Sie sind auf Seite 1von 23

IMA Journal of Applied Mathematics (2012) 77, 887–909

doi:10.1093/imamat/hxs048
Advance Access publication on July 26, 2012

Numerical scheme for coupling two-phase compositional porous-media flow


and one-phase compositional free flow

Katherina Baber∗ , Klaus Mosthaf, Bernd Flemisch and Rainer Helmig


Department of Hydromechanics and Modelling of Hydrosystems, University of Stuttgart,
Pfaffenwaldring 61, 70569 Stuttgart, Germany

Corresponding author: katherina.baber@iws.uni-stuttgart.de
Steffen Müthing
Institute for Visualization and Interactive Systems, University of Stuttgart, Universitätsstraße 38,
70569 Stuttgart, Germany
and
Barbara Wohlmuth
M2 - Zentrum Mathematik , TU Munich, Boltzmannstraße 3, 85748 Garching, Germany
[Received on 10 May 2011; accepted on 7 November 2011]

We present the numerical implementation of a model that couples single-phase free and two-phase
porous-media flow and transport with the focus on the treatment of the interface conditions. The model
concept is based on the two-domain approach and on non-isothermal compositional submodels. These
are coupled by interface conditions accounting for mass, momentum and energy transfer, and guaran-
teeing continuity of fluxes. A vertex-centred finite-volume scheme is used throughout the domain and
a global matrix is solved for the whole system incorporating the coupling matrices. The fluxes at the
interface are calculated indirectly via a flux balance from the adjacent finite-volume boxes. Numerical
examples, representing evaporation from partially saturated porous media influenced by an ambient air
flow, illustrate the evolution of saturation and temperature in a reference case and demonstrate the cou-
pling concept. Furthermore, the effect of temperature, Beavers–Joseph coefficient and permeability on
the evaporation process are examined with a series of simulations. In the presented set-ups, the choice of
the Beavers–Joseph coefficient has a negligible influence on the evaporation rate across the interface.

Keywords: coupling conditions, free flow, porous media, interface, compositional flow, two-phase flow,
box-method, evaporation.

1. Introduction
Flow and transport processes in domains composed of a porous medium and an adjacent free-flow region
appear in a wide range of industrial, medical and environmental applications. Industrial applications
range from flow in fuel cells to industrial drying processes, while the interplay of distribution processes
in blood vessels and in the surrounding tissue is a possible medical application.
In environmental systems, evaporation is an important process, since evaporation rates and patterns
affect the energy balance of terrestrial surfaces and drive a multitude of climatic processes. The predic-
tion of evaporative drying rates from porous media remains a challenge due to the ambient conditions at
the interface (radiation, humidity, temperature, air velocity, turbulent conditions). Moreover, the inter-
nal porous-medium properties lead to abrupt transitions and rich flux dynamics. The factors involved


c The authors 2012. Published by Oxford University Press on behalf of the Institute of Mathematics and its Applications. All rights reserved.
888 K. BABER ET AL.

Fig. 1. Relevant interface processes for evaporation: solar radiation, small and larger scale turbulence, vaporization and conden-
sation at the interface and exchange fluxes (Mosthaf et al., 2011).

are coupled by the complex interactions between the porous medium and the free-flow system (see
Fig. 1). Other examples for environmental applications are infiltration of overland flow during rainfall
and groundwater contamination due to infiltrating pollutants.
Modelling such coupled systems while accounting for the respective processes in both domains is a
challenging task, especially since many of these systems are dominated by multi-phase compositional
flow. In Mosthaf et al. (2011), a coupling concept for non-isothermal two-phase compositional flow in
the porous medium in contact with a laminar single-phase non-isothermal compositional system in the
free-flow region was developed. It is based on existing approaches and valid on the REV scale. The
employed coupling conditions for mass, momentum and energy account for the physics at the interface
and are based on flux continuity and thermodynamic equilibrium and justified by phenomenological
explanations. The resulting coupled model is flexible with respect to the subdomain models which are
combined by clearly defined coupling conditions. It allows a detailed description of transfer processes
like, e.g. evaporation influenced by a laminar wind field.
The developed model is based on existing approaches, which to our knowledge have been applied
only to single-phase systems. Coupling free flow and flow in porous media has been extensively stud-
ied and coupling concepts have been developed based on volume averaging, upscaling and matched
asymptotic expansion (Ochoa-Tapia & Whitaker, 1995; Jäger et al., 2001; Chandesris & Jamet, 2006,
2007, 2009; Jäger & Mikelić, 2009). The resulting coupling conditions range from continuity of thermo-
dynamic properties to discontinuities defined by jump coefficients (Layton et al., 2002; Valdés-Parada
et al., 2009a). Valdés-Parada et al. (2009b) describe diffusive and convective transport processes in a
coupled system. A comprehensive overview of the developments in the field of coupling free flow and
porous media can be found in the special issue ‘Transport Phenomena at the Interface Between Fluid
and Porous Domains’ (Shavit, 2009). Mosthaf et al. (2011) give a summary of the existing approaches.
Various discretization techniques, such as finite elements (FE), mixed methods, control-volume
techniques and Discontinuous-Galerkin (DG) approaches have been developed and analysed for coupled
COUPLING TWO-PHASE COMPOSITIONAL POROUS-MEDIUM FLOW 889

single-phase single-component systems. In Discacciati et al. (2002) and Discacciati et al. (2007),
a formulation based on the Beavers–Joseph–Saffman interface condition (Beavers & Joseph, 1967;
Saffman, 1971) is studied, continuous finite elements are used in both regions and subdomain iterative
methods for solving coupled problems are developed. Cao et al. (2010) apply the Beavers–Joseph con-
dition without the Saffman simplification. The application of the continuous FE method for the Stokes
problem with a mixed FE method for the Darcy problem is considered in Layton et al. (2002), where the
coupling conditions are imposed with the help of Lagrange multipliers. The existence and uniqueness
of a weak solution of the coupled system is proved. DG methods for the coupled system are proposed
in Rivière (2005) and Girault et al. (2009), and the existence of a weak solution with some a priori
estimates is shown. Discacciati & Quarteroni (2004) use a Galerkin Finite Element approximation in
both domains and propose iterative subdomain methods based on either a Dirichlet–Neumann method
or a Conjugate Gradient method for the solution of the Steklov–Poincaré interface problem. In Rivière
& Yotov (2005), the DG method is used in the Stokes region and coupled to a mixed FE method in the
Darcy region. Mixed FE methods are developed and analysed for coupled systems with large cavities
in the porous medium by Arbogast et al. (2007). Quite recently, new types of mixed DG techniques are
successfully applied to the Stokes–Darcy coupling (Rivière & Kanschat, 2010). Chidyagwai & Riviére
(2010) compare different combinations of DG and finite-element methods with respect to accuracy,
robustness and mass conservation. However, all these techniques have been developed for a single-
phase single-component system.
For the discretization of the present compositional coupled model, the implicit Euler time-
integration and a vertex-centred control-volume finite-element method (also called box method, Huber
& Helmig, 2000) in space is used. In the free-flow domain, a stabilization technique similar to Franca
et al. (1993) is generalized and implemented for the Stokes equations. Special care at the interface is
required to avoid spurious oscillations. The arising non-linear system is solved fully implicitly applying
a Newton solver.
In the following, the developed model concept consisting of subdomain models and coupling
conditions will be briefly explained (see Section 2). The focus of this work, however, is on the numerical
concept and on the implementation of the coupled model into the modelling toolbox DuMux (Flemisch
et al., 2011). The discretization and handling of the coupling conditions at the interface are explained
in Section 3. Finally, numerical examples illustrate the model concept. Several parameters (tempera-
ture, Beavers–Joseph coefficient, permeability) are varied and the influence on the evaporation rate is
demonstrated.

2. Model concept
The focus of this paper is on the coupling of free flow and porous-media flow systems. The developed
concept is based on existing model concepts in the subdomains. To simplify the free-flow system, we
assume Stokes flow, whereas Darcy’s law is employed in the porous medium. For the coupling, two
basic strategies exist (Jamet et al., 2009; Shavit, 2009). In the one-domain approach, one system of
equations is used in the whole domain and the transition between the two domains is achieved by spa-
tial variation of physical properties such as permeability and porosity (Brinkman, 1947; Goyeau et al.,
2003; Rosenzweig & Shavit, 2007). The two-domain approach employs coupling conditions at a dis-
tinct interface and allows the use of different sets of equations in the two domains. The challenge is to
define coupling conditions that account for the transfer processes at the interface. Various approaches
can be found ranging from conditions involving jump coefficients (Ochoa-Tapia & Whitaker, 1995;
Chandesris & Jamet, 2009) to the continuity of thermodynamic quantities at the interface (Layton et al.,
890 K. BABER ET AL.

Fig. 2. Schematic depiction of the two-domain coupling concept (Mosthaf et al., 2011).

2002). Here, the coupling of the two domains is achieved using the two-domain approach and assuming
continuity of fluxes and local thermodynamic equilibrium at the interface. The interface is assumed to
be simple in the sense that it cannot store mass, momentum or energy (Hassanizadeh & Gray, 1989).
Furthermore, the Beavers–Joseph–Saffman condition is employed in the knowledge of its theoretical
limitation to single-phase, parallel flow (Beavers & Joseph, 1967; Saffman, 1971). This condition is
theoretically justified in Jäger & Mikelić (2000). The developed coupling conditions are valid on the
REV scale and have been developed based on phenomenological explanations. In a first step, interface
processes like the formation of boundary layers, flow separation or radiation, and the influence of tur-
bulent flow are neglected. In the following, the subdomain models and the coupling concept are briefly
explained. For more details we refer to Mosthaf et al. (2011).
Figure 2 illustrates the problem setting: two domains Ω ff and Ω pm are separated by the interface
Γ = ∂Ω ff ∩ ∂Ω pm with the outward unit normal vectors nff and npm . For simplicity of notation, the
superscripts (ff) and (pm) are only applied for the quantities at the interface, where (ff) refers to the
values in the free-flow subdomain and (pm) stands for the porous-medium side. As was stated above, a
non-isothermal one-phase compositional free-flow system is coupled to a respective two-phase system
in the porous medium. In both domains, balance equations for mass, momentum and energy are solved.
In order to simplify the system, slow, creeping flow of Newtonian fluids is assumed.

2.1 Subdomain model: porous medium


The porous-medium model is based on the following assumptions:
• local thermodynamic (mechanical, thermal and chemical) equilibrium,
• a rigid solid phase (subscript s),
• two-phase flow consisting of a liquid phase (subscript l) and a gas phase (subscript g),
• two components being present in each phase: water (superscript w) and air (superscript a),
• air considered to be an ideal gas,
• slow flow velocities (Re  1) allowing an application of multi-phase Darcy’s law for the phase
velocities,
COUPLING TWO-PHASE COMPOSITIONAL POROUS-MEDIUM FLOW 891

Table 1 Porous-medium equations and associated primary variables

Balance equations Primary variable


Mass balance for component κ ∈ {w, a}:
 ∂(α X κ Sα ) 
α
φ + ∇ · Fκ − qκα = 0 Sl or Xακ
α∈{l,g} 
∂t α∈{l,g}
κ
F = (α vα Xα − Dα,pm α ∇Xακ )
κ κ

α∈{l,g}
Mass balance:
 ∂(α Sα ) 
φ + ∇ · Fm − qα = 0 pg
α∈{l,g} 
∂t α∈{l,g}
Fm = (α vα )
α∈{l,g}
Darcy’s law:
krα
vα = − K(∇pα − α g), α ∈ {l, g}
μα
Energy balance:
 ∂(α uα Sα ) ∂(s cs T)
φ + (1 − φ) + ∇ · FT − qT = 0 T
α∈{l,g} 
∂t ∂t
FT = α hα vα − λpm ∇T
α∈{l,g}

• binary diffusion,
• negligible influence of dispersion due to slow flow velocities and comparatively high diffusion coef-
ficients in the gas phase.
The model allows the transfer of components from one phase into another and is thus able to describe
vaporization and condensation as well as dissolution and degassing within the porous medium. Table 1
summarizes the employed balance equations and chosen primary variables. The non-isothermal model
comprises two mass-balance equations and an energy balance. The definitions of the respective param-
eters, material laws and equations of state are listed in Table 2.
The mass conservation in Ω pm is expressed by one mass-balance equation for each component
κ ∈ {w, a}, where φ is the porosity, Sα is the saturation of the phase α, Xακ stands for the mass fraction of
the component κ in the phase α, α is the phase density (see Table 2) , Dκα,pm is the diffusion coefficient
in phase α in the porous medium (see Table 2) and qκα are given source or sink terms. The Darcy
velocity vα is calculated using the multi-phase Darcy law with the relative permeability krα of phase α,
the dynamic viscosity μα of phase α, the tensor of intrinsic permeability K and the gradient of the phase
pressure ∇pα . Summing up the mass-balance equations for the two components, with Xαw + Xαa = 1, and
assuming binary diffusion, we obtain the total mass-balance equation. In this case, the source and sink
terms are also summed up for the components, qα = qwα + qaα . It is now possible to choose equivalently
two of the three balance equations for a complete model description.
Under the assumption of local thermal equilibrium (Tl = Tg = Ts = T), justified by slow flow
conditions, one single energy-balance equation that accounts for convective and conductive heat fluxes,
the heat sources qT within the domain and the storage of heat in the fluid phases and the porous material
892 K. BABER ET AL.

Table 2 Material laws and equations of state dependent on primary variables

Equations of state References


Parameters
Density α α = f (κ , xκα , pα , T) IAPWS (2009)
Component density κ Incompressible fluid and ideal gas Reid et al. (1987)
Diffusion coefficient Dα,pm Dα,pm = τ φSα Dα Millington & Quirk (1960)
(φSα )7/3
Tortuosity τ τ= Millington & Quirk (1960)
φ2
Capillary pressures pc Brooks–Corey model pc = f (Sl ) Brooks & Corey (1964)
Relative permeability krα Brooks–Corey√ model Brooks & Corey (1964)
Eff. heat conductivity λpm λpm = λeff,g + Sl (λeff,l − λeff,g ) Somerton et al. (1974)
Internal energy uα uα = hα − pα /α
Enthalpy h
Gas phase hg = Xgw hw + Xga ha
Water phase hl = hw
Component air ha = 1005(T − 273.15 K)
Component water hw = f (pα , T) see IAPWS (2009)
Secondary variables
Saturation Sg Sg = 1 − Sl
Liquid-phase pressure pl pc = pg − pl Brooks & Corey (1964)
Air in water phase xal Henry’s law: xal = pag /Hgla
Vapour in gas phase xwg xwg = pwsat,Kelvin /pg  
pc
where psat,Kelvin = psat exp −
w w
Kelvin equation
l Rl T
Mass fractions Xαw + Xαa = xwα + xaα = 1
Xακ = xκα Mκ /(xwα Mw + xaα Ma )

can be applied. In the energy equation, uα is the internal energy of the phase α, s and cs are the density
and specific heat capacity of the porous medium, hα is the specific enthalpy and λpm is the effective
heat conductivity of the porous medium which depends on the liquid saturation Sl and the effective
heat conductivities of a gas- or water-saturated porous medium λeff,g and λeff,l (see Table 2). A detailed
explanation of the model concept applied in the porous medium and of the underlying assumptions can
be found in Class et al. (2002).
Primary variables have to be chosen and the secondary variables have to be expressed as functions
of the primary variables. Here, the gas-phase pressure pg , the saturation of the liquid phase Sl and the
temperature T are chosen as primary variables for the two-phase system. Based on this, the secondary
variables can be calculated using the relations in Table 2. Here, pc is the capillary pressure, xκα and
Xακ are the mole and mass fractions of component κ in phase α, and M κ are the molar masses of the
components. If one phase disappears, the system of equations is significantly simplified. However, the
saturation cannot be used as a primary variable anymore. Possible choices for the numerical algorithm
are then a primary-variable switch (Class et al., 2002) or the introduction of additional primary variables
and additional constraints in the form of non-linear complementarity functions, as proposed in Lauser
et al. (2011). We use the former approach here.
COUPLING TWO-PHASE COMPOSITIONAL POROUS-MEDIUM FLOW 893

Table 3 Free-flow equations and associated primary variables

Balance equation Primary variable


Mass balance for component κ:
∂(g Xgκ )
+ ∇ · Fκ − qκg = 0, Fκ = g vg Xgκ − Dκg g ∇Xgκ Xακ
∂t
Mass balance:
∂g
+ ∇ · Fm − qg = 0, Fm = (g vg ) pg
∂t
Momentum balance (Stokes equation):
∂(g vg )
+ ∇ · Fv − g g = 0, vg
∂t
Fv = pg I − μg (∇vg + ∇vTg )
Energy balance:
∂(g ug )
+ ∇ · FT − qT = 0, FT = g hg vg − λg ∇T T
∂t

2.2 Subdomain model: free flow


The free-flow domain is based on the following setup:
• laminar, creeping flow,
• binary diffusion,
• neglect of inertial forces due to low Reynolds numbers,
• single-phase flow (here a gas phase is considered: α = g),
• two components in the gas phase: water (w) and air (a).
The four equations which are solved in the free-flow domain are shown in Table 3: two scalar
mass-balance equations (one for the water component and one for the total mass), one vector-valued
momentum-balance equation and one scalar energy-balance equation. The same notation as in the
porous-medium region is used and a compressible gas phase, which behaves like an ideal gas, is con-
sidered. Neglecting the non-linear inertial term of the Navier–Stokes equation and assuming gravity to
be the only external force, the Stokes equation results as momentum-balance equation. The unsteady
Stokes equation shown in Table 3 is based on Newton’s law and includes the symmetrized gradient of
the velocity. In the energy balance, the equations of state shown in Table 2 are again used to calculate
enthalpy and internal energy.
As primary variables, the pressure of the gas phase pg , the mass fraction of water in the gas phase
Xgw , the velocity of the gas phase vg and the temperature T are chosen.

2.3 Coupling conditions


Suitable coupling conditions have to be found that account for the physics at the interface and the
transfer processes of mass, momentum and energy. The coupling conditions should be valid on the
REV scale but still account for the pore-scale processes. By making reasonable assumptions based on
phenomenological explanations, coupling conditions have been deduced in Mosthaf et al. (2011) that
are motivated by thermodynamic equilibrium and get as close as possible to a description for a so-called
894 K. BABER ET AL.

simple interface that has no thickness and is devoid of thermodynamic properties (Hassanizadeh & Gray,
1989). The reasoning behind the coupling conditions is explained in detail in Mosthaf et al. (2011). For
brevity, we only list them here.
In particular, the mechanical equilibrium (equilibrium of forces) is given by

(1) the continuity of the normal stresses resulting in a possible jump in the gas-phase pressure

n · [((pg I − μg (∇vg + ∇vTg ))n)]ff = [pg ]pm , (2.1)

(2) the continuity of the normal mass fluxes

[g vg · n]ff = −[(g vg + l vl ) · n]pm , (2.2)

(3) the Beavers–Joseph–Saffman condition for the tangential component of the free-flow velocity
(Beavers & Joseph, 1967; Saffman, 1971)
 √  ff
ki
vg + (∇vg + ∇vg )n · ti = 0,
T
i ∈ {1, . . . , d−1}, (2.3)
αBJ

where ki = t · (K t) is the tangential component of the permeability tensor.

The thermal equilibrium is given by (see e.g. Alazmi & Vafai (2001))

(1) the continuity of the temperature


[T]ff = [T]pm , (2.4)

(2) the continuity of the heat fluxes

[(g hg vg − λg ∇T) · n]ff = −[(g hg vg + l hl vl − λpm ∇T) · n]pm . (2.5)

The chemical equilibrium is represented by

(1) the continuity of mass fractions


[Xgκ ]ff = [Xgκ ]pm , (2.6)

(2) the continuity of the component fluxes across the interface

[(g vg Xgκ − Dg g ∇Xgκ ) · n]ff = −[(g vg Xgκ − Dg,pm g ∇Xgκ + l vl Xlκ − Dl,pm l ∇Xlκ ) · n]pm .
(2.7)
Based on the choice in the submodels, two of the three conditions (2.2), (2.7) have to be
employed.

3. Numerical model
In this section, the employed numerical model is presented. The focus is on the numerical details of
the implementation of the coupling conditions. We present the discretization scheme (Section 3.1),
the choice of primary variables (Section 3.2), the handling of interface and boundary conditions
(Section 3.3) and some details on the computational algorithm (Section 3.4).
COUPLING TWO-PHASE COMPOSITIONAL POROUS-MEDIUM FLOW 895

(a) (b)

(c)

Fig. 3. Schematic description of the spatial discretization (box scheme).

3.1 Discretization scheme


For the discretization of the problem, we use the control-volume finite-element method (CVFEM or
box method) in both subdomains Ω ff and Ω pm for the space discretization (Masson et al., 1994; Huber
& Helmig, 2000) and an implicit Euler time discretization. The method is locally mass conservative.
We are aware of the fact that more advanced discretization methods are available. For our purposes,
however, a simple scheme is sufficient since the intention of this work is to demonstrate the numerical
coupling concept.
The computational domain is covered by a structured finite-element grid (primal mesh). The values
of all primary variables are computed at the vertices, which are the corners of the elements. The dual
mesh which describes the finite volumes is obtained by connecting the centres of gravity of each element
with the associated edge midpoints. Thus, each finite volume is associated with a vertex (see box Bi in
Fig. 3(a)). The FE mesh divides the boxes Bi into subcontrol volumes vki (see Fig. 3(b)). We consider
a quadrilateral grid and approximate all primary variables using piecewise bilinear functions (equal
order method), which are mapped from the standard-reference-element basis functions. The equations
are integrated over the control volumes and the divergence theorem is applied. The method of weighted
residuals with piecewise constant weighting functions characteristic for box Bi and mass lumping of the
accumulation terms leads to the general discretized form for one box Bi :

ûn+1 − ûni  
|Bi | i
+ FA (ũn+1 (xf )) · nf |sf | + FD (ũn+1 (xf )) · nf |sf | − |Bi |qn+1 = 0, (3.1)
Δt f ∈s(i) f ∈s(i)
i

where |Bi | is the volume of box Bi , ûn+1


i is the solution at node i with u standing for one of the primary
variables listed in Tables 1 and 3, s(i) are all subcontrol-volume faces of the box Bi , nf is the unit
outward normal of the respective subcontrol-volume face in direction of the neighbouring node j, |sf | is
the size of the subcontrol-volume face and qn+1 i is the source and sink term. Equation (3.1) consists of
storage, flux and source/sink terms. The fluxes may be composed of an advective and a diffusive part
FA , FD and are computed on the control-volume boundaries sf at the integrationpoints xf (see Fig. 3(c)).
A finite-element approximation for each primary variable is done, ũn+1 (xf ) = k∈ηE Nk (xf )ûn+1k , where
896 K. BABER ET AL.

ηE are the nodes of the element E. Furthermore, a fully upwind scheme is applied to the advective part.
Following the general procedure employed for finite-element methods, local operators are assembled
for each element and then merged for the whole domain. This means that the discretized equations are
evaluated element-wise, looping over the subcontrol volumes inside an element E (see Fig. 3(c)). In the
following, we will refer to the left-hand side of the equations as residuum R. The residuum denotes the
assembled discretized equations for one node including storage, flux, and source and sink terms.
For the mass-balance equation in the porous medium, the box method yields:
⎛ ⎞
    n+1 
(αi Ŝαi ) n+1
− (αi Ŝαi ) n
krα
|Bi |Φi − α Kij ⎝ (∇Nαk (xf )p̂n+1 n+1 ⎠
αk ) + αi g
α∈l,g
Δt f ∈s(i) α∈l,g
μ α ups(i,j) k∈η
E

· nf |sf | − |Bi |qαi
n+1
= 0. (3.2)
α∈l,g

An upstream weighting of the non-linear terms is performed, where the subscript ups(i, j) is the upstream
node depending on the direction of the velocity which is used for the upwinding of the relative perme-
abilities, the viscosities and the densities. Soil parameters are defined at the nodes (valid for one box).
Kij is the harmonic mean of the intrinsic permeabilities of the neighbouring nodes i and j.
In the free-flow region, the box method leads to possible oscillations in the pressure, resulting from
the use of equal-order approximations for pressure and velocity, which does not guarantee full stability.
To overcome this problem, we apply a similar stabilization technique to the continuity equation as the
one proposed in Franca et al. (1993):

gi
n+1
− gi
n  n+1
|Bi | + (Nk (xf )v̂gk ) · nf |sf | + |Bi |qn+1
Δt f ∈s(i) k∈ηE
gi

 
− αstab h2f (∇Nk (xf )p̂n+1
gk ) · nf |sf | − |Bi |αstab hf ∇ · (gi g) = 0.
2 n+1
(3.3)
f ∈s(i) k∈ηE
 
stabilization

The stabilization terms are obtained by taking the divergence of the Stokes equation ∇ · (∇pg − μ∇ 2 v −
g) multiplied by a stabilization parameter αstab h2 , involving the mesh width h.
The Stokes equation has the following discretized form:

(gi v̂gi )n+1 − (gi v̂gi )n    n+1 n+1 T


|Bi | + |Bi | (∇Nk (vki )p̂n+1
gk ) − μgi
n+1
(∇Nk (xf )v̂gk + ∇Nk (xf )T v̂gk )
Δt k∈η f ∈s(i) k∈η
E E

· nf |sf | − |Bi |qn+1


i = 0. (3.4)

Due to stability reasons, the pressure term in the Stokes equation is approximated as a volume term in
(3.4) instead of including it in the flux term. Consequently, the pressure term and the viscous term are
of similar order of magnitude.
Furthermore, spurious oscillations occur at the corners of the Stokes domain, both at the points
where the exterior boundary meets the coupling interface (∂Ω ff ∩ Γ ) and at the corners at the top of the
domain. The oscillations at the interface nodes are due to the fact that the boundary conditions of the
free-flow and the porous-medium domain are linked via the coupling conditions and do not necessarily
COUPLING TWO-PHASE COMPOSITIONAL POROUS-MEDIUM FLOW 897

2 3

BC free−flow region
(e.g. Dirichlet)
0 1
coupling

BC porous-medium region
(e.g. Neumann)

Fig. 4. Treatment of the nodes where the coupling interface meets the boundaries of the subdomains.

match (see Fig. 4). Since temperature and mass fractions are continuous quantities, reasonable boundary
conditions can be defined. Due to the discontinuity in the gas-phase pressure (2.1), oscillations occur
which are avoided by replacing the residuum at the corner node by an interpolation and thus, depriving
the corner node of its degrees of freedom. This procedure is applied only for the mass balance and must
not be done if Dirichlet conditions are set for the pressure.

3.2 Primary variables


In order to solve the system described above, suitable primary variables have to be chosen. In the
porous domain Ω pm , the pressure of the gas phase pg , the saturation of the liquid phase Sl and the
temperature T are used. But, since fluid phases can appear and disappear, a primary-variable switch has
to be implemented in the model (Class et al., 2002). If the liquid phase is not present, Sl = 0, and we
have a single-phase system, the liquid saturation is replaced by the mass fraction of water in the gas
phase Xgw as primary variable. The switch back to the liquid saturation is triggered when the partial
pressure of water in the gas phase exceeds the saturation vapour pressure.
In the free-flow domain Ω ff , the pressure of the gas phase pg , the mass fraction of water in the gas
phase Xgw , the velocity of the gas phase vg and the temperature T are chosen as primary variables.

3.3 Interface and boundary conditions


Both the porous-medium (Table 1) and the free-flow (Table 3) model have to be completed by
appropriate boundary conditions. Based on the chosen set of primary variables, the boundary condi-
tions can be specified as

fNκ , on ΓN ,
pm pm
pg = pD on ΓD , Fm · n =
κ
pm pm
Sl = SD on ΓD , F ·
w
n = fNw , on ΓN ,
pm pm
T = TD on ΓD , FT · n = fN,T on ΓN ,
898 K. BABER ET AL.

Table 4 Interpretation of coupling conditions as boundary conditions

Property Porous medium Free flow


Mechanical equilibrium:
Normal stress (2.1) Neumann for vg · n
Tangential velocity (2.3) Robin for vg · t
Thermal equilibrium:
Temperature (2.4) Dirichlet for T
Heat flux (2.5) Robin for T
Chemical equilibrium:
Mole fraction (2.6) Dirichlet for Xgw
Component flux w (2.7) Mass flux (2.2) Non-linear coupled
conditions for p, S, T

pm pm pm pm
where Γ pm = ∂Ω pm \ Γ , Γ pm = ΓD ∪ ΓN , ΓD ∩ ΓN = ∅, in the porous medium region, and

vg = vD on ΓDff , Fv n = fN,v on ΓNff ,


Xgw = XD on ΓDff , Fκ · n = fN,X on ΓNff , (3.5)
T = TD on ΓDff , FT · n = fN,T on ΓNff ,

where Γ ff = ∂Ω ff \ Γ , Γ ff = ΓDff ∪ ΓNff , ΓDff ∩ ΓNff = ∅, in the free-flow domain. For the Neumann
conditions, the normal flux across the boundary F · n is specified by the Neumann flux fN . The respective
flux terms are defined in Tables 1 and 3. Note that the Neumann flux for the Stokes equation is defined as
fN,v = [pg I − μg (∇vg + ∇vTg )]n. Since the pressure term is chosen to be discretized as a volume and not
as a flux term in (3.4), a correction for the pressure contribution has to be applied when a Neumann con-
dition is set for the Stokes equation. The free-flow mass-balance equation requires a Dirichlet condition
for pressure in at least one point and all other boundaries are set to outflow conditions.
The boundary conditions are supplemented by conditions at the interface Γ . The interface conditions
play the role of internal boundary conditions for the coupled model, meaning that the information is
handed over from the free-flow to the porous-medium side and vice versa. An overview of a possi-
ble combination and implementation of the presented conditions can be found in Table 4. An outflow
condition is set for the mass balance equation in the free-flow domain at the interface.
Figure 5 shows the discretization at the interface. The interface is located at the border of two
elements on the nodes, cutting a FV box in half. The interface nodes belong to both domains and thus,
contain two sets of primary variables. In contrast to Neumann-type conditions which enter the weak
form in a natural way as term on the right side, Dirichlet-like coupling conditions are imposed in a
strong form. Hence, the unknown u in the free-flow domain (e.g. (2.4) or (2.6)) replaces the residuum
of the respective free-flow equation with uff − upm = 0.
The implementation of the flux conditions requires a special treatment due to the characteristics of
the discretization scheme. The flux conditions (2.5) and (2.7) are third-type or Robin conditions which
are a combination of the gradient of the primary variables directly at the interface and the primary
variable itself. As illustrated in Fig. 5, these gradients have to be approximated at the interface between
two elements and not, as normally done, at the subcontrol-volume faces sf . At the edge of the elements,
however, the gradient of the Ansatz functions is discontinuous causing a discontinuity in the fluxes. As a
consequence, these fluxes cannot be used as condition for the other domain.
COUPLING TWO-PHASE COMPOSITIONAL POROUS-MEDIUM FLOW 899

Fig. 5. Illustration of the spatial discretization (box scheme) at the interface.

An alternative approach for calculating the normal fluxes across the interface is based on the fact
that the total sum of the change in storage and all fluxes across the faces of a control volume has to be
zero. Hence, the sum of the storage term and the interior fluxes across one half of the control volume
adjacent to the interface (see fluxes across the dotted line in Fig. 5) is equal to the normal fluxes across
the interface F · n (see (3.6)).
  n+1  
 Bi  ûi − ûni
  + F(ũ (xf )) · n dΓinternal =
n+1
F · n dΓ (3.6)
2 Δt f ∈s(i) Γ
internal

That means that the flux directly at the interface can be expressed by balancing storage and flux
contributions. The calculated normal flux is then added to the right-hand side of the respective node
in the other domain. Note that where the coupling interface touches the boundaries ∂Ω ∩ Γ , the fluxes
across the external boundaries have to be considered in addition to the internal fluxes to balance the
fluxes across the interface correctly. Moreover, as it was said above, the pressure is interpolated at these
corner points. Furthermore, in case of condition (2.1) and (2.3) employed as Neumann-like condition
for the Stokes equation, a pressure correction has to be applied as it was explained above.

3.4 Computational algorithm


Our implementation of the numerical scheme is based on the modelling toolbox DuMux . DuMux is a free
and open-source simulator for flow and transport processes in porous media, based on the Distributed
and Unified Numerics Environment (DUNE) (Bastian et al., 2008). Its main intention is to provide a
sustainable and consistent framework for the implementation and application of model concepts, con-
stitutive relations, discretizations and solvers (Flemisch et al., 2011). The coupled problem (free-flow
domain, porous-medium domain and interface) is written in the following operator form
∂M(u)
− ∇ · F(u) = Q(u), (3.7)
∂t
pm
where u = (pffg , Xgw,ff , vffg , T ff , ppm
g , Sl or Xgw,pm , T pm )T is the solution vector, M(u) is the storage, F(u)
the flux and Q(u) the source/sink term. The spatial discretization of the geometric domain and the
designation of subdomains are based on the software package DUNE-Multidomaingrid (Müthing &
Bastian, 2011), which allows us to represent arbitrarily complex subdomain shapes with conforming
900 K. BABER ET AL.

Model setup Matrix

Fig. 6. Depiction of the global matrix for the coupled system. Aff and Apm stand for the submatrixes of the free-flow and porous-
medium subproblems. Aff/pm and Apm/ff are the coupling matrixes, where Aff/pm contains the coupling conditions for the free-flow
domain from the porous medium Apm/ff contains the coupling conditions for the porous medium.

interfaces. For the management of the different subdomains and their associated function spaces, we
employ PDELab (Bastian et al., 2010), a general discretization framework based on DUNE, and its
extension DUNE-Multidomain, which allows for a transparent definition of local operators and variables
on the individual subdomains and supports the assembly of the global system matrix for the coupled
problem. The global matrix contains the submatrixes for the two subdomains and the coupling matrixes
(see Fig. 6).
The ability to do a full system assembly enables us to solve the strongly coupled highly non-linear
system (3.7) in a fully implicit fashion without any subdomain iteration scheme. While such an iteration
scheme might become necessary in large-scale applications to efficiently handle the different time scales
in the subdomains, we chose the strongly coupled solution. On the one hand, it eliminates the substantial
task of designing, implementing and testing an iteration scheme, and on the other hand, it provides us
with a high-fidelity benchmark solution to compare against future, weakly-coupled simulations.
In order to solve the non-linear algebraic system at each time step, we apply a standard Newton
solver  
∂R
(un+1,m − un+1,m−1 ) = −R(un+1,m ), (3.8)
∂u n+1,m  
  Δu
J(un+1,m )

leading to a stable and robust scheme. Here, J(un+1,m ) is the Jacobi matrix calculated by numerical
differentiation, Δu is the correction to the primary variables u, and R(un+1,m ) is the residuum at time-
level n + 1 and iteration m. The linear problem at each Newton iteration step is solved using the direct
linear solver SuperLU (Demmel et al., 1999). For the time integration, we use a fully implicit Euler
scheme with a heuristic time-step control based on the convergence rate of the Newton solver. A target
number Ntarget of Newton iterations is defined. Depending on the actual number of iterations N, the new
time-step size is chosen:
⎧  

⎪ Δt n 1
Nmax > N > Ntarget ,

⎨ 1 + (N − Ntarget )/Ntarget
Δt n+1
=   (3.9)

⎪ n Ntarget − N

⎩Δt 1 + N  Ntarget ,
1.2Ntarget

where Δtn is the time-step size of time step n. The time-step size is halved if N exceeds the maximum
number of time-steps, Nmax .
COUPLING TWO-PHASE COMPOSITIONAL POROUS-MEDIUM FLOW 901

Fig. 7. Model setup with respective initial and boundary conditions of the reference case.

4. Numerical examples and parameter study


Several numerical examples demonstrate the implementation of the model concept in DuMux . A
reference case is chosen and single parameters (the temperature, the permeability and the Beavers–
Joseph coefficient) are varied and their influence on the evaporation rate is examined. In all cases, a
two-dimensional setup is chosen where relatively dry air blows horizontally over the surface of a par-
tially water-saturated porous medium that is closed on all other sides. This represents e.g. a sand-filled
vessel in contact with the ambient air flow. The setup is illustrated in Fig. 7. The following simplifica-
tions have been used:
• gravity is not considered;
• the fluid viscosities and densities are set constant

μg = 1.71e−5 Pa s
μl = 0.0013 Pa s
ρg = 1.189 kg/m3
ρl = 1000 kg/m3 .

The domain size is 0.25 m × 0.5 m, discretized with one meta-grid consisting of 20 × 50 rectangular
cells. The meta-grid is bisected into two subgrids, one for the porous medium and one for the free flow
taking advantage of the module DUNE-Multidomaingrid (Müthing & Bastian, 2011). A graded mesh
is employed that becomes finer towards the interface. The finest cells at the interface have a height
902 K. BABER ET AL.

of 5 mm. The influence of this relatively coarse choice of the grid has been examined in a small grid
convergence study, which is presented at the end of this chapter.

4.1 Reference case


The porous medium is initially 50% water-saturated and has homogeneous Neumann conditions on all
sides except at the top boundary (see Fig. 7(a)), where it is coupled to the adjacent free-flow domain.
The gas pressure pg , the water saturation Sl and the temperature T are selected as primary variables.
The initial pressure of the gas phase is chosen as atmospheric, 1e5 Pa, and the temperature is set to
293.15 K. As soil properties, we choose an isotropic permeability of K = 1e−10 m2 and a porosity of
φ = 0.1. A Brooks–Corey parameterization is employed for the capillary pressure—saturation and the
relative permeability—saturation relationship with the parameters pd = 100 Pa and λ = 2. As effective
heat conductivities, λeff,g = 2.5 W/(mK) for the air-saturated soil and λeff,l = 2.65 W/(mK) for the water-
saturated soil are used.
In the free-flow domain, relatively dry air (Xgw = 0.005) flows mainly horizontally over the surface
of the porous medium. Initially, the pressure is set to 1e5 Pa and the temperature to 293.15 K, which is
conform with the initial conditions in the porous medium. A linear velocity profile is set as Dirichlet
condition on the left inflow boundary. The maximum horizontal velocity of vx,max = 2 m/s is at the
top boundary and decreases to the Beavers–Joseph slip velocity at the coupling boundary. It is varied
sinusoidally in the range of 2 ± 0.1 m/s over time with a period of 1 h. This variation of velocity is
meant to illustrate the dynamics of the system, namely the influence of varying wind velocities on
the evaporation rate. Furthermore, a variation in horizontal velocity causes variations in the vertical
velocity field and consequently improves the convergence, since small y-velocities cause problems with
the non-linear Newton solver. The temperature and the mass fraction of vapour in the gas flow at the
left inflow boundary are chosen as T = 293.15 K and Xgw = 0.005 to obtain the Neumann boundary
fluxes as fN,X = (g vg Xgw ) · n or fN,T = (g vg hg ) · n. The gas pressure is fixed as Dirichlet condition to
atmospheric conditions with pg = 1e5 Pa on the right boundary of the domain, on all other boundaries
an outflow condition is set for the total mass balance. All other equations have outflow conditions on
the right boundary. As Beavers–Joseph coefficient for the slip velocity at the coupling interface, we use
αBJ = 1.0.

Fig. 8. Evolution of the saturation. The upper parts of the images show the mass fraction of vapour and the velocity vectors of
the gas phase, the lower parts show the evolution of the saturation distribution in the porous medium. Different points in time are
shown and the scales are fitted to the data range. The first image shows the the initial conditions. In Fig. 8(c,d), a certain region at
the interface has already switched to a single-phase system.
COUPLING TWO-PHASE COMPOSITIONAL POROUS-MEDIUM FLOW 903

Fig. 9. Evolution of temperature at the same points in time as in Fig. 8. Due to the vaporization enthalpy, the porous medium
is cooled as long as two phases are present. In Fig. 9(c), the interface has already dried out and the temperature readjusts to the
ambient temperature in the free flow.

6e-05
T=283 K
T=293 K
T=303 K
5e-05
Evaporation Rate [kg/(m s)]
2

4e-05

3e-05

2e-05

1e-05

0
0 2 4 6 8 10 12 14 16 18
Time [days]

Fig. 10. Variation of temperature. The highest computed evaporation rate is in the case of 303.15 K.

The evolution of temperature, saturation and mass fraction of vapour in the gas phase are very
similar to the ones already presented in Mosthaf et al. (2011). For completeness, the figures for the ref-
erence case are depicted in Figs 8 and 9. The vapour concentration in the ambient gas phase increases
along the interface of the two compartments as air flows over the evaporating porous surface. Conse-
quently, the porous medium desaturates. Due to the vaporization enthalpy, the porous medium is cooled
as long as two phases are present at the interface. As soon as the water phase disappears, the temperature
of the porous medium starts to adapt to the temperature in the free flow and the porous medium becomes
warmer again.
The evaporation rate is determined as integral of the fluxes of the component water across the inter-
face. This is evaluated as the sum of the residual of the mass balance for the water component at the
interface, thus in the same manner as the coupling is performed. Alternatively, the evaporation rate can
be calculated as temporal change of the amount of the component water that is stored in the porous-
medium system. This can be obtained by summing up the storage terms for all subcontrol volumes
in the porous-medium domain and determining the change of mass over time as Δmw /Δt, with mw
904 K. BABER ET AL.

being the total mass of the component water in the porous medium. The evaporation curves resulting
from the simulations (Fig. 10, reference case T = 293.15 K) show a similar qualitative behaviour as the
rates observed in physical experiments (Lehmann et al., 2008) with two distinct evaporation regimes.
A capillary-dominated regime, where the water phase at the interface makes the main contribution to
the evaporation process and capillary forces provide fresh supply from within the porous medium, is
followed by a diffusion-dominated regime, where the interface has dried out and water is predomi-
nantly transported in the form of vapour in the gas phase across the interface between free flow and
porous medium. The wavelike variations of the evaporation rate in the diffusion-dominated regime are
a numerical artefact. They are caused by the variable switch when cell rows dry out.

4.2 Variation of the temperature


Additionally to the reference case, temperatures of 283.15 and 303.15 K are tested. In case of a higher
temperature as in the reference case and as long as two fluid phases are present at the interface, the gas
phase in the porous medium contains more vapour. Thus, the difference of the vapour concentration
within the two subdomains becomes larger and leads to an increased evaporation rate, whereas in the
case of a lower temperature, the opposite happens. Figure 10 provides a comparison of the evaporation
rates with the different temperature levels. The highest rate is in the warm case, where the equilibrium
concentration of vapour in the gas phase in the porous medium is highest and thus direct evaporation
and the vapour transport is maximal. As a consequence, the interface dries out faster and the transition
from a capillary-dominated evaporation regime to a diffusion-dominated regime happens earlier.

4.3 Variation of the Beavers–Joseph coefficient


The influence of the Beavers–Joseph coefficient is examined by varying αBJ in the range of 0.1–5.0. We
apply the Beavers–Joseph condition only on the gas phase, whereas no-slip conditions are used for the
water and the solid phase in the porous medium. That is why a saturation-dependent Beavers–Joseph
coefficient was suggested in Mosthaf et al. (2011).

2.5e-05
alpha=0.1
alpha=1.0
alpha=5.0
2e-05
Evaporation Rate [kg/(m´†s)]

1.5e-05

1e-05

5e-06

0
0 2 4 6 8 10 12 14
Time [days]

Fig. 11. Computed evaporation rate for three different Beavers–Joseph coefficients.
COUPLING TWO-PHASE COMPOSITIONAL POROUS-MEDIUM FLOW 905

2.5e-05 2
K=1e-9 m2
K=1e-10 m2
K=1e-11 m
2
2e-05 K=1e-12 m

Evaporation Rate [kg/(m s)]


2
1.5e-05

1e-05

5e-06

0
0 2 4 6 8 10 12
Time [days]

Fig. 12. Evaporation rates for four different intrinsic permeabilities.

In Fig. 11, we present the evaporation curves for the two extreme cases and compare them to the
reference case (αBJ = 1.0). The slip velocity in the middle of the interface after 1 day is for the refer-
ence case 9.75e−5 m/s, for αBJ = 0.1 it is 9.0e−4 m/a and for αBJ = 5.0 it amounts to 1.8e−5 m/s. It is
influenced by the sinusoidal variation of the inflow velocity and by the fluxes normal to the interface,
because the Beavers–Joseph condition is implemented as Neumann-type condition. Although the slip
velocity is very sensitive to the choice of the BJ coefficient as can be expected from Equation (2.3), the
evaporation rate in the simulated cases is hardly influenced by this parameter. Thus, the precise value
and the saturation dependency of this parameter, which depends on many important properties such as
the surface roughness, may be of minor importance concerning the predicted evaporation rate. For our
quantities of interest, it is sufficient to use a physically reasonable choice of this parameter.

4.4 Variation of the permeability


Four different permeabilities have been tested: K = 1e−9 m2 , K = 1e−10 m2 (reference case),
K = 1e−11 m2 and K = 1e−12 m2 . The evaporation curves for these cases are depicted in Fig. 12.
With a permeability of K = 1e−9 m2 , the high evaporation rate which is mainly driven by capillary
forces can be sustained longest, whereas in the case of K = 1e−12 m2 , the resistance of the perme-
ability is so strong that the capillary forces cannot sustain the water demand at the interface and the
evaporation rate drops already after about 1 day. Please note that the permeability influences the dura-
tion of the capillary-driven evaporation regime, whereas the temperature affects also the magnitude of
the evaporation rate.
The variation of the parameters yields results that can be physically interpreted. Moreover, they
qualitatively match the behaviour of the rates measured in physical experiments (Lehmann et al., 2008).
We would like to emphasize that the evaporation rate is a direct outcome of our model and not imposed
as boundary condition as it was often done in existing models.
906 K. BABER ET AL.

2.5e-05
10x20 cells
20x40 cells
40x80 cells
2e-05

Evaporation Rate [kg/(m s)]


2
1.5e-05

1e-05

5e-06

0
0 1 2 3 4 5 6
Time [days]

Fig. 13. Effect of grid refinement on the computed evaporation rate.

4.5 Grid refinement


The reference case has been calculated using meshes with different numbers of cells in order to assess
the effect of the grid resolution on the modelled evaporation rate. This is shown in Fig. 13. As expected,
the very coarse grid with 10 × 20 cells is inaccurate. However, it can still represent the principal
behaviour of the evaporation process, but it shows the largest discrepancy at the transition from the
capillary- to the diffusion-dominated regime. The curves with a stronger refinement have less and less
differences to each other.

5. Summary and outlook


The numerical implementation of a coupled model describing interacting free and porous-media flow
is shown with a focus on the treatment of the interface conditions. The coupled model is based on
the two-domain approach, on a simple interface which cannot store mass, momentum or energy, and
on non-isothermal compositional submodels. These are coupled by interface conditions accounting for
mass, momentum and energy transfer and guaranteeing the continuity of fluxes.
The entire domain is discretized in the modelling framework DuMux using a vertex-centred finite-
volume (box) scheme and an implicit Euler time discretization. It is solved fully implicitly, relying
on the software tools DUNE-PDELab and DUNE-Multidomain for the assembly of the global system
matrix. The designation of subgrids and efficient handling of the interface are done using the module
DUNE-Multidomaingrid. The coupling conditions are implemented as a combination of Dirichlet and
flux conditions for the respective subdomains. The coupling fluxes are calculated indirectly via a flux
balance for the respective finite-volume box.
The presented coupling concept is applied in order to model evaporation processes influenced by
wind blowing over the porous surface. Numerical examples show the evolution of saturation and tem-
perature and illustrate the coupling concept. Furthermore, it is possible to study the influence of various
parameters on the interface processes. Exemplarily, the influence of the temperature, the permeabil-
ity and the Beavers–Joseph coefficient on the evaporation rate are shown. Thereby, the choice of the
COUPLING TWO-PHASE COMPOSITIONAL POROUS-MEDIUM FLOW 907

Beavers–Joseph coefficient turned out to have a minor influence on the evaporation rate across the
interface.
A comparison with evaporation experiments is envisaged in the close future. Besides, the model
concept and the numerical scheme have to be adapted to allow the description of turbulent gas flow
and radiation. In this context, it might be necessary to consider the use of more advanced discretization
schemes such as the DG method (Vassilev & Yotov, 2009; Rivière & Kanschat, 2010).
Moreover, the applicability to other scenarios, like flow in diffusion layers of PEM fuel cells or
transport processes between vascular and interstitial space in biological tissues, will be tested. This
requires a further development of the model to allow the resolution of the interface processes and
structure in more detail. Furthermore, the benefit of the implementation of a subdomain iteration
scheme needs to be investigated, especially because time and length scales may strongly differ in the
subdomains.
In order to improve the scalability of our implementation, we are currently in the process of adding
support for parallel computations based on domain decomposition of the grid. As we will still be solving
the fully coupled system, this domain decomposition will be independent of the subdomain structure.
Furthermore, solving the algebraic problem (3.7) might prove more challenging, as we will have to
investigate the use of iterative instead of direct solvers for the individual Newton steps.

Funding
This work was supported by the German research foundation (DFG). Some of the authors are mem-
bers of the DFG International Research Unit MUSIS (FOR 1083), the Cluster of Excellence SimTech
(EXC310) and of the International Research Training Group NUPUS (GRK 1398), all funded by the
German Research Foundation (DFG) and The Netherlands Organization for Scientific Research (NWO).
We thank the Royal Netherlands Academy of Arts and Sciences (KNAW) for facilitating the KNAW
colloquium in Amsterdam 2010, and the DFG for their support.

References
Alazmi, B. & Vafai, K. (2001) Analysis of fluid flow and heat transfer interfacial conditions between a porous
medium and a fluid layer. Int. J. Heat Mass Transfer, 44, 1735–1749.
Arbogast, T. & Brunson, D. (2007) A computational method for approximating a Darcy–Stokes system govern-
ing a vuggy porous medium. Comput. Geosci., 11, 207–218.
Bastian, P., Blatt, M., Dedner, A., Engwer, C., Klöfkorn, R., Kornhuber, R., Ohlberger, M. & Sander,
O. (2008) A generic grid interface for parallel and adaptive scientific computing. Part II: implementation and
tests in DUNE. Computing, 82, 121–138.
Bastian, P., Heimann, D. & Marnach, S. (2010) Generic implementation of finite element methods in the Dis-
tributed and Unified Numerics Environment (DUNE). Kybernetika, 46, 294–315.
Beavers, G. S. & Joseph, D. D. (1967) Boundary conditions at a naturally permeable wall. J. Fluid Mech., 30,
197–207.
Brinkman, H. (1947) A calculation of the viscous force exerted by a flowing fluid on a dense swarm of particles.
Appl. Sci. Res., A1, 27–34.
Brooks, R. & Corey, A. (1964) Hydraulic properties of porous media. Hydrology Paper, Fort Collins: Colorado
State University.
Cao, Y., Gunzburger, M., Hu, X., Hua, F., Wang, X. & Zhao, W. (2010) Finite element approximations for
Stokes–Darcy flow with Beavers–Joseph interface conditions. SIAM J. Numer. Anal., 47, 4239–4256.
Chandesris, M. & Jamet, D. (2006) Boundary conditions at a planar fluid-porous interface for a Poiseuille flow.
Int. J. Heat Mass Transfer, 49, 2137–2150.
908 K. BABER ET AL.

Chandesris, M. & Jamet, D. (2007) Boundary conditions at a fluid-porous interface: an a priori estimation of the
stress jump coefficients. Int. J. Heat Mass Transfer, 50, 3422–3436.
Chandesris, M. & Jamet, D. (2009) Jump conditions and surface-excess quantities at a fluid/porous interface: a
multi-scale approach. Transp. Porous Med., 78, 419–438.
Chidyagwai, P. & Riviére, B. (2010) Numerical modelling of coupled surface and subsurface flow systems. Adv.
Water Res., 33, 92–105.
Class, H., Helmig, R. & Bastian, P. (2002) Numerical simulation of nonisothermal multiphase multicomponent
processes in porous media - 1. An efficient solution technique. Adv. Water Res., 25, 533–550.
Demmel, J. W., Eisenstat, S. C., Gilbert, J. R., Li, X. S. & Liu, J. W. H. (1999) A supernodal approach to sparse
partial pivoting. SIAM J. Matrix Anal. Appl., 20, 720–755.
Discacciati, M. & Quarteroni A. (2004) Convergence analysis of a subdomain iterative method for the finite
element approximation of the coupling of Stokes and Darcy equations. Comput. Visual Sci., 6, 93–103.
Discacciati, M., Miglio, E. & Quarteroni A. (2002) Mathematical and numerical models for coupling surface
and groundwater flows. Appl. Num. Math., 43, 57–74.
Discacciati, M., Quarteroni, A. & Valli, A. (2007) Robin–Robin domain decomposition methods for the
Stokes–Darcy coupling. SIAM J. Numer. Anal., 45, 1246–1268.
Flemisch, B., Darcis, M., Erbertseder, K., Faigle, B., Lauser, A., Mosthaf, K., Müthing, S., Nuske, P.,
Tatomir, A., Wolff, M. & Helmig, R. (2011) DUMUX: DUNE for Multi-{Phase, Component, Scale, Physics,
. . .} flow and transport in porous media. Adv. Water Res., 34, 1102–1112.
Franca, L. P., Hughes, T. J. R. & Stenberg, R. (1993) Stabilized finite element methods for the Stokes problem.
In Incompressible Computational Fluid Dynamics (M. Gunzburger & R. A. Nicolaides, eds). Cambridge:
Cambridge University Press, pp. 87–108.
Girault, V. & Rivière, B. (2009) DG approximation of coupled Navier–Stokes and Darcy equations by Beaver–
Joseph–Saffman interface condition. SIAM J. Numer. Anal., 47, 2052–2089.
Goyeau, B., Lhuillier, D., Gobin, D. & Velarde, M. (2003) Momentum transport at a fluid–porous interface.
Int. J. Heat Mass Transfer, 46, 4071–4081.
Hassanizadeh, S. M. & Gray, W. G. (1989) Derivation of conditions describing transport across zones of reduced
dynamics within multiphase systems. Water Resour. Res., 25, 529–539.
Huber, R. & Helmig, R. (2000) Node-centered finite volume discretizations for the numerical simulation of mul-
tiphase flow in heterogeneous porous media. Comput. Geosci., 4, 141–164.
IAPWS (2009) Revised release on the IAPWS formulation 1995 for the thermodynamic properties of ordinary
water substance for general and scientific use. Available at http://www.iapws.org.
Jamet, D., Chandesris, M. & Goyeau, B. (2009) On the equivalence of the discontinuous one- and two-domain
approaches for the modeling of transport phenomena at a fluid/porous interface. Transp. Porous Media, 78,
403–418.
Jäger, W. & Mikelić, A. (2000) On the interface boundary condition of Beavers, Joseph and Saffman. SIAM J.
Appl. Math., 60, 1111–1127.
Jäger, W. & Mikelić, A. (2009) Modeling effective interface laws for transport phenomena between an unconfined
fluid and a porous medium using homogenization. Transp. Porous Media, 78, 489–508.
Jäger, W., Mikelić, A. & Neuss, N. (2001) Asymptotic analysis of the laminar viscous flow over a porous bed.
SIAM J. Sci. Comput., 22, 2006–2028.
Lauser, A., Hager, C., Helmig, R. & Wohlmuth, B. (2011) A new approach for phase transitions in miscible
multi-phase flow in porous media. Adv. Water Res., 34, 957–966.
Layton, W. J., Schieweck, F. & Yotov, I. (2002) Coupling fluid flow with porous media flow. SIAM J. Numer.
Anal., 40, 2195–2218.
Lehmann, P., Assouline, S. & Or, D. (2008) Characteristic lengths affecting evaporative drying of porous media.
Phys. Rev. E, 77, Paper ID. 056309.
Masson, C., Saabas, H. J. & Baliga, B. R. (1994) Co-located equal-order control-volume finite element method
for two-dimensional axisymmetric incompressible fluid flow. Int. J. Numer. Meth. Fluids, 18, 1–26.
Millington, R. & Quirk, J. (1960) Permeability of porous solids. Trans. Faraday Soc., 57, 1200–1207.
COUPLING TWO-PHASE COMPOSITIONAL POROUS-MEDIUM FLOW 909

Mosthaf, K., Baber, K., Flemisch, B., Helmig, R., Leijnse, A., Rybak, I. & Wohlmuth, B. (2011) A coupling
concept for two-phase compositional porous-medium and single-phase compositional free flow. Water Resour.
Res., 47, W10522.
Müthing, S. & Bastian, P. (2011) Dune-multidomaingrid: a metagrid approach to subdomain modeling. Advances
in Dune. Heidelberg: Springer.
Ochoa-Tapia, J. A. & Whitaker, S. (1995) Momentum transfer at the boundary between a porous medium and a
homogeneous fluid – I. Theoretical development. Int. J. Heat Mass Transfer, 38, 2635–2646.
Reid, R. C., Prausnitz, J. M. & Poling, B. E. (1987) The Properities of Gases & Liquids. New York: McGraw-
Hill, Inc.
Rivière, B. (2005) Analysis of a discontinuous finite element method for the coupled Stokes and Darcy problems.
J. Sci. Comput., 22, 479–500.
Rivière, B. & Kanschat, G. (2010) A strongly conservative finite element method for the coupling of Stokes and
Darcy flow. J. Comput. Phys., 229, 5933–5943.
Rivière, B. & Yotov, I. (2005) Locally conservative coupling of Stokes and Darcy flow. SIAM J. Numer. Anal.,
42, 1959–1977.
Rosenzweig, R. & Shavit, U. (2007) The laminar flow field at the interface of a Sierpinski carpet configuration.
Water Resour. Res., 43, W10402.
Saffman, R. (1971) On the boundary condition at the surface of a porous medium. Stud. Appl. Math., 50, 93–101.
Shavit, U. (2009) Special issue: transport phenomena at the interface between fluid and porous domains. Transp.
Porous Media, 78, 327–540.
Somerton, W., El-Shaarani, A. & Mobarak, S. (1974) High temperature behavior of rocks associated with
geothermal type reservoirs. Society of Petroleum Engineers, pp. SPE–4897, presented at 48th Annual Califor-
nia Regional Meeting of the Society of Petroleum Engineers.
Valdés-Parada, F. J., Alvarez-Ramirez, J. A., Goyeau, B. & Ochoa-Tapia, J. A. (2009a) Computation of jump
coefficient for momentum transfer between a porous medium and a fluid using a closed generalized transfer
equation. Transp. Porous Media, 78, 439–457.
Valdés-Parada, F. J., Alvarez-Ramirez, J. A., Goyeau, B. & Ochoa-Tapia, J. A. (2009b) Jump condition
for diffusive and convective mass transfer between a porous medium and a fluid involving adsorption and
chemical reaction. Transp. Porous Media, 78, 459–476.
Vassilev, D. & Yotov, I. (2009) Coupling Stokes–Darcy flow with transport. SIAM J. Sci. Comput., 31,
3661–3684.

Das könnte Ihnen auch gefallen