180034/Comp/APPFile/SA-J###180034


J182607 DOI: 10.2118/182607-PA Date: 8-June-18 Stage: Page: 17 Total Pages: 18

Li, X., Wu, S., Song, J. et al. 2011. Numerical Simulation of Pore-Scale Flow in Chemical Flooding Process. Theoretical and Applied Mechanics Letters
1 (2): 022008. https://doi.org/10.1063/2.1102208.
Li, J. and Sultan, A. S. 2015. Permeability Computations of Shale Gas by the Pore-Scale Monte Carlo Molecular Simulations. Presented at the Interna-
tional Petroleum Technology Conference, Doha, Qatar, 6–9 December. IPTC-18263-MS. https://doi.org/10.2523/IPTC-18263-MS.
Lowengrub, J. and Truskinovsky, L. 1998. Quasi-Incompressible Cahn–Hilliard Fluids and Topological Transitions. In Proc. of the Royal Society of
London A: Mathematical. Physical and Engineering Sciences 454: 2617–2654.
Luo, H., Segawa, H., and Visbal, M. R. 2012. An Implicit Discontinuous Galerkin Method for the Unsteady Compressible Navier-Stokes Equations.
Computers & Fluids 53: 133–144. https://doi.org/10.1016/j.compfluid.2011.10.009.
Meakin, P. and Tartakovsky, A. M. 2009. Modeling and Simulation of Pore-Scale Multiphase Fluid Flow and Reactive Transport in Fractured and Porous
Media. Reviews of Geophysics 47 (3): 1–47. https://doi.org/10.1029/2008RG000263.
Montlaur, A., Fernandez-Mendez, S., Peraire, J. et al. 2009. Discontinuous Galerkin Methods for the Navier-Stokes Equations Using Solenoidal Approx-
imations. International Journal for Numerical Methods in Fluids 64 (5). https://doi.org/10.1002/fld.2161.
Moseler, M. and Landman, U. 2000. Formation, Stability and Breakup of Nanojets. Science 289 (5482): 1165–1169. https://doi.org/10.1126/
science.289.5482.1165.
Myshkis, A. D., Babskii, V. G., Kopachevskii, N. D. et al. 1987. Low-Gravity Fluid Mechanics: Mathematical Theory of Capillary Phenomena. Berlin,
Germany: Springer-Verlag.
Nie, X. B., Chen, S. Y., Robbins, M. O. et al. 2004. A Continuum and Molecular Dynamics Hybrid Method for Micro- and Nano-Fluid Flow. Journal of
Fluid Mechanics 500: 55–64. https://doi.org/10.1017/S0022112003007225.
Nourgaliev, R. R. and Theofanous, T. G. 2007. High Fidelity Interface Tracking in Compressible Flows: Unlimited Anchored Adaptive Level Set. Jour-
nal of Computational Physics 224: 836–866.
Osher, S. and Fedkiw, S. 2003. Level Set Methods and Dynamic Implicit Surfaces. New York: Springer.
Pagonabarraga, I. and Frenkel, D. 2001. Dissipative Particle Dynamics for Interacting Systems. Journal of Chemical Physics 115 (11): 5015–5026.
https://doi.org/10.1063/1.1396848.
Persson, P.-O., Bonet, J., and Peraire, J. 2009. Discontinuous Galerkin Solution of the Navier–Stokes Equations on Deformable Domains. Computer
Methods in and Engineering 198: 1585–1595. https://doi.org/10.1016/j.cma.2009.01.012.
Pigeonneau, F. and Saramito, P. 2016. Discontinuous Galerkin Finite Element Method Applied to the Coupled Navier-Stokes/Cahn-Hilliard Equations.
Presented at the 9th International Conference on Multiphase Flow (ICMF), Firenze, Italy, 22–27 May.
Porter, M. L., Schaap, M. G., and Wildenschild, D. 2009. Lattice-Boltzmann Simulations of the Capillary Pressure–Saturation–Interfacial Area Relation-
ship for Porous Media. Advances in Water Resources 32 (11): 1632–1640. https://doi.org/10.1016/j.advwatres.2009.08.009.
Prodanovic, M. and Bryant, S. L. 2006. A Level Set Method for Determining Critical Curvatures for Drainage and Imbibition. Journal of Colloid and
Interface Science 304 (2): 442–458. https://doi.org/10.1016/j.jcis.2006.08.048.
Prodanovic, M., Bryant S. L., and Karpyn, Z. T. 2010. Investigating Matrix/Fracture Transfer via a Level Set Method for Drainage and Imbibition. SPE
J. 15 (1): 125–136. SPE-116110-PA. https://doi.org/10.2118/116110-PA.
Purcell, W. R. 1949. Capillary Pressures—Their Measurement Using Mercury and the Calculation of Permeability Therefrom. J Pet Technol 1 (2):
39–48. SPE-949039-G. https://doi.org/10.2118/949039-G.
Raeini, A. Q., Bijeljic, B., and Blunt, M. J. 2012. Modelling Two-Phase Flow in Porous Media at the Pore Scale Using the Volume-of-Fluid Method.
Journal of Computational Physics 231 (17): 5653–5668. https://doi.org/10.1016/j.jcp.2012.04.011.
Raeini, A. Q., Bijeljic, B., and Blunt, M. J. 2014a. Numerical Modelling of Sub-Pore Scale Events in Two-Phase Flow Through Porous Media. Transport
in Porous Media 101 (2): 191–213. https://doi.org/10.1007/s11242-013-0239-6.
Raeini, A. Q., Blunt, M. J., and Bijeljic, B. 2014b. Direct Simulations of Two-Phase Flow on Micro-CT Images of Porous Media and Upscaling of Pore-
Scale Forces. Advances in Water Resources 74: 116–126. https://doi.org/10.1016/j.advwatres.2014.08.012.
Rider, W. J. and Koth, D. B. 1998. Reconstructing Volume Tracking. Journal of Computational Physics 141 (2): 112–152. https://doi.org/10.1006/
jcph.1998.5906.
Riviere, B. 2008. Discontinuous Galerkin Methods for Solving Elliptic and Parabolic Equations: Theory and Implementation (Frontiers in Applied
Mathematics). Philadelphia, Pennsylvania, USA: Society for Industrial and Applied Mathematics. https://doi.org/10.1137/1.9780898717440.
Ryazanov, A. V., van Dijke, M. I. J., and Sorbie, K. S. 2009. Two-Phase Pore-Network Modelling: Existence of Oil Layers During Water Invasion.
Transport in Porous Media 80 (1): 79–99. https://doi.org/10.1007/s11242-009-9345-x.
Sedghi, M., Piri, M., and Goual, L. 2014. Molecular Dynamics of Wetting Layer Formation and Forced Water Invasion in Angular Nanopores With
Mixed Wettability. Journal of Chemical Physics 141: 194703. https://doi.org/10.1063/1.4901752.
Sethian, J. A. 1999. Level Set Methods and Fast Marching Methods: Evolving Interfaces in Computational Geometry, Fluid Mechanics, Computer
Vision, and Material Science. Cambridge, UK: Cambridge University Press.
Sivanesapillai, R., Falkner, N., Hartmaier, A. et al. 2016. A CSF-SPH Method for Simulating Drainage and Imbibition at Pore-Scale Resolution While
Tracking Interfacial Areas. Advances in Water Resources 95: 212–234. https://doi.org/10.1016/j.advwatres.2015.08.012.
Spanne, P., Thovert, J. F., Jacquin, C. J. et al. 1994. Synchrotron Computed Microtomography of Porous-Media—Topology and Transports. Physical
Review Letters 73 (14): 2001–2004. https://doi.org/10.1103/PhysRevLett.73.2001.
Thiele, C., Araya-Polo, M., Alpak, F. O. et al. 2017. Inexact Hierarchical Scale Separation: A Two-Scale Approach for Linear Systems From Discontinu-
ous Galerkin Discretizations. Computers and Mathematics With Applications 74 (8): 1769–1778. https://doi.org/10.1016/j.camwa.2017.06.025.
Ubbink, O. and Issa, R. I. 1999. Method for Capturing Sharp Fluid Interfaces on Arbitrary Meshes. Journal of Computational Physics 153 (1): 26–50.
https://doi.org/10.1006/jcph.1999.6276.
Unverdi, S. O. and Tryggvason, G. 1992. A Front-Tracking Method for Viscous, Incompressible, Multifluid Flows. Journal of Computational Physics
100 (1): 25–37. https://doi.org/10.1016/0021-9991(92)90307-K.
Wells, G. N., Kuhl, E., and Garikipati, K. 2006. A Discontinuous Galerkin Method for the Cahn-Hilliard Equation. Journal of Computational Physics
218 (2): 860–877. https://doi.org/10.1016/j.jcp.2006.03.010.
Wildenschild, D., Vaz, C. M. P., Rivers, M. L. et al. 2002. Using X-ray Computed Tomography in Hydrology: Systems, Resolutions, and Limitations.
Journal of Hydrology 267 (3–4): 285–297. https://doi.org/10.1016/S0022-1694(02)00157-9.
Wildenschild, D. and Sheppard, A. P. 2013. X-ray Imaging and Analysis Techniques for Quantifying Pore-Scale Structure and Processes in Subsurface
Porous Medium Systems. Advances in Water Resources 51: 217–246. https://doi.org/10.1016/j.advwatres.2012.07.018.
Wodo, O. and Ganapathysubramanian, B. 2011. Computationally Efficient Solution to the Cahn–Hilliard Equation: Adaptive Implicit Time Schemes,
Mesh Sensitivity Analysis and the 3D Isoperimetric Problem. Journal of Computational Physics 230 (15): 6037–6060. https://doi.org/10.1016/
j.jcp.2011.04.012.

2018 SPE Journal 17

ID: jaganm Time: 17:25 I Path: S:/J###/Vol00000/180034/Comp/APPFile/SA-J###180034


J182607 DOI: 10.2118/182607-PA Date: 8-June-18 Stage: Page: 18 Total Pages: 18

Yue, P., Feng, J. J., Liu, C. et al. 2004. A Diffuse-Interface Method for Simulating Two-Phase Flows of Complex Fluids. Journal of Fluid Mechanics
515: 293–317. https://doi.org/10.1017/S0022112004000370.

Florian Frank was a senior post-doctoral-degree research associate at Rice University in the Department of Computational and
Applied Mathematics and a contractor at Shell International Exploration and Production (USA). With expertise in numerical anal-
ysis, mathematical modeling, and computer science, Frank contributed to several publications in the field of digital rock. He
holds a diploma in mathematics and a PhD degree in applied mathematics from the Friedrich-Alexander-Universität Erlangen-
Nürnberg, where he currently works as a deputy professor.
Chen Liu is a PhD-degree candidate in the Department of Computational and Applied Mathematics at Rice University. His
research interests include DG methods, numerical analysis, the NS equation, pore-scale modeling, and scientific computing.
Faruk O. Alpak is a senior research reservoir engineer in the Upstream Computational Science Team of Shell International Explo-
ration and Production (USA) and an adjunct associate professor at Rice University in the Computational and Applied Mathemat-
ics Department. Before joining Shell in 2005, he worked at the Schlumberger-Doll Research Center as a visiting scientist on
mathematical modeling and inversion projects. Alpak’s specialization areas include reservoir-simulator development and appli-
cation (linear and nonlinear solvers, flow and geomechanics coupling, upscaling and multiscale methods, parallelization,
assisted history matching, and thermal enhanced oil recovery), in-situ conversion and upgrading processes, CFD-based DNS for
digital rock applications, model-based optimization, inverse problems, computational geomechanics, and computational elec-
tromagnetics. He holds a BS degree in petroleum and natural-gas engineering from the Middle East Technical University, Turkey,
and MS and PhD degrees in petroleum engineering from The University of Texas at Austin. Alpak is an associate editor for SPE
Journal. He received an SPE A Peer Apart Award and the 2008 SPE Journal Outstanding Associate Editor Award and the Best Pa-
per Award from the Society of Well Log Analysts Petrophysics Journal in 2003 and 2006. Alpak serves currently on the organizing
committees of the SPE Annual Technical Conference and Exhibition and the SPE Reservoir Simulation Conference.
Steffen Berg is a senior reservoir engineer at Shell Global Solutions International in the Netherlands. He is also a visiting reader in
the Earth Science and Engineering Department and the Chemical Engineering Department at Imperial College London. Berg
holds an MS degree in materials science from the University of the Saarland and a PhD degree in physics from the University of
Mainz/Max Planck Institute for Polymer Research at Mainz, Germany. After a post-doctoral period at Princeton University, he
joined Shell as a research scientist in the Subsurface Research and Development Department. Current research interests are dig-
ital rock, pore-scale physics, and imaging multiphase flow in porous media by X-ray computed tomography including
improved-oil-recovery and enhanced-oil-recovery processes. Berg has authored or coauthored more than 50 journal articles
and 120 conference contributions. He serves as a technical editor for SPE Journal and SPE Reservoir Evaluation & Engineering for
which he has received multiple recognition awards.
Beatrice Riviere holds the Noah Harding Chair and is a professor in the Department of Computational and Applied Mathematics
at Rice University, where she has been on the faculty for 10 years. Riviere has worked extensively in the development and analy-
sis of numerical methods applied to problems in porous media and fluid mechanics. She has authored or coauthored more than
90 scientific publications in numerical analysis and scientific computation. Riviere is the author of a book on the theory and
implementation of DG methods. She holds a PhD degree in computational and applied mathematics from the University of
Texas at Austin and an engineering degree from École Centrale in France.

18 2018 SPE Journal

ID: jaganm Time: 17:25 I Path: S:/J###/Vol00000/180034/Comp/APPFile/SA-J###180034

You might also like