Você está na página 1de 19

Symmetry properties of macroscopic transport coefficients in porous media

D. Lasseux and F. J. Valdés-Parada

Citation: Physics of Fluids 29, 043303 (2017); doi: 10.1063/1.4979907


View online: http://dx.doi.org/10.1063/1.4979907
View Table of Contents: http://aip.scitation.org/toc/phf/29/4
Published by the American Institute of Physics
PHYSICS OF FLUIDS 29, 043303 (2017)

Symmetry properties of macroscopic transport coefficients


in porous media
D. Lasseux1,a) and F. J. Valdés-Parada2
1 CNRS, UMR 5295, University of Bordeaux, Esplanade des Arts et Métiers, 33405 Talence Cedex, France
2 Departamento de Ingenierı́a de Procesos e Hidráulica, Universidad Autónoma Metropolitana-Iztapalapa,
Ave. San Rafael Atlixco 186, 09340 Mexico D.F., Mexico
(Received 25 November 2016; accepted 26 March 2017; published online 17 April 2017)

We report on symmetry properties of tensorial effective transport coefficients characteristic of many


transport phenomena in porous systems at the macroscopic scale. The effective coefficients in the
macroscopic models (derived by upscaling (volume averaging) the governing equations at the under-
lying scale) are obtained from the solution of closure problems that allow passing the information
from the lower to the upper scale. The symmetry properties of the macroscopic coefficients are
identified from a formal analysis of the closure problems and this is illustrated for several differ-
ent physical mechanisms, namely, one-phase flow in homogeneous porous media involving inertial
effects, slip flow in the creeping regime, momentum transport in a fracture relying on the Reynolds
model including slip effects, single-phase flow in heterogeneous porous media embedding a porous
matrix and a clear fluid region, two-phase momentum transport in homogeneous porous media, as
well as dispersive heat and mass transport. The results from the analysis of these study cases are
summarized as follows. For inertial single-phase flow, the apparent permeability tensor is irreducibly
decomposed into its symmetric (viscous) and skew-symmetric (inertial) parts; for creeping slip-flow,
the apparent permeability tensor is not symmetric; for one-phase slightly compressible gas flow in
the slip regime within a fracture, the effective transmissivity tensor is symmetric, a result that remains
valid in the absence of slip; for creeping one-phase flow in heterogeneous media, the permeability
tensor is symmetric; for two-phase flow, we found the dominant permeability tensors to be symmetric,
whereas the coupling tensors do not exhibit any special symmetry property; finally for dispersive heat
transfer, the thermal conductivity tensors include a symmetric and a skew-symmetric part, the latter
being a consequence of convective transport only. A similar result is achieved for mass dispersion.
Beyond the physical mechanisms under consideration in the present work, the reported technique can
be viewed as a general methodology applicable to any type of upscaled model obtained by volume
averaging. Published by AIP Publishing. [http://dx.doi.org/10.1063/1.4979907]

I. INTRODUCTION sense, a tensorial transport coefficient relates the flux to the


force in the macroscopic model. When this coefficient is sym-
Modeling transport phenomena in porous media at the metric, it indicates that the flux in the ith-direction due to a
pore-scale is often not sufficient for applications as a macro- force in the jth-direction is exactly the same as the flux in
scopic description of the physical mechanisms through flux- the jth-direction due to the same force in the ith-direction,
to-force relationships is of more practical interest. Derivation whatever the microstructure of the medium, and this repre-
of macroscale models from the underlying physics at the pore sents fundamental information on the transport process. In
scale is hence of major importance, and this can be carried out other words, accounting for the symmetry properties of the
with the aid of an upscaling technique such as volume averag- coefficients involved in macroscale models is essential from
ing,1 homogenization,2 thermodynamically constrained aver- a mathematical viewpoint, especially in terms of the manipu-
aging theory,3 or stochastic approaches.4 For most of them, lations related to the macroscale flux terms. The importance
the idea consists in the derivation of an upscaled model that of the symmetry properties is also evident for the characteri-
involves macroscopic coefficients together with the appropri- zation of transport phenomena in multiscale systems from an
ate boundary-value problems allowing the prediction of these experimental or numerical point of view, as the knowledge
coefficients for a given microstructure and transport conditions of the number of independent components of the tensors that
taking place in the porous medium. To complete the descrip- need to be determined is of prime interest. Moreover, if an
tion and the understanding of the physics at the macroscopic effective-medium equation is taken as the departing model for
scale, knowledge of the symmetry properties of the macro- a subsequent upscaling process, knowledge of the symmetry
scopic transport coefficients is very important. In a general properties of the associated coefficients is crucial to analyze the
properties of the coefficients involved in the resulting larger-
scale model. A clear example of the above is the study of
a) Email: didier.lasseux@ensam.eu momentum transport in porous media. Since the derivation

1070-6631/2017/29(4)/043303/18/$30.00 29, 043303-1 Published by AIP Publishing.


043303-2 D. Lasseux and F. J. Valdés-Parada Phys. Fluids 29, 043303 (2017)

of the vector form of Darcy’s law involves a permeability are many other transport processes, apart from those cited
tensor, the properties of this tensor appear to be ambiguous above, for which, to the best of our knowledge, no results at all
when this macroscale model is applied to situations different were reported so far on the symmetry properties of the macro-
from steady and incompressible creeping flow in rigid and scopic coefficients present in the upscaled models. The case of
homogeneous porous media, where it is well-known that this single-phase incompressible flow through fractures featuring
tensor is symmetric as will be recalled below. Under inertial heterogeneous aperture fields, or through strongly heteroge-
flow conditions, there may be situations in which this ten- neous porous materials embedding porous matrix and clear
sor is quasi-symmetric or even skew-symmetric, depending fluid regions, together with heat dispersion are amongst the
on the value of the Reynolds number. If more than one phase principal ones.
is involved, the porous medium structure is heterogeneous, or The focus of the present work is hence laid upon an
the flow is not incompressible, the properties of the perme- efficient method dedicated to the analysis of the symme-
ability tensor become more obscure. Certainly, these doubts try properties of macroscopic coefficients arising in upscaled
are extensible to the dispersion tensor involved in heat or mass transport models within the framework of the volume aver-
transport in porous media. aging method (detailed in Appendix A). With this purpose
To be more precise, a quick review of the literature dealing in mind, a series of problems dealing with momentum, heat,
with the analysis of the symmetry of macroscopic transport and mass transport in porous media are analyzed. In all
coefficients shows that this property has been reported for cases, we limit our presentation to the pore-scale model, the
few transport mechanisms only. The intrinsic permeability ten- upscaled model, and the ancillary closure problems, so that
sor for one-phase incompressible momentum transport in the attention can be directed to the reformulation and symme-
creeping regime is probably the one for which symmetry has try properties analysis of the corresponding effective-medium
been studied most frequently, from energy considerations,5 coefficients.
double-scale homogenization,6–9 stochastic approach,10 or The paper is organized as follows. In Section II, study
volume averaging.1,11,12 When inertia is present, there is no cases related to momentum transport are considered, namely,
formal result available regarding the symmetry properties of inertial one-phase flow, slip flow in the creeping regime, flow
the apparent permeability tensor except some observation that in a fracture including slip effects, one-phase flow in a medium
it is not symmetric in general as obtained from the volume embedding a porous matrix and a clear fluid region, and finally
averaging procedure applied to the incompressible Navier- creeping two-phase flow. In Section III, dispersive heat and
Stokes equations11 or from numerical simulations.13,14 Non- mass transport in homogeneous porous media are subsequently
symmetry of the apparent permeability tensor for single-phase analyzed, highlighting the relevance of the symmetric and
slip-flow was only mentioned in the work by Skjetne and Auri- skew-symmetric parts of the associated macroscopic transport
ault 8 using homogenization. It was recently addressed in more coefficients. Finally, the corresponding conclusions are pre-
detail by Lasseux et al.12 by means of the volume averag- sented in Section IV. The derivations in all the study cases
ing method. In the case of two-phase flow in homogeneous require the use of tensor analysis; thus for the sake of brevity
porous media, symmetry of the effective (or dominant) perme- in presentation, a list of the most relevant tensor identities is
ability tensors was reported together with reciprocity relation- provided in Appendix B, together with some important recalls
ships for the coupling permeability tensors in accordance with on vector and tensor algebra necessary to recover them.
Onsager’s principle, both in the context of homogenization9,15
and volume averaging.16,17 The case of mass dispersion of a
tracer is strikingly a situation where confusion on the symme- II. MOMENTUM TRANSPORT
try properties of the dispersion tensor reported in the literature A. Inertial one-phase flow
has lasted for several decades. The dispersion tensor was some-
Consider the steady, incompressible, and Newtonian flow
times assumed (without any proof) to be non-symmetric in the
of a single fluid phase β saturating a rigid and homogeneous
general case.1,18–20 Some proof was provided while employ-
porous medium, for which the solid matrix is referred to
ing double-scale homogenization21 or volume averaging.22
as the σ-phase. The configuration is schematized in Fig. 1
Conversely, symmetry of this tensor is often implicitly pos-
where we have represented the macroscopic region VM of
tulated,23 demonstrated when determined from the method of
scale L and the averaging domain V of radius r 0 over which
moments,24–26 or simply considered as such because any non-
the pore-scale physical model is averaged and where the β-
symmetric component in this tensor would be unimportant in
phase and σ-phase can be identified. The governing equations
the estimation of the dispersive flux in the macroscopic dis-
for total mass and momentum transport at the pore-scale are
persion equation.1,18,20 It is sometimes invoked on the basis
given by
of Onsager’s reciprocity arguments.27,28 Onsager’s arguments
have been widely employed in the upscaling process from the
∇ · vβ = 0, in Vβ , (1a)
molecular scale to the continuum scale to justify reciprocity of
the coupling at the continuum scale.29,30 However, they do not ρvβ · ∇vβ = −∇pβ + µ∇ vβ , 2
in Vβ , (1b)
necessarily apply for subsequent upscaling procedures leading vβ = 0, at Aβσ . (1c)
to macroscopic transport models in porous media.21,31 In all
cases mentioned above, it must be noted that the symmetry As indicated by Eq. (1c), the fluid velocity is subject to
analysis of macroscopic tensorial transport coefficients does the nonslip boundary condition at the solid-fluid interface
not rely on a well-defined and systematic methodology. There Aβσ . Throughout this work, ρ and µ are used to denote
043303-3 D. Lasseux and F. J. Valdés-Parada Phys. Fluids 29, 043303 (2017)

is a reasonable solution domain for the above closure problem


and thus the periodic boundary condition given in Eq. (5d). In
this problem, the vector e is another closure variable that maps
hvβ iβ onto the spatial variations of the fluid pressure, p̃β . Note
that ∇E is a third-order tensor which is defined in Eq. (B3)
of Appendix B. This closure problem has been numerically
solved by Lasseux et al.14 on unit cells of model structures
for several Reynolds number values and incident flow angle
orientations.
Our objective in this section is to analyze the structure of
the apparent permeability tensor H. From Eq. (4), it follows
that, in order to gain more knowledge about this coefficient, it
is necessary to have more insight into the closure variable E.
With this in mind, let us pre-multiply Eq. (5b) by ET and take
into account the solenoidal nature of the fluid velocity fields,
in order to obtain
ρ T  
E · ∇ · vβ E = −ET · ∇e + ET · ∇2 E + ET . (6)
FIG. 1. Sketch of a system, consisting of a porous medium saturated with µ
a single fluid phase including characteristic lengths and a sample of the
averaging domain. Applying the superficial average operator to this equation leads
to
the density and dynamic viscosity of the β-phase for single ρD T  E D E D E
E · ∇ · vβ E = − ET · ∇e + ET · ∇2 E + HT , (7)
phase-flow. µ
As detailed by Whitaker,11 the resulting upscaled model
where, for the last term of the above equation, we used the
arising from the volume averaging method consists of the
definition given in Eq. (4). Moreover, employing the identity
macroscale continuity equation,
in Eq. (B18) of Appendix B, taking into account Eq. (5a),
∇ · hvβ i β = 0 (2a) and after using the spatial averaging theorem in its divergence
form as indicated in Eq. (A2) of Appendix A, the first term
and the following Darcy-like equation: on the right-hand side (rhs) of the above equation takes the
H form
hvβ i = − · ∇hpβ i β . (2b) 
µ D T E

1
E · ∇e = ∇ · Ee = ∇ · Ee +

n · EedA, (8)
Here the superficial averaged velocity and pressure are defined V
Aβ σ
as (see Appendix A)
  where n denotes the unit normal vector to the Aβσ inter-
1 β 1
hvβ i = vβ dV , hpβ i = pβ dV . (3) face directed from the β- to the σ-phase (see Fig. 1). Note
V Vβ also that Ee is a third-order tensor resulting from the outer
Vβ Vβ
product defined in Eq. (B11) of Appendix B. When the inter-
β
Note that hvβ i = εhvβ i , with ε being the porous medium facial boundary condition given by Eq. (5c) is considered and
porosity. In addition, the apparent permeability tensor H is while limiting the analysis to the bulk of the porous medium
defined as where
D T average
E quantities are position-invariant, it results that
H = hEi. (4) E · ∇e = 0 and Eq. (7) can be written as
ρD T  E D E
In the above expression, the second-order tensor E is a clo- E · ∇ · vβ E = ET · ∇2 E + HT . (9)
sure variable that maps hvβ i β onto the spatial variations of µ
the fluid velocity, ṽβ , and it satisfies the following boundary- Directing the attention to the first term on the rhs of this equa-
value problem (see the work of Whitaker 11 for details in the tion, and using successively the identities reported in Appendix
derivation): B in Eqs. (B21), (B12), and (B22) (taking A to be ET and B
to be E) and Eq. (B16), we have
∇ · E = 0, in Vβ , (5a) ET
ρ
D T E D
E · ∇2 E = ∇2 ET · E
vβ · ∇E = −∇e + ∇2 E + I, in Vβ , (5b)
µ D  ET D T1 ET
E = 0, at Aβσ , (5c) = ∇ · ∇ET · E − (∇ET ) : (∇E)T 1
ET D ET
ψ(r) = ψ(r + li ), ψ = E, e,
D 
(5d) = ∇ · ∇ET · E − (∇E)T 3 : ∇E . (10)
β
hei = 0. (5e)
In this result, one must keep in mind that ∇E (or ∇ET ) is
At this point, it is worth mentioning that the developments a third-order tensor. Since the transposition operation is not
presented this far involve adopting a set of length-scale con- unique for this type of tensors, we have used the superscript
straints and assumptions as explained in item 5 of Appendix T 1 to denote the transpose that permutes the first and second
A. One of these scaling postulates is that a periodic unit cell indices as defined in Eq. (B8) of Appendix B, i.e. (C is a
043303-4 D. Lasseux and F. J. Valdés-Parada Phys. Fluids 29, 043303 (2017)

 
third-order tensor), CT 1 = Cjik , while the superscript T 3 this conclusion should be carefully considered because it is
ijk
denotes the transpose that permutes the first and only applicable whenever it is safe to impose nonslip condi-
 third
 indices
as defined in Eq. (B9) of Appendix B, i.e., CT 3 = Ckji . tions at the solid-fluid interface. To prove this point, consider
ijk the steady, non-inertial, Newtonian, and slightly compress-
In addition, throughout this work, the nested convention is
ible flow of a barotropic fluid saturating the pores of a rigid
adopted for double inner products (see Eqs. (B5) and (B7) in
and homogeneous porous medium. The situation is again the
Appendix B).
one schematically represented in Fig. 1. Under isothermal
Applying the spatial averaging theorem to the first term on
conditions, the governing total mass, momentum, and state
the rhs of the last above equation and using similar arguments
equations at the pore-scale are
as those used to simplify Eq. (8) (i.e., the nonslip-like boundary
condition in Eq. (5c) and the fact that average quantities are ∂ρ
+ ∇ · (ρvβ ) = 0, in Vβ , (16a)
considered as constants at the closure level), it is not hard to ∂t
conclude that this term is zero. Consequently, Eq. (9) can now
be written as 0 = −∇pβ + µ∇2 vβ , in Vβ , (16b)
D E ρD T  E
HT = (∇E)T 3 : ∇E + E · ∇ · vβ E . (11)
µ
 
ρ = f pβ , in Vβ . (16c)
The first term on the rhs of this result is clearly symmetric. As a
At the solid-fluid interface, the following slip-condition is
matter of fact, under creeping-flow conditions, this is the only
applicable:12,32–35
term remaining in the definition of the intrinsic permeability
2 − σ3
!
tensor and it is the proof that this tensor is symmetric7 (see     T 
vβ = − λ β I − nn · n · ∇vβ + ∇vβ ,
also Problems 4 and 5 in the work of Whitaker 1 ). Therefore, σ3
the rest of this section is dedicated to the analysis of the second
| {z }
ξ
term on the rhs of Eq. (11) (which is originated from inertial
(16d)
effects at the pore-scale).
Let us commence the analysis by noting that, due to the with σ3 and λ β being the accommodation coefficient and the
continuity equation, the last term in Eq. (11) may also be mean free path, respectively. Upscaling of this problem with
written as (see Eq. (B19)) the use of the volume averaging method was addressed ear-
D T  E D E lier by Lasseux et al.12 where the following set of macroscale
E · ∇ · vβ E = ET vβ : ∇E . (12)
equations was obtained:
At this point, it is convenient to note that
∂hρiβ
+ ∇ · hρiβ hvβ iβ = 0,
 
D T  ET D   E (17a)
E · ∇ · vβ E = ∇ · vβ ET · E ∂t
D  E D E Ks
= ∇ · vβ ET · E − ET vβ : ∇E . (13) hvβ i = − · ∇hpβ iβ , (17b)
µ
In order to obtain the last equality, we took into account Eq.
hρiβ = f hpβ iβ ,
 
(B20). As shown above, the use of the spatial averaging theo- (17c)
rem together with the interfacial boundary condition leads to where Ks is an apparent slip-corrected permeability tensor,
conclude that the first term on the rhs of the above equation which can be computed from
is null. In this way, substitution of Eq. (12) on the remaining
term on the rhs of Eq. (13) yields Ks = hDi. (18)
D T  E D 
E · ∇ · vβ E = − ET · ∇ · vβ E .
ET
(14) Here, the closure variable D is obtained from the solution of
the following boundary-value problem:
This is the proof that the second term in the decomposition of
the apparent permeability tensor, given in Eq. (11), is skew- ∇ · D = 0, in Vβ , (19a)
symmetric and the tensor H can finally be written as 0 = −∇d + ∇ D + I, in Vβ ,
2
(19b)
E ρD T f  g
D = −ξ λ β (I − nn)· n· ∇D + (∇D)T 1 , at Aβσ ,
D  E
H = (∇E)T 3 : ∇E − E · ∇ · vβ E . (15) (19c)
} |µ
ψ(r) = ψ(r + li ); ψ = D, d,
| {z
Symmetric part
{z } (19d)
Skew-symmetric part
β
In this way, it can be deduced that the first term on the rhs of this hdi = 0, (19e)
last expression, which results from dissipative-like transport, is where, in the boundary condition in Eq. (19c), the superscript
the symmetric part of the apparent permeability tensor and that T 1 denotes the transpose of a third-order tensor as defined in
the second term, arising from convective transport, is its skew- Eq. (B8) in Appendix B. The definitions of the closure vari-
symmetric part. In other words, the decomposition shown in ables D and d are analogous to those of E and e presented
Eq. (15) readily provides the irreducible parts of the tensor H. above, respectively. In addition, λ β represents the mean free
path at the intrinsic-average density and it can thus be regarded
B. Slip flow
as a constant at the closure level.35
From the results presented in the previous paragraphs, In Appendix B of the work by Lasseux et al.,12 an analysis
it may be inferred that the permeability tensor remains sym- similar to the one used in Sec. II A for the properties of the
metric for creeping flow, whatever the conditions. However, apparent permeability tensor for inertial momentum transport
043303-5 D. Lasseux and F. J. Valdés-Parada Phys. Fluids 29, 043303 (2017)

is provided. Here, only the final reformulation of the appar- last term in Eq. (20) is only symmetric if the solid-fluid inter-
ent slip-corrected permeability reported in this reference is face exhibits symmetries across the three planes parallel to the
recalled. It is given by edges of the periodic unit cell and passing through its centroid.
E 1  Furthermore, after performing an order of magnitude analy-
sis, Lasseux et al.12 concluded that, for conditions in which
D T3
Ks = (∇D) : ∇D − (∇D)T 3 : nD dA. (20)

V ξKn  ε (with Kn = λ β /` β being the macroscale Knudsen
Aβ σ
number), Ks can be considered to be quasi-symmetric. This
Just like for flow with inertia, the first term on the rhs of this analysis thus shows that, under creeping slip-flow conditions,
expression results from the dissipation-like term in Eq. (19b) it cannot be concluded that the apparent permeability tensor
and it is not difficult to demonstrate that it is symmetric. The in Darcy’s law is a symmetric tensor.
remaining term is, in general, non-symmetric. This observa- Finally, using the irreducible decomposition of any
tion is consistent with the study by Skjetne and Auriault 8 second-order tensor and the identity given in Eq. (B17),
using the homogenization method. As a matter of fact, the Eq. (20) may be expressed as
 f
D T3
E 1 g
Ks = (∇D) : ∇D − (∇D)T 3 : nD + ∇D : (nD)T 3 dA

2V
Aβ σ
| {z }
Symmetric part
 f
1 g
(∇D)T 3 : nD − ∇D : (nD)T 3 dA,

− (21)
2V
Aβ σ
| {z }
Skew-symmetric part

where the symmetric and skew-symmetric terms are clearly for the leak-rate determination through mechanical seals36–39
identified. The existence of the skew-symmetric term may or for the lubrication of rough surfaces.40 The boundary-value
contribute to a misalignment of the velocity with the macro- problem governing flow within the fracture is the one provided
scopic pressure gradient as it is the case, in general, for inertial by Equation (16), except a steady version of the mass conserva-
flow. tion Equation (16a) is considered. When the local slope of the
fracture walls’ roughness (i.e., tan α, see Fig. 2(a)) remains
C. Slightly compressible slip-flow in a fracture small compared to unity, the form of the above-mentioned
mass and momentum conservation equations, pre-integrated
The conclusion reached in Sec. II B regarding slip flow in
in the direction of the thickness of the fracture, can be easily
the bulk of a porous medium raises the question on whether
obtained, reducing the dimension of the problem from three to
it remains valid in the case of slip-flow within the confined
two:
space between two surfaces, i.e., in a fracture. The purpose of
the present section is hence to investigate the properties of the ∇ · qβ = 0, in AβM , (22a)
effective-medium coefficient present in the upscaled model k
for the steady Newtonian flow of a fluid ( β-phase) saturat- qβ = −ρ ∇pβ , in AβM , (22b)
µ
ing a fracture. The study is directed to slightly compressible  
slip-flow under isothermal conditions. The fracture, of local ρ = f pβ , in AβM , (22c)
aperture h, consists of two impermeable rough walls (σ-phase) qβ · n = 0, at Cβσ , (22d)
as sketched in Fig. 2(a). Contact spots may also occur between
the upper and lower surfaces of the fracture as it would be the where qβ represents the mass flow rate per unit length of the
case, for instance, for the contact under normal stress between fracture in each direction, while k stands for the so-called local
two manufactured solid surfaces (see Fig. 2(a)). When this con- transmissivity of the fracture aperture field that includes slip
tact exists, their contours shall be denoted by Cβσ in the mean effects, i.e.,
ξ λβ
!
plane of the fracture (see Fig. 2(b)). The incompressible flow 1 3
k= h 1+6 . (23)
version of this problem has been the object of many studies 12 h

FIG. 2. (a) Sketch of a fracture between two rough sur-


faces for which the aperture field is h. The local slope
is such that tan α  1 everywhere. (b) Top view of the
fracture with contact areas which contours are C β σ and
averaging surface A.
043303-6 D. Lasseux and F. J. Valdés-Parada Phys. Fluids 29, 043303 (2017)

Moreover, in this last equation, as in Sec. II B, λ β denotes the maps ∇hpβ isβ onto the spatial variations of the fluid pressure
mean-free path of the fluid molecules at the pressure and tem- and is the solution of the following boundary-value problem:
perature under consideration. In addition, in Eqs. (22), AβM  
∇ · k ∇b + I = 0, in Aβ , (30a)
represents the surface where no effective contact takes place
and n is the unit normal vector at Cβσ directed from the β-phase −n · ∇b + I = 0, at Cβσ ,

(30b)
to the σ-phase. β
hbis = 0, (30c)
Equations (22), which basically correspond to the
Reynolds model including slip, are still operating at the scale of b(r + li ) = b(r), i = 1, 2. (30d)
roughnesses, and, from a practical point of view, it is useful to The interest shall now be dedicated to the analysis of the
provide a description of the flow over a representative elemen- symmetry properties of the transmissivity tensor K. With this
tary surface A in which the portion that excludes the effective purpose in mind, the outer product of Eq. (30a) with b may
contact area is denoted by Aβ . To this goal, an upscaling can be formed, and while taking the superficial area average of the
be carried out using the following definitions of the averages, result, one obtains
analog to the ones employed for volume averaging, namely, 
 1 
the superficial area averaging operator ∇ · hk ∇b + I bis + kn · ∇b + I bd`
 A
1 Cβ σ
hψβ is = ψβ dA (24a)  
A − hk ∇bT + I · ∇bis = 0. (31)

To arrive at this expression, the identity given in Eq. (B18) of
and intrinsic averaging operator
 Appendix B was employed as well as the averaging theorem
β 1 in Eq. (25).
hψβ is = ψβ dA, (24b)
Aβ Because of periodicity and due to the slightly compress-


ible flow assumption, the average hk ∇b + I bis can be treated
where A and Aβ represent the areas of A and Aβ , respectively. as a constant at the closure level so that the first term on the
Similarly, the surface version of the spatial averaging theorem left-hand side of Eq. (31) is zero. The second term vanishes
takes the form also when the boundary condition given in Eq. (30b) is taken
 into account and this yields
1
h∇ψβ is = ∇hψβ is + nψβ d`. (25) β β
A hk∇bis = −hk∇bT · ∇bis (32)
Cβ σ
where we have employed the intrinsic average instead of the
When the superficial average of the problem in Eqs. (22) is car- superficial average. When this last result is introduced back
ried out, it can be shown that the steady macroscopic Reynolds into Eq. (28), the effective transmissivity tensor takes the form
model for slightly compressible slip flow may be written as
β β
K = hkIis − hk∇bT · ∇bis (33)
∇ · hqβ is = 0, (26a)
and this final result proves that K is a symmetric tensor.
β K As important remarks following the development per-
hqβ is = −hρis ε s · ∇hpβ isβ , (26b) formed in this section, it should be noted that:
µ
• The expression for the effective transmissivity tensor,
β
hρis = f hpβ isβ
 
(26c) given in Eq. (33), remains completely valid if the flow
takes place without slip effects and/or if the flow is
A
in which ε s = Aβ = 1 − δs , where δs stands for the so-called perfectly incompressible.
load-bearing capacity. At this point, it is opportune to clarify • More generally, the conclusions may be extended to the
that the slightly compressible hypothesis is supported by the effective diffusivity tensor resulting from the upscaling
constraint of a diffusion process within a heterogeneous medium
β β when the local diffusivity can be considered as a con-
ρ̃ = ρ − hρis  hρis . (27)
tinuous function of space. In particular, it is readily
In the expression of the macroscale flow-rate in Eq. (26b), K is applicable to the permeability tensor resulting from the
the effective transmissivity tensor characterizing the fracture large-scale averaging process carried out on the one-
that is given by phase Darcy flow in a heterogeneous porous medium
 β when the permeability field is considered as a contin-
K = hk ∇b + I is , (28) uous function of space as reported by Quintard and
Whitaker.41
where k is defined as
ξ λ βs + D. One-phase flow in a medium embedding
1 3*
k= h 1+6 , (29) a porous matrix and a clear fluid region
12 , h -
Up to this point of the investigation, momentum trans-
β
λ βs being now the mean-free path at the average density hρis . port has not been considered in a heterogeneous porous
In the expression of K above, b is the closure variable that medium featuring discontinuous properties and the question
043303-7 D. Lasseux and F. J. Valdés-Parada Phys. Fluids 29, 043303 (2017)

now remains on whether the conclusions reached so far are K∗  ∗ 


Vβ∗ = − · ∇Pβ − ρg , (35b)
still applicable in this situation. To illustrate this, we are hence µ
interested to study one-phase creeping flow of a fluid-phase β where Vβ∗ and Pβ∗ represent the megascale average velocity
within a heterogeneous system in which a porous matrix (ω- and pressure defined as
region) and a clear fluid (η-region) can be distinctly identified 
as for example in a vugular (see Fig. 3) or a fractured porous 1 η
Ψβ = {Ψβ } =

Ψβ dV = {Ψβω }ω + {Ψβ }η
medium. This problem has been analyzed by Arbogast and V
V
Lehr 42 and more recently by Golfier et al.43 The macroscale  
1 1 η
model used in the latter work to describe the incompressible = Ψβω dV + Ψβ dV , (36)
Newtonian flow in such a material is the result of a first up- V V
Vω Vη
scaling of the continuity and Stokes equations at the microscale
and it is given by where Vκ (κ = ω, η) represents the portion of the κ-region
contained within the averaging volume V = Vω + Vη . In the
∇ · Vβω = 0 in the ω-region, (34a)
megascale momentum Equation (35b), K∗ represents the effec-
tive permeability tensor that can be computed from the solution
µ 2 ω
0 = −∇Pβω + ρg + ∇ Vβ − µK−1 ω
ω · Vβ in the ω-region, of the associated closure problem given by
ε
(34b) ∇ · Dω = 0, in Vω , (37a)
η
Vβω = Vβ at Aωη , (34c) 1
0 = −∇dω + ∇2 Dω − K−1
ω · Dω + I, in Vω , (37b)
ε
η η
Pβ − Pβω I + µ ε −1 ∇Vβω − ∇Vβ
f  f 

  B.C. 1 Dω = D η at Aωη , (37c)
η T
+ ε −1 ∇Vβω − ∇Vβ

= 0 at Aωη , (34d)
" #
1 T1
η B.C. 2 n · −Idω + ∇Dω + ∇Dω
∇ · Vβ = 0 in the η-region, (34e) ε
   T1  
η η = n · −Idη + ∇Dη + ∇Dη at Aωη , (37d)
0 = −∇Pβ + ρg + µ∇2 Vβ in the η-region. (34f)
In these equations, Vβκ and Pβκ (κ = ω, η) are the macroscale
∇ · Dη = 0, in Vη , (37e)
velocity and pressure in the κ-region, ε and Kω are the poros-
ity and intrinsic permeability tensor of the ω-porous region. In
the boundary condition given in Eq. (34d), n denotes the unit 0 = −∇dη + ∇2 Dη + I, in Vη , (37f)
normal vector at the ω – η boundary Aωη , directed from the ω-
region toward the η-region (i.e., nωη as shown in Fig. 3). After d∗ = 0, (37g)
applying the averaging procedure to the above equations,43
the megascale model operating on the effective medium D∗ = K ∗ , (37h)
reads
∇ · Vβ∗ = 0, (35a) ψk (r + li ) = ψk (r), ψ = d, D, k = ω, η, i = 1, 2, 3. (37i)
In these equations, Dκ and dκ (κ = ω, η) are closure
variables that map Vβ∗ onto the velocity and pressure varia-
tions in the κ-region, respectively, whereas D∗ and d∗ are the
megascale averages based on the definition given in Eq. (36).
The analysis of the symmetry properties of the effec-
tive permeability tensor K∗ starts by first reformulating the
momentum-like Equations (37b) and (37f), which, while tak-
ing into account the fact that Dκ , (κ = ω, η) are solenoidal
tensor fields as indicated by Eqs. (37a) and (37e), may indeed
be written as
1   
0 = −∇dω + ∇ · ∇Dω + ∇Dω T 1 − K−1 ω · Dω + I (38)
ε
and
1   T1
0 = −∇dη + ∇ · ∇Dη + ∇Dη + I. (39)
ε
When the former of these two equations is pre-multiplied
FIG. 3. Sketch of a megascopic vugular medium saturated with a single fluid
phase that contains a homogeneous porous medium (ω-region) and clear fluid by DTω and when the ω-regional average of the result is formed,
zones (η-region). the following expression is obtained:
043303-8 D. Lasseux and F. J. Valdés-Parada Phys. Fluids 29, 043303 (2017)

1 11 transport coefficients at the macroscale is extremely impor-
0= DTω · ndω dA +
V εV tant. Clearly, the first two terms in Eq. (43) are also symmetric
Aω η
  T as shown in Appendix B (see Eqs. (B25) and (B26)) making
the effective permeability tensor K∗ a symmetric tensor and
  T3
× n · ∇DTω + ∇DTω · Dω dA
this represents the final important result of this section.
Aω η
(
1  T  T T3T1
)T E. Creeping two-phase flow in homogeneous
: ∇Dω T 1

− ∇Dω + ∇Dω porous media
ε ω
( T
− Dω · K−1 · D
)
+
( T)
D . (40) So far, the analysis of macroscopic coefficients has been
ω ω ω
ω ω restricted to single phase-flow in porous structures. As a final
This last result is subject to the use of the averaging theorem momentum transport case, let us then consider incompressible,
for the ω-regional average, which has exactly the same form immiscible, and Newtonian two-phase momentum transport in
as the one given in Eq. (A2), together with the fact that dω the bulk region of a rigid porous medium as sketched in Fig. 4.
and Dω are periodic. Moreover, the first term on the rhs of the The governing pore-scale equations and interfacial boundary
above equation results from the fact that Dω is solenoidal. The conditions for this transport process are given by
second term on the rhs of Eq. (40) results from the use of the
∇ · vk = 0, in the k-phase (k = β, γ), (44a)
identities given in Eqs. (B12) and (B14) in which the arbitrary
twice-derivable second-order tensor A is identified as Dω . It
also makes use of the identity given in Eq. (B22), in which ∇A 0 = −∇pk + µk ∇2 vk , in the k-phase(k = β, γ), (44b)
 T3
is identified as ∇DTω + ∇DTω and B to Dω .
  T B.C. 1 vk = 0, at Akσ (k = β, γ), (44c)
At this point, it is worth noticing that n · ∇DTω
T3T B.C. 2 vβ = vγ , at Aβγ , (44d)
 
= n · ∇Dω and n · ∇DTω = n · ∇Dω T 1 , so that taking


into account Eqs. (B15) and (B16), Eq. (40) can be rewritten B.C. 3 nβγ · Tβ = nβγ · Tγ + 2σHnβγ , at Aβγ . (44e)
as
 " # Here, the nonslip condition has been imposed for both fluid
1 T 1 T1
0= Dω · n · −Idω + ∇Dω + ∇Dω dA phases in contact with the solid phase. In addition, assum-
V ε ing that no mass transport is taking place between the fluid
Aω η
phases, it is reasonable to impose continuity of the velocity
1( f  g )T
∇Dω T 3 : ∇Dω + ∇Dω T 1 fields at Aβγ . Moreover, at this interface, the normal stress


(ε T
ω
−1
) ( T) is exactly compensated by the Laplace pressure as expressed
− Dω · K ω · Dω + Dω . (41)
ω ω in Eq. (44e) fwhere the totalg stress tensor is denoted as Tk
Let now attention be directed toward the η-region. Using = −Ipk + µk ∇vk + (∇vk )T (k = β, γ), while σ and H rep-
the same procedure employed above, it follows that resent the interfacial tension and the mean curvature of Aβγ ,
 respectively. This problem has been studied by Whitaker 16,44
1 and by Lasseux et al.17 using the volume averaging method.
    T 1 
0=− DTη · n · −Idη + ∇Dη + ∇Dη dA
V The resulting upscaled model for momentum transport can be
Aη η
expressed as follows:
 T3   T 1  T ( T )
K∗ββ K∗βγ

− ∇Dη : ∇Dη + ∇Dη + Dη . (42)
η η hvβ i = − · ∇hpβ iβ − · ∇hpγ iγ , (45a)
µβ µγ
When Eqs. (41) and (42) are added and when the boundary con-
ditions B.C.1 and B.C.2 in Eqs. (37c) and (37d) are employed ∗
Kγγ ∗
Kγβ
together with the definition of the effective permeability tensor hvγ i = − · ∇hpγ iγ − · ∇hpβ iβ . (45b)
in Eq. (37h), the following expression for K∗ arises: µγ µβ

1( f  g)
K∗ = ∇Dω T 3 : ∇Dω + ∇Dω T 1

ε ω
 T3   T1 
+ ∇Dη : ∇Dη + ∇Dη
η
( T −1
)
+ Dω · K ω · Dω . (43)
ω

In this last expression, we have made( use of the fact) that Kω


is a symmetric tensor to infer that DTω · K−1 ω · D ω ω is also
symmetric. To reach this conclusion, the symmetry property
of Kω , which is known in the creeping flow regime with the
no-slip boundary condition at the fluid-solid interface, as men-
tioned earlier, is indeed fundamental. This is a clear example FIG. 4. Sketch of a porous medium saturated with two immiscible fluid
showing why the knowledge of the symmetry properties of phases.
043303-9 D. Lasseux and F. J. Valdés-Parada Phys. Fluids 29, 043303 (2017)

This model was also derived using the homogenization method These closure problems are restricted to cases in which the
by Auriault.15 In Eqs. (45), the permeability tensors K∗αk surface tension has no effects at the closure problem level.15,44
(α, k = β, σ) are defined in terms of the corresponding closure This is a reasonable assumption whenever the capillary number
variables as is much smaller than unity; i.e., Ca = µh3i/σ  1, with µh3i
being the largest of O(µβ hvβ iβ ) or O(µγ hvγ iγ ).
K∗αk = −hDαk i, α, k = β, γ. (46)
To commence the derivations, let attention be focused to
Following the work of Lasseux et al.,17 we shall refer to the Problem I, in specific, consider Eq. (47b) for k = β, pre-
tensors K∗ββ and Kγγ∗
as the dominant permeability tensors multiply it by DTββ , and apply the superficial average defined

and to the tensors K∗βγ and Kγβ as the coupling permeability in Eq. (A1a) of Appendix A on the resulting equation in order
tensors. The closure variables Dββ and Dγβ are the solution of to obtain
the following boundary-value problem. D E     T 1 
− DTββ · ∇dββ + DTββ · ∇ · ∇Dββ + ∇Dββ
β β
1. Problem I
+ K∗T
ββ = 0. (49)
∇ · Dkβ = 0, in Vk (k = β, γ), (47a) Here, we have used the subscript β in the superficial average to
  T1 clearly indicate that this average is taken only in the β-phase.
−∇dkβ + ∇ · ∇Dkβ + ∇Dkβ = δkβ I, in Vk (k = β, γ), At this point, one may follow the same line of derivations
performed in Section II D in order to obtain
(47b) 
1     T 1 
0= DTββ · nβγ · −Idββ + ∇Dββ + ∇Dββ dA
B.C. 1 Dkβ = 0, at Akσ (k = β, γ), (47c) V
A βγ
 T 1 T
Dββ = Dγβ ,
 T3 
B.C. 2 at Aβγ , (47d)

− ∇Dββ : ∇Dββ + ∇Dββ + K∗T
ββ . (50a)
β
 T1
Let us return the attention to Eq. (47b) but now for k = γ with

B.C. 3 µβ nβγ · −Idββ + ∇Dββ + ∇Dββ T
  T1 a pre-multiplication by Dγβ ; performing the same procedure
= µγ nβγ · −Idγβ + ∇Dγβ + ∇Dγβ , at Aβγ , (47e) used just above, it is not hard to deduce that

1 T
   T1 
0=− Dγβ · nβγ · −Idγβ + ∇Dγβ + ∇Dγβ dA
ψkβ (r + li ) = ψkβ (r), i = 1, 2, 3; ψ = d, D; k = β, γ, V
A βγ
(47f)  T3    T 1 T
− ∇Dγβ : ∇Dγβ + ∇Dγβ . (50b)
hdkβ ik = 0, k = β, γ. (47g) γ

When this last result is multiplied by µγ and it is added to Eq.


In Eq. (47b), δkβ is the Dirac delta function. Note that in this
(50a) multiplied by µβ , the resulting equation takes the form
equation, we have taken into account the solenoidal  nature
 µγ   T3   T 1 
of the tensors Dkβ (k = β, γ), so that ∇ ∇ · Dkβ = ∇ ·

K∗ββ = ∇Dγβ : ∇Dγβ + ∇Dγβ
 T1 µβ γ
∇Dkβ = 0. In a similar manner, the closure variables Dβγ  T3    T 1 
and Dγγ must satisfy the following closure problem. + ∇Dββ : ∇Dββ + ∇Dββ . (51)
β

2. Problem II Here, the interfacial boundary conditions given in Eqs. (47d)


and (47e) were taken into account. The two average terms
∇ · Dkγ = 0, in Vk (k = β, γ), (48a) on the rhs of this equation can be easily shown to be sym-
metric (see the proof in Appendix B, Eqs. (B25) and (B26)),
thus making K∗ββ a symmetric tensor. This conclusion is con-
  T1
−∇dkγ + ∇ · ∇Dkγ + ∇Dkγ = δkγ I, in Vk (k = β, γ),
sistent with those obtained separately by Auriault9,15 and
(48b) Whitaker.16
The above procedure can be repeated with Eq. (47b) (for
B.C. 1 Dkγ = 0, at Akσ (k = β, γ), (48c) k = β) when it is pre-multiplied by DTβγ and also with this
same equation, when evaluated for k = γ and pre-multiplied
B.C. 2 Dβγ = Dγγ , at Aβγ , (48d) T
by Dγγ to obtain, after few steps, which are analogues to those
  T1 used above,
B.C. 3 µβ nβγ · −Idβγ + ∇Dβγ + ∇Dβγ µγ 
*  T1T3   T 1+

K βγ = ∇Dγβ + ∇Dγβ : ∇Dγγ
  T1 µβ γ
= µγ nβγ · −Idγγ + ∇Dγγ + ∇Dγγ , at Aβγ , (48e) *  T1T3   T 1+
+ ∇Dββ + ∇Dββ : ∇Dβγ . (52)
ψkγ (r + li ) = ψkγ (r), i = 1, 2, 3; ψ = d, D; k = β, γ, (48f) β

To arrive at this result, we have made use of the identity given


hdkγ ik = 0, k = β, γ. (48g) in Eq. (B13).
043303-10 D. Lasseux and F. J. Valdés-Parada Phys. Fluids 29, 043303 (2017)

The remaining tensors in the macroscale model can be equations:


obtained by performing a similar analysis to Problem II. For   ∂Tβ    
the sake of brevity in presentation, we only report the final ρcp + ρcp vβ · ∇Tβ = ∇ · kβ ∇Tβ , in Vβ ,
β ∂t β
results, which are given by
(55a)
µβ 
* T1T3   T 1+

  ∂Tσ
=

Kγβ
µγ
∇Dβγ + ∇Dβγ : ∇Dββ ρcp = ∇ · (kσ ∇Tσ ) , in Vσ , (55b)
β σ ∂t
* T1T3   T 1+
+ ∇Dγγ + ∇Dγγ

: ∇Dγβ , (53a) where convection only takes place in the fluid-phase experi-
γ encing an incompressible and Newtonian flow governed by
the Stokes equations. The above equations are coupled at
µβ   T3    T 1  the solid-fluid interface by means of the following boundary

Kγγ = ∇Dβγ : ∇Dβγ + ∇Dβγ conditions:
µγ β
Tβ = Tσ , at Aβσ , (55c)
 T3    T 1 
+ ∇Dγγ : ∇Dγγ + ∇Dγγ . (53b)
γ

This last result on ∗


Kγγ clearly exhibits the same symmetry n · kβ ∇Tβ = n · kσ ∇Tσ , at Aβσ . (55d)
properties as K∗ββ (see Eq. (51)). Unfortunately, the tensors The result of applying the volume averaging method, without

Kγβ and K∗βγ do not show any particular symmetry properties. the imposition of the local thermal equilibrium assumption, is
Nevertheless, a reciprocity relationship can still be derived. To the following set of upscaled equations:
prove this point, it is useful to consider the identities given in
Eqs. (B23) and (B24), so that   ∂hTβ iβ
+ ε β ρcp hvβ iβ · ∇hTβ iβ
 
ε β ρcp
β ∂t β
µβ K∗βγ = µγ Kγβ
∗T
. (54)
− uββ · ∇hTβ i − uβσ · ∇hTσ iσ
β

The reciprocity relationship given in Eq. (54), which is in = ∇ · K ββ · ∇hTβ iβ + K βσ · ∇hTσ iσ


 
agreement with Onsager’s arguments, has been previously
− a3 h hTβ iβ − hTσ iσ ,
 
obtained by Auriault9,15 and Lasseux et al.17 (56a)
As a summary for this case study, it is worth recalling
that, for the upscaled model given in Eqs. (45), the tensors   ∂hTσ iσ
K∗ββ and Kγγ∗
are symmetric. In addition, the coupling per- ε σ ρcp − uσβ · ∇hTβ iβ − uσσ · ∇hTσ iσ
σ ∂t
meability tensors K∗βγ and Kγβ∗
, although they do not exhibit
= ∇ · K σβ · ∇hTβ iβ + K σσ · ∇hTσ iσ
 
a special symmetry or skew-symmetry property, are related
+ a3 h hTβ iβ − hTσ iσ ,
 
as shown in Eq. (54), thus reducing the number of effec- (56b)
tive medium coefficients to be computed to 3 instead of 4.
The irreducible decomposition of second-order tensors can where the scalar a3 h, the vectors uαk , and the tensors Kαk
be used for the coupling tensors in order to express them in (α, k = β, σ) are defined in terms of the corresponding closure
terms of their symmetric and skew-symmetric parts. However, variables (see Eq. (19) in the work of Quintard et al.48 ). In
for the sake of brevity, this decomposition is not presented this section, the attention is focused on the effective thermal
here. As a final note, it is worth pointing out that the rele- conductivity tensors defined as
vance of the coupling permeability tensors has been addressed

in several works,45–47 from which a definite conclusion is still 1 +  
K ββ = kβ ..ε I + nbββ dA// − ρcp hṽβ bββ i,
*
lacking. (57a)
V β
, Aβ σ -
III. DISPERSIVE HEAT AND MASS TRANSPORT 
kβ  
This last part of the article is devoted to the analysis of K βσ = nbβσ dA − ρcp hṽβ bβσ i, (57b)
V β
the effective transport coefficients involved in the upscaled Aβ σ
equations for dispersive heat and mass transport in the bulk
region of a rigid and homogeneous porous medium. The con- 

figuration is the one depicted in Fig. 1 showing the different K σβ = − nbσβ dA, (57c)
V
scales and the domain over which the pore-scale equations are Aβ σ
averaged. Heat transport, involving conduction and convec-
tion, is first considered under non-equilibrium conditions and
subsequently passive mass dispersion is addressed.
  
1
= kσ (1 − ε)I −
 
K σσ nbσσ dA . (57d)
 V 
A. Heat transport  Aβ σ 
In their study of heat dispersion in homogeneous porous The closure variables bββ and bσβ are the solution of the
media, Quintard et al.48 considered the following pore-scale following boundary-value problem.
043303-11 D. Lasseux and F. J. Valdés-Parada Phys. Fluids 29, 043303 (2017)

1. Problem I Applying the superficial averaging operator (see Eq. (A1a) in


    Appendix A) corresponding to the β-phase to Eqs. (62) leads
ρcp ṽ + ρcp vβ · ∇bββ = k β ∇2 bββ − ε −1 cββ , in Vβ , to
β β β
(58a) (ρcp )β hṽβ bββ iβ + (ρcp )β h(vβ · ∇bββ )bββ iβ
= h∇ · (kβ ∇bββ bββ )iβ − kβ h∇bTββ · ∇bββ iβ
0 = k σ ∇ bσβ − (1 − ε) cσβ ,
2 −1
in Vσ , (58b)
− hε −1 cββ bββ iβ . (64)
bββ = bσβ , at Aβσ , (58c) The last term on the above equation can be readily dropped
because both the porosity and cββ can be taken out from the
n · kβ ∇bββ = n · kσ ∇bσβ − nkβ , at Aβσ , (58d) averaging operator, leaving only bββ . Therefore, on the basis
of Eq. (58f), this last term is zero. Directing the attention to
bkβ (r + li ) = bkβ (r), i = 1, 2, 3; k = β, σ, (58e) the first term on the rhs of the above equation, and applying
the spatial averaging theorem, yields
hbkβ ik = 0, k = β, σ. (58f) h∇ · [kβ (∇bββ )bββ ]iβ = ∇ · hkβ ∇bββ bββ iβ

In the above expressions, 1
+ n · kβ ∇bββ bββ dA. (65)
 V
1
cββ = −cσβ = n · kβ ∇bββ dA. (59) Aβ σ
V
Aβ σ As already noticed in this work, since closure variables are
periodic, there are no spatial variations of average quantities.
In addition, the closure variables bβσ and bσσ are given
After application of the interfacial boundary conditions given
by the solution of the following boundary-value problem.
in Eq. (58d), this last result can be consequently expressed as
2. Problem II follows:


h∇ · [kβ (∇bββ )bββ ]iβ = −
 
ρcp v · ∇bβσ = k β ∇2 bβσ − ε −1 cβσ , in Vβ , (60a) nbββ dA
β β V
Aβ σ

0 = k σ ∇ bσσ − (1 − ε) cσσ ,
2 −1
in Vσ , (60b) 1
+ n · kσ ∇bσβ bσβ dA. (66)
V
Aβ σ
bβσ = bσσ , at Aβσ , (60c)
The first term on the rhs of this equation is present in the
n · kβ ∇bβσ = n · kσ ∇bσσ + nkσ , at Aβσ , (60d) definition of K ββ as indicated in Eq. (57a). Note that in the
second term, the boundary condition given in Eq. (58c) was
used. Moreover, since this term involves the closure variable
bkσ (r + li ) = bkσ (r), i = 1, 2, 3; k = β, σ, (60e)
bσβ , it is opportune to make the outer product of Eq. (58b)
with bσβ and then apply the superficial averaging operator
hbkσ ik = 0, k = β, σ. (60f) corresponding to the σ-phase, in order to obtain
In this case, 0 = h∇ · (kσ ∇bσβ )bσβ iσ − h(1 − ε)−1 cσβ bσβ iσ . (67)

1
cβσ = −cσσ = n · kβ ∇bβσ dA. (61) Following developments similar to those used in the β-phase,
V
Aβ σ it is not hard to deduce that

Note that, at the closure problem level, all the cαk (α, k = β, σ) 1
h∇ · (kσ ∇bσβ )bσβ iσ = − n · (kσ ∇bσβ )bσβ dA
vectors are constants within the unit cell. V
Aβ σ
Let us commence by analyzing the properties of the tensor
K ββ . From Eq. (57a), it follows that, in order to study the nature − kσ h∇bTσβ · ∇bσβ iσ = 0. (68)
of this tensor, it is convenient to make the outer product of Eq.
Using this expression in Eq. (66) and substituting the result
(58a) with bββ , in order to obtain
into Eq. (64) lead to

     
ρcp ṽβ bββ + ρcp vβ · ∇bββ bββ kβ
β β
  nbββ dA = −(ρcp )β hṽβ bββ iβ
= ∇ · kβ ∇bββ bββ − kβ ∇bTββ · ∇bββ V
Aβ σ
− ε −1 cββ bββ , in Vβ , (62) − (ρcp )β hvβ · ∇bββ bββ iβ
where we used the following identity in the diffusive term: − kσ h∇bTσβ · ∇bσβ iσ − kβ h∇bTββ · ∇bββ iβ .
   
k β ∇2 bββ bββ = ∇ · kβ ∇bββ bββ − kβ ∇bTββ · ∇bββ . (63) (69)
043303-12 D. Lasseux and F. J. Valdés-Parada Phys. Fluids 29, 043303 (2017)

In this way, Eq. (57a) can now be written as K σσ = kσ (1 − ε)I − kβ h∇bTβσ · ∇bβσ iβ
− kσ h∇bTσσ · ∇bσσ iσ − (ρcp )β hvβ · ∇bβσ bβσ iβ .
K ββ = kβ εI − kβ h∇bTββ · ∇bββ iβ − kσ h∇bTσβ · ∇bσβ iσ
(75)
− (ρcp )β hvβ · ∇bββ bββ iβ − 2(ρcp )β hṽβ bββ iβ . (70)
Finally, taking into account the interfacial boundary condition
Taking into account the spatial decomposition of the velocity given in Eq. (60c), the result given in Eq. (74) can also be used
and the average constraint in Eq. (58f), we may group the last to express K βσ as follows:
two terms of the above equation into a single one and express
K ββ in its final form, kσ
K βσ = kβ h∇bTβσ · ∇bβσ iβ + kσ h∇bTσσ · ∇bσσ iσ

* ! +
K ββ = kβ εI − kβ h∇bTββ · ∇bββ iβ − kσ h∇bTσβ · ∇bσβ iσ kσ
+ (ρcp )β vβ · ∇bβσ − I bβσ . (76)
kβ β
− (ρcp )β hvβ · (∇bββ + 2I)bββ iβ . (71)
Observation of the derived expressions for all the effective
Furthermore, on the basis of the interfacial boundary thermal conductivity coefficients shows that they all contain
condition in Eq. (58c), it can also be deduced that symmetric terms and one volume-averaged term that van-
ishes under non-convective conditions. It is thus interesting
kβ to carry out a deeper analysis in order to determine the struc-
K σβ = kσ h∇bTσβ · ∇bσβ iσ + kβ h∇bTββ · ∇bββ iβ
kσ ture of this last term. As a matter of fact, for all the effec-
+ (ρcp )β hvβ · (∇bββ + I)bββ iβ . (72) tive conductivities,
  this term involves the following average,
hvβ · ∇bβk bβk iβ (k = β, σ), which, due to the solenoidal
With the aim of analyzing the remaining effective- nature of the velocity field, can be written in the following
medium coefficients, let us now make the outer product of alternative form:
Eq. (60a) with bβσ and apply the superficial averaging oper- D   E D   E
ator for the β-phase. After performing similar steps to those vβ · ∇bβk bβk = ∇ · vβ bβk bβk
β β
used above, one easily arrives at D  E D E
= ∇ · vβ bβk bβk − bβk vβ · ∇bβk .
β β

kσ (77)
h∇ · (kβ ∇bβσ )bβσ iβ = nbσσ dA
V
Aβ σ Here we have also taken into account the identity given in Eq.
 (B18). It is not hard to prove that the first term on the rhs of
1
+ n · (kσ ∇bσσ ) bσσ dA the last equality of the above expression is null as a result of
V
Aβ σ the use of the spatial averaging theorem together with the non-
slip boundary condition at the β-σ interface. Finally, taking
− kβ h∇bTβσ · ∇bβσ iβ . (73) once again into account the solenoidal nature of vβ , it follows
that
To make further progress, let us make the outer product of
Eq. (60b) with bσσ and apply the superficial average in the
D E
∇ · (vβ bβk )bβk = −hbβk ∇ · (vβ bβk )iβ
σ-phase. After few steps that also involve the use of Eq. (73), β

the following expression is obtained: = −h∇ · (vβ bβk )bβk iTβ , (78)
 hence showing that this term is skew-symmetric. Unfortu-

nbσσ dA = (ρcp )β hvβ · ∇bβσ bβσ iβ nately, since the outer product between the velocity and the
V closure variable bβk has no particular symmetry properties,
Aβ σ
the term hvβ bβk iβ , which is present in all the effective thermal
+ kβ h∇bTβσ · ∇bβσ iβ + kσ h∇bTσσ · ∇bσσ iσ . conductivities (except in K σσ ) is neither symmetric nor skew-
(74) symmetric. However, this term can be simply decomposed into
its irreducible form. In this way, all the thermal conductivity
Substitution of this result into Eq. (57d) yields tensors can be expressed as

K ββ = kβ εI − kβ h∇bTββ · ∇bββ iβ − kσ h∇bTσβ · ∇bσβ iσ − (ρcp )β hvβ bββ + bββ vβ iβ


| {z }
Symmetric part

−(ρcp )β hvβ · ∇bββ bββ iβ − (ρcp )β hvβ bββ − bββ vβ iβ , (79a)


| {z }
Skew-symmetric part
043303-13 D. Lasseux and F. J. Valdés-Parada Phys. Fluids 29, 043303 (2017)

kβ 1
K σβ = kσ h∇bTσβ · ∇bσβ iσ + kβ h∇bTββ · ∇bββ iβ + (ρcp )β hvβ bββ + bββ vβ iβ
kσ | {z 2 }
Symmetric part
1
+ ρcp β hvβ · ∇bββ bββ iβ + (ρcp )β hvβ bββ − bββ vβ iβ ,

(79b)
| 2
{z }
Skew-symmetric part

kσ kσ D E
K βσ = kβ h∇bTβσ · ∇bβσ iβ + kσ h∇bTσσ · ∇bσσ )iσ − (ρcp )β vβ bβσ + bβσ vβ
kβ 2kβ β
| {z }
Symmetric part
 D E kσ D E
+ ρcp β vβ · ∇bβσ bβσ − (ρcp )β vβ bβσ − bβσ vβ , (79c)
β 2kβ β
| {z }
Skew-symmetric part

K σσ = kσ (1 − ε)I − kβ h∇bTβσ · ∇bβσ iβ − kσ h∇bTσσ · ∇bσσ iσ


| {z }
Symmetric part

− ρcp

β hvβ· ∇bβσ bβσ iβ . (79d)
| {z }
Skew-symmetric part

From the above results, the following comments are in order. equation involving diffusive and convective mass transport of
(1) The analysis carried out here allows us to exhibit the species A at the pore-scale is
dependence on convection of all the coefficients. This ∂cAβ  
contrasts with the original form given in Eqs. (57), + vβ · ∇cAβ = ∇ · Dβ ∇cAβ , in Vβ . (80a)
∂t
where the effective coefficients associated with the σ-
Assuming the solid phase to be impermeable to mass transport,
phase (K σσ and K σβ ) do not contain a hydrodynamic
the following interfacial boundary condition is applicable:
dispersion contribution as it is the case for the K ββ and
K βσ coefficients.
 
−n · Dβ ∇cAβ = 0, at Aβσ . (80b)
(2) The reformulation of the thermal conductivity tensors
shows that none of them are symmetric. However, the As in the heat transport analysis, this problem is coupled
present developments operate decomposition into the to momentum transport and the fluid velocity is subject to
symmetric and skew-symmetric parts in each tensor. nonslip conditions at the fluid-solid interface. As detailed by
The skew-symmetric parts of the tensors only contain Whitaker 1 (see Chapter 3), the result from using the volume
convective transport terms in the β-phase. averaging method to this problem is the following upscaled
(3) For transport conditions in which the thermal Péclet model:
number, defined as Pe = (ρcP )β khvβ iβ k` β /kβ , is much ∂hcAβ iβ
smaller than unity, all the coefficients are quasi- ε + hvβ i · ∇hcAβ iβ = εD∗β : ∇∇hcAβ iβ , (81)
∂t
symmetric and they are perfectly symmetric under
purely conductive conditions. where the total dispersion tensor is defined as

(4) For situations in which the assumption of local ther- D∗β 1 1
mal equilibrium is reasonable, there is only one effec- =I+ nfβ dA − hṽβ fβ iβ . (82)
Dβ Vβ DAβ
tive thermal conduction coefficient, Keff = K ββ + K βσ Aβ σ
+ K σβ + K σσ (see Eq. (15) in the work of Quintard et
al.48 ) to which all the above observations are applicable. In order to predict the values of this coefficient, it is thus
(5) As explained by Whitaker 1 (see Problems 3-7), only necessary to solve the following closure problem:
the symmetric parts of the effective medium coeffi- ṽβ + vβ · ∇fβ = Dβ ∇2 fβ , in Vβ , (83a)
cients influence the transport equation. However, if
the macroscopic boundary conditions are of Neumann n · ∇fβ = −n, at Aβσ , (83b)
or Robin-type, then the skew-symmetric parts of the
tensors cannot be discarded.
fβ (r + li ) = fβ (r), i = 1, 2, 3, (83c)
B. Mass dispersion
hfβ iβ = 0. (83d)
Finally, from the previous derivations, it is of interest to
address the passive mass dispersion problem in the bulk region As recently shown by Valdés-Parada et al.,22 the dispersion
of a rigid and homogeneous porous medium. The governing tensor can be re-written in an equivalent dimensional form,
043303-14 D. Lasseux and F. J. Valdés-Parada Phys. Fluids 29, 043303 (2017)

using exactly the same procedure employed in the previous Nevertheless, as pointed out during the analysis of heat trans-
paragraphs for heat transfer, as follows: port, this result is not extensible to situations in which mass
D∗β transport near the porous medium boundaries is prescribed
β
= I − h(∇fβ )T · ∇fβ i by Neumann or Robin-type boundary conditions.1 Indeed, in
Dβ this situation the skew-symmetric part of the dispersion tensor
1 2 cannot be neglected in the estimation of mass flux and this is
− hvβ · (∇fβ )fβ iβ − hvβ fβ iβ . (84)
Dβ Dβ even the dominant part under strongly convective conditions
The last term in this equation is exactly twice the so-called (i.e., for sufficiently large values of the mass Péclet number,
hydrodynamic dispersion tensor in the original work by say Pe = khvβ iβk` β /Dβ ). In this way, the conclusions from the
Whitaker.1 This term has no specific symmetry properties in work by Bear et al.27,28 can only hold if their total dispersion
the general case, except when the periodic unit cell represen- tensor is meant to be the same as the one arising from the
tative of the structure is symmetric and when the flow is along method of moments.
one of the symmetry axes. The third term on the rhs of Eq.
(84) can be easily proven to be skew-symmetric and this can
be obtained by following exactly the same procedure and argu- IV. CONCLUSION
ments as those used for Eqs. (77) and (78). Certainly, after In this work, we have used the framework of volume aver-
using the irreducible decomposition in this last term, the total aging as an upscaling tool to study the symmetry properties
dispersion tensor may be expressed as of effective-medium coefficients arising in macroscopic equa-
D∗β β 1 tions in many transport processes. In all cases we have followed
= I − h(∇fβ )T · ∇fβ i − hvβ fβ + fβ vβ iβ a similar procedure, which can be summarized as follows.
Dβ Dβ
| {z }
Symmetric part (1) Use the governing equations for the closure variables
1 1 that define the macroscopic transport coefficients and
− hvβ · (∇fβ )fβ iβ − hvβ fβ − fβ vβ iβ . (85) perform an appropriate (inner or outer) product with
Dβ Dβ
| {z } the adequate closure variable. Since the definition of
Skew-symmetric part the effective-medium coefficient is to be recovered from
Here, it is not hard to identify the same properties as for the the differential equation that governs the fields of the
effective thermal conductivity tensors, namely, a symmetric closure variable, the appropriate product to be used in
part present in the first three terms on the rhs of Eq. (85) and a each case is the one that leads to a tensorial equation of
skew-symmetric part (fourth and fifth terms). Not surprisingly, the same rank as the desired coefficient. For example,
the skew-symmetric part of D∗β only contains convective terms. in momentum transport the corresponding differential
There has been some confusion in the literature not only equations are already tensorial; therefore, the corre-
about the definition of the hydrodynamic dispersion part of D∗β sponding product is an inner product, whereas for heat
but also about the properties of the total dispersion tensor itself. and mass transport the equation is vectorial and thus an
For example, in the works by Bear et al.,27,28 the total disper- outer product is in order.
sion tensor is argued to be totally symmetric on the basis of the (2) Apply the superficial average to the previous result
conjugated thermodynamic force and flux relation. The present along with the averaging theorem and correspond-
development evidences that this cannot be the case due to the ing boundary and periodicity conditions in the closure
presence of skew-symmetric terms, which are only negligible problem to obtain, after few algebraic manipulations
under poorly or non-convective conditions. This conclusion involving tensor analysis, an alternative expression of
is in agreement with the results presented by Auriault et al.21 the corresponding macroscopic transport coefficient in
using the homogenization method and confirmed by numerical terms of the closure variables.
simulations. As remarked by these authors, Onsager’s relations (3) The resulting expression for the effective-medium
cannot be invoked here due to the irreversible character of the coefficients has the nice feature that it allows a straight-
mass and momentum transport processes taking place at the forward identification of the symmetry and skew-
pore-scale. symmetric parts of the tensors.
It should be noted that the symmetric part of D∗β in Eq. (85)
The technique outlined above was widely illustrated over
is the one resulting from the method of moments, as reported
several study cases, from which the following comments
by Brenner 24 (see Sections 5 and 6 therein) and later on by
arise.
Salles et al.25 This is consistent with what has been pointed
out by Koch and Brady,26 who showed that the method of • For inertial one-phase flow in homogeneous porous
moments, or the Lagrangian approach, can only predict the media, the method performs the irreducible decom-
symmetric part of the total dispersion tensor. Importantly, position of the apparent permeability tensor, showing
the present analysis clearly evidences the complementary (i.e., that the symmetric part results from viscous dissipa-
the skew-symmetric) parts of this tensor that are missing from tion, while the skew-symmetric part originates from
the method of moments analysis. As a matter of fact, these inertial transport only. Hence, under creeping flow con-
parts of D∗β do not contribute to the macroscopic mass trans- ditions, the apparent permeability tensor corresponds
port as can be easily inferred from the rhs of Eq. (81); this is to the intrinsic one, which is fully symmetric as it is
simply due to the fact that ∇∇hcAβ iβ is a symmetric tensor. well-known in the literature.
043303-15 D. Lasseux and F. J. Valdés-Parada Phys. Fluids 29, 043303 (2017)

• When the slip-boundary condition is applicable, even the upscaling process does not require any specification of the
in the creeping-flow regime, the apparent permeability porous medium geometry except that, at the closure level, the
tensor is not symmetric. structure is made pseudo-periodic.
• For one-phase slightly compressible gas flow, in the slip
regime within a fracture, the effective transmissivity ACKNOWLEDGMENTS
tensor is symmetric despite the presence of slip effects.
The result obtained in this case can be generalized to Inspiration of this development is intimately related to
any diffusion process involving continuously position wonderful exchanges with F.T.C.P. to whom D.L. would like
dependent diffusivities. to warmly dedicate this work. F.J.V.-P. is thankful to the
• The effective permeability tensor for one-phase flow in CNRS for the facilities provided to carry out a scientific
a heterogeneous porous medium embedding a porous stay with D.L. during spring 2016. D.L. acknowledges the
matrix and a clear fluid region was shown to be sym- financial support from CONACyT to perform a two-week
metric when the flow is assumed to be governed by the stay in UAM-Iztapalapa to work with F.J.V.-P. during autumn
Darcy-Brinkman equations in the porous medium and 2016.
by the Stokes equations in the clear fluid region.
• In the case of two-phase creeping flow in homoge- APPENDIX A: OUTLINE OF THE VOLUME
neous porous media, the dominant permeability ten- AVERAGING METHOD
sors are shown to be symmetric, while the coupling
permeability tensors do not exhibit specific symme- In this section, the method of volume averaging is briefly
try properties. A reciprocity relationship between the outlined. The steps presented here are consistent with the
latter was recovered in consistency with a previous monograph by Whitaker 1 and also with the more recent ver-
analysis. sion reported by Wood and Valdés-Parada.49 In Fig. 5, the
• While studying conductive and convective heat transfer main steps, assumptions, and tools involved in this upscaling
under non-local thermal equilibrium conditions, it was method are schematized and listed below.
demonstrated that the effective thermal conductivity (1) First, a set of starting assumptions are introduced with
tensors include a symmetric and a skew-symmetric part, the aim of defining the transport problem at the pore
the latter involving convective effects only. When con- scale.
vection is unimportant or absent, the effective thermal (2) An averaging domain V (of measure V ) that contains
conductivity tensors become symmetric. portions of all the phases involved in the system is then
• The results from the previous item are extensible to defined along with the corresponding superficial
passive mass dispersion in homogeneous porous media. 
1
More importantly, our developments allow us to explic- hψα i = ψα dV (A1a)
V
itly identify the skew-symmetric part of the total dis- Vα
persion tensor that is otherwise missing when derived
and intrinsic
from the method of moments. 
α 1
hψα i = ψα dV (A1b)
Beyond the study cases included in this work, the tech- Vα
nique used here can be applied to any other transport process Vα

that is upscaled using the volume averaging method. Hence, the averaging operators for a piece-wise continuous func-
procedure can be regarded to be systematic for the analysis of tion, ψα , defined everywhere in the α-phase. Actu-
the symmetry properties of macroscopic transport coefficients ally, these averaging operators are coupled by means
as it was the goal of this work. Finally, it is worth mention- of the Dupuit-Forchheimer relation hψα i = ε α hψα iα ,
ing that the conclusions reached in this analysis remain valid with ε α = Vα /V being the volume fraction of the
whatever the microstructure of the porous medium. Indeed, α-phase within the averaging domain. For the case in

FIG. 5. Scheme of the upscaling process using the


method of volume averaging.
043303-16 D. Lasseux and F. J. Valdés-Parada Phys. Fluids 29, 043303 (2017)

which there is only one fluid phase (the β-phase) (see Using orders of magnitude estimates and taking into
Fig. 1) saturating the porous matrix, the fluid volume account this separation of length scales, the differen-
fraction corresponds to the porosity, and throughout this tial equations and boundary conditions in the closure
work, it is denoted by ε β ≡ ε. problem can be considerably simplified. In this way, the
(3) The superficial averaging operator defined in Eq. (A1a) average quantities present in the closure problem are
is applied to the pore-scale equations. In addition, the position-invariant. A corollary of this approximation
interchange of spatial integration and differentiation is that the deviations fields must satisfy the following
can be achieved by means of the spatial averaging average constraint:
theorem,50 which, for any quantity ψα defined in the α
hψ̃α i = 0. (A5)
α-phase, is given by
 A convenient, although non-mandatory, simplification
1
h∇ψα i = ∇hψα i + nα ψα dA (A2) of the closure problem is that, in many situations, its
V solution domain can be reduced to a single periodic unit

cell. Therefore, one may impose the following bound-
and a completely similar expression in its divergence ary condition at the entrances and exits of the unit
form. cell:
It is worth adding that the application of the averaging
ψ̃α (r + li ) = ψ̃α (r), i = 1, 2, 3, (A6)
theorems often allows substitution of the correspond-
ing interfacial boundary conditions. In Eq. (A2), Aα with r and li being a position vector and each of the
denotes all the surfaces with which the α-phase is in lattice vectors in the unit cell, respectively. Under these
contact and nα is a unit normal vector directed from the conditions, let us refer to this version of the closure
α-phase toward each of the other phases in contact. For problem as the simplified closure problem, which has the
the particular case in which there is a single fluid-phase nice quality of requiring knowledge of less information
saturating the porous medium, Aβ = Aβσ and nβσ than the pore-scale model.
≡ n. (6) Substitution of the formal solution of the simplified
(4) At this point, the average equations are expressed in closure problem into the filters terms of the unclosed
terms of both average and pointwise quantities. To elim- average model leads to its closed version, which can
inate the latter, the spatial decomposition introduced by be subsequently reduced by performing orders of mag-
Gray51 is used, nitude analyses taking into account the separation of
length scales already imposed in the simplified closure
ψα = hψα iα + ψ̃α , (A3) problem. After performing this final development, the
ψ̃α being the spatial deviations of ψα about its intrin- resulting model may be referred to as the simplified
sic average. The resulting expression is the unclosed upscaled model.
average model because it lacks of a relation between
the spatial deviations and average quantities. Following APPENDIX B: TENSOR IDENTITIES
the work of Whitaker,1 the (surface and/or volumet-
ric) integrals containing deviation terms may be con- In this manuscript, many identities involving tensor alge-
ceived as filters of information coming from the pore bra are used that are listed in the present appendix. For the sake
scale. of clarity, basic definitions and necessary material to recover
(5) With the aim of closing the average model, a boundary- them are first recalled in Gibbs’ notation together with the
value problem for the spatial deviations is derived Einstein summation convention. Let A and b be respectively
and formally solved. Since no assumptions have been a second-order tensor and a vector having all the necessary
imposed so far, let us refer to this boundary-value prob- properties of regularity in the three dimensions of space for
lem as the exact closure problem. However, the price to the derivation operators to be defined on their fields; these
∂ψ
be paid is that its complexity is the same as (or even basic definitions are given by (we use the notation ψ,i = ∂x i
)
greater than) that of the pore-scale model. It is thus (∇b)ij = bj,i , (B1)
convenient to introduce a first set of scaling postulates
∇ · A j = Aij,i ,

with the aim of filtering out the redundant information (B2)
∇A ijk = Ajk,i .

present in the pore-scale. In typical volume averag- (B3)
ing applications, these scaling postulates consist of a
Note here that ∇A is a third-order tensor,
set of reasonable length-scale constraints and assump-  
∇2 A = ∇ · ∇A ij = ∇A kij,k = Aij,k,k .
 
tions among which it is worth recalling the following (B4)
ij
one:
Throughout the article, the nested convention for inner
`  r0  L (A4) products is employed. This means that the indices that are
closest together are those on which summation applies. As a
in which r 0 is the characteristic size of the averag- consequence, the double inner product between two second-
ing domain, while ` and L represent the largest char- order tensors A and B is defined as
acteristic length of the pore-scale and the smallest
characteristic length of the macroscale, respectively. A : B = Aij Bji = B : A. (B5)
043303-17 D. Lasseux and F. J. Valdés-Parada Phys. Fluids 29, 043303 (2017)

Similarly, if C is a third-order tensor, we have (3) For any arbitrary second-order tensors A and B hav-
C · A ijk = Cijl Alk

(B6) ing the required regularity properties, the following
identities hold:
and further, if F is another third-order tensor
A · B T = BT · A T ,

(B21)
C : F ik = Cijl Fljk .

(B7)
Importantly, while dealing with third-order tensors, one must ∇ · ∇A · B = ∇2 A · B + ∇A T 1 : ∇B T 1 ,
  
be clear about the transposes of such algebraic quantities
(B22)
as three different transpose operators can be defined. In this
manuscript, two of them are employed. The transpose denoted T T3 T3
by the superscript T 1 permutes the first and second indices, ∇A : ∇B = ∇B : ∇A , (B23)
whereas the transpose denoted by the superscript T 3 permutes
f  gT3
the first and third indices, namely (C is again a third-order ∇A + ∇A T 1 : ∇B T 1

tensor), f  gT3
 T1 = ∇A + ∇A T 1 : ∇B
C = Cjik , (B8)
ijk
1 f  T3 f
g  g
= ∇A + ∇A T 1 : ∇B + ∇B T 1 .
2
 
CT 3 = Ckji . (B9)
ijk
(B24)
In addition, throughout this work, the outer product
between two vectors, a and b, is defined as (4) The last part of this appendix is dedicated to the proof
that, for any second-order tensor A having the required
(ab)ij = ai bj . (B10) regularity properties, the second-order tensorg given by
f
It is worth mentioning that some authors52 the expression B = ∇A T 3 : ∇A + ∇A T 1 is sym-
 
denote this product
as a ⊗ b. The outer product between a second-order tensor A metric. The ij component of this tensor is given by
and a vector b is given in a similar manner by      
B ij = ∇A T 3 ∇A + ∇A T 1

ikl lkj
Ab ijk = Aij bk .

(B11)
= ∇A lki ∇A lkj + ∇A lki ∇A klj .
   
(B25)
With this at hand, it is not hard to deduce the following
formulas that are used in the article and which can be listed as The ji component of this tensor is obtained by inverting
follows. the i and j indices, yielding
(1) For a second-order tensor, A, which is a regular enough B ji = ∇A lkj ∇A lki + ∇A lkj ∇A kli .
    
(B26)
function of position, the following identities are applica-
ble (the superscript T represents The first term on the rhs of both Eqs. (B25) and (B26) is
 the
 classical transpose
of a second-order tensor, i.e., AT = Aji ), obviously the same. Moreover, for the second term, one
ij can notice that the two indices on which summation is
T
performed are dummy and can hence be interchanged
  T  
∇ A = ∇ · ∇A
2
= ∇2 AT , (B12)
so that the second term on the rhs of both the above
 T1  T1T3 expressions is also the same, completing the proof that
∇AT = ∇A , (B13) T3 f T1g
∇A : ∇A + ∇A is a symmetric tensor.
T1   T
∇ · ∇A = ∇ · ∇A T 3 , (B14)
1 S. Whitaker, The Method of Volume Averaging (Kluwer Academic Publish-

 T3T1 ers, 1999).


T3
∇AT = ∇A
2 J. L. Auriault, C. Boutin, and C. Geindreau, Homogenization of Coupled
, (B15)
Phenomena in Heterogeneous Porous Media (Wiley, Hoboken, 2009).
3 W. Gray and C. Miller, Introduction to the Thermodynamically Constrained
 T1 T1 T3
∇AT : ∇A = ∇A : ∇A. (B16) Averaging Theory for Porous Medium Systems (Springer, 2014).
4 L. Gelhar, Stochastic Subsurface Hydrology (Prentice Hall, 1993).
5 G. Matheron, Elements pour une Théorie des Milieux Poreux (Masson &
(2) For a second-order tensor A and a vector b, which are
Cie, 1967).
both continuous and derivable functions of position, the 6 E. Sanchez-Palencia, Nonhomogeneous Media and Vibration Theory, Lec-
following identities are applicable: ture Notes in Physics (Springer, 1980).
7 J. Auriault, L. Borne, and R. Chambon, “Dynamics of porous saturated
AT b = bA T 3 ,

(B17)
media, checking of the generalized law of Darcy,” J. Acoust. Soc. Am. 77,
1641–1650 (1985).
∇ · (Ab) = (∇ · A)b + AT · ∇b. (B18) 8 E. Skjetne and J.-L. Auriault, “Homogenization of wall-slip gas flow through

porous media,” Transp. Porous Media 36, 293–306 (1999).


When b is a solenoidal vector field, the following 9 J. Auriault, “Transport in porous media: Upscaling by multiscale asymptotic
identities are valid: expansions,” in Applied Micromechanics of Porous Materials, Courses and
Lectures, edited by L. Dormieux and F. J. Ulm (Springer Wien, New-York,
AT · ∇ · (bA) = AT b : ∇A,
 
(B19) 2005), Vol. 480, pp. 3–56.
10 G. Dagan, Flow and Transport in Porous Formations (Springer-Verlag,
   
∇ · bAT · A = ∇ · bAT · A + AT b : ∇A. 1989).
11 S. Whitaker, “The Forchheimer equation: A theoretical development,”

(B20) Transp. Porous Media 25, 27–61 (1996).


043303-18 D. Lasseux and F. J. Valdés-Parada Phys. Fluids 29, 043303 (2017)

12 D. Lasseux, F. Valdés-Parada, and M. Porter, “An improved macroscale 33 J. C. Maxwell, “On stresses in rarified gases arising from inequal-
model for gas slip flow in porous media,” J. Fluid Mech. 805, 118–146 ities of temperature,” Philos. Trans. R. Soc. London 170, 231–256
(2016). (1879).
13 D. L. Koch and A. J. C. Ladd, “Moderate Reynolds number flows through 34 E. Lauga, M. P. Brenner, and H. A. Stone, “Microfluidics: The no-slip bound-

periodic and random arrays of aligned cylinders,” J. Fluid Mech. 349, 31–66 ary condition,” in Handbook of Experimental Fluid Dynamics, edited by
(1997). J. Foss, C. Tropea, and A. Yarin (Springer, New York, 2007), Chap. 19, pp.
14 D. Lasseux, A. A. Abbasian-Arani, and A. Ahmadi, “On the stationary 1219–1240.
35 D. Lasseux, F. Valdés-Parada, J. Ochoa-Tapia, and B. Goyeau, “A macro-
macroscopic inertial effects for one phase flow in ordered and disordered
porous media,” Phys. Fluids 23, 073103 (2011). scopic model for slightly compressible gas slip-flow in homogeneous porous
15 J. Auriault, “Nonsaturated deformable porous media: Quasistatics,” Transp. media,” Phys. Fluids 26, 053102 (2014).
36 C. Marie, D. Lasseux, H. Zahouani, and P. Sainsot, “An inte-
Porous Media 2, 45–64 (1987).
16 S. Whitaker, “The closure problem for two-phase flow in homogeneous grated approach to characterize liquid leakage through metal contact
porous media,” Chem. Eng. Sci. 49, 765–780 (1994). seal,” Eur. J. Mech. Environ. Eng. 48, 81–86 (2003), see https://
17 D. Lasseux, M. Quintard, and S. Whitaker, “Determination of permeability www.researchgate.net/publication/289479488.
37 C. Marie and D. Lasseux, “Experimental leak-rate measurement through a
tensors for two-phase flow in homogeneous porous media: Theory,” Transp.
Porous Media 24, 107–137 (1996). static metal seal,” J. Fluids Eng. 129, 799–805 (2007).
18 R. Carbonell and S. Whitaker, “Dispersion in pulsed systems II: Theoretical 38 C. Vallet, D. Lasseux, P. Sainsot, and H. Zahouani, “Real versus synthesized

developments for passive dispersion in porous media,” Chem. Eng. Sci. 38, fractal surfaces: Contact mechanics and transport properties,” Tribol. Int.
1795–1802 (1983). 42, 250–259 (2009).
19 J. L. Auriault and P. M. Adler, “Taylor dispersion in porous media: Analysis 39 F. Pérez-Ràfols, R. Larsson, and A. Almqvist, “Modelling of leakage on

by multiple scale expansions,” Adv. Water Resour. 18, 217–226 (1995). metal-to-metal seals,” Tribol. Int. 94, 421–427 (2016).
20 H. P. Amaral-Souto and C. Moyne, “Dispersion in two-dimensional periodic 40 M. Prat, F. Plouraboué, and N. Letalleur, “Averaged Reynolds equation for

porous media. Part I. Hydrodynamics,” Phys. Fluids 9, 2243–2252 (1997). flows between rough surfaces in sliding motion,” Transp. Porous Media 48,
21 J. Auriault, C. Moyne, and H. Amaral-Souto, “On the asymmetry of the 291–313 (2002).
41 M. Quintard and S. Whitaker, “Écoulement monophasique en milieu poreux:
dispersion tensor in porous media,” Transp. Porous Media 85, 771–783
(2010). Effet des hétérogénéités locales,” J. Mec. Theor. Appl. 6, 691–726 (1987),
22 F. Valdés-Parada, D. Lasseux, and F. Bellet, “A new formulation of the see https://www.researchgate.net/publication/230704130.
42 T. Arbogast and H. Lehr, “Homogenization of a Darcy-Stokes system
dispersion tensor in homogeneous porous media,” Adv. Water Resour. 90,
70–82 (2016). modeling vuggy pous media,” Comput. Geosci. 10, 291–302 (2006).
23 L. G. Fel and J. Bear, “Dispersion and dispersivity tensors in saturated 43 F. Golfier, D. Lasseux, and M. Quintard, “Investigation of the effective

porous media with uniaxial symmetry,” Transp. Porous Media 85, 259–268 permeability of vuggy or fractured porous media from a Darcy-Brinkman
(2010). approach,” Comput. Geosci. 19, 63–78 (2014).
24 H. Brenner, “Dispersion resulting from flow through spatially periodic 44 S. Whitaker, “Flow in porous media II: The governing equations for

porous media,” Philos. Trans. R. Soc., A 297, 81–133 (1980). immiscible, two-phase flow,” Transp. Porous Media 1, 105–125 (1986).
25 J. Salles, J. Thovert, R. Delannay, L. Prevors, J. Auriault, and P. Adler, 45 C. Zarcone and R. Lenormand, “Détermination expérimentale du couplage

“Taylor dispersion in porous media. Determination of the dispersion tensor,” visqueux dans les écoulements diphasiques en milieu poreux,” C. R. Acad.
Phys. Fluids A 5, 2348–2376 (1993). Sci., Ser. IIb 318, 1429–1435 (1994).
26 D. Koch and J. Brady, “The symmetry properties of the effective dif- 46 T. Ramakrishnan and P. Goode, “Measurement of off-diagonal transport

fusivity tensor in anisotropic porous media,” Phys. Fluids 30, 642–650 coefficients in two-phase flow in porous media,” J. Colloid Interface Sci.
(1987). 449, 392–398 (2015).
27 J. Bear and A. Cheng, Modeling Groundwater Flow and Contaminant 47 H. Li, C. Pan, and C. Miller, “Pore-scale investigation of viscous coupling

Transport (Springer, 2010). effects for two-phase flow in porous media,” Phys. Rev. E 72, 026705 (2005).
28 J. Bear, L. G. Fel, and Y. Zimmels, “Effects of material symmetry on the 48 M. Quintard, M. Kaviany, and S. Whitaker, “Two-medium treatment of heat

coefficients of transport in anisotropic porous media,” Transp. Porous Media transfer in porous media: Numerical results for effective properties,” Adv.
82, 347–361 (2010). Water Resour. 20, 77–94 (1997).
29 L. Onsager, “Reciprocal relations in irreversible processes,” Phys. Rev. 37, 49 B. Wood and F. Valdés-Parada, “Volume averaging: Local and nonlocal

405–426 (1931). closures using a Green’s function approach,” Adv. Water Resour. 51, 139–
30 H. B. G. Casimir, “On onsager’s principle of microscopic reversibility,” 167 (2013).
50 F. Howes and S. Whitaker, “The spatial averaging theorem revisited,” Chem.
Rev. Mod. Phys. 17, 343–350 (1945).
31 J. Auriault and Y. Lewandowska, “On the cross-effects of coupled macro- Eng. Sci. 40, 1387–1392 (1985).
51 W. Gray, “A derivation of the equations for multiphase transport,” Chem.
scopic transport equations in porous media,” Transp. Porous Media 16,
31–52 (1994). Eng. Sci. 30, 229–233 (1975).
32 M. Navier, Mémoire sur les Lois du Mouvemet des Fluides (Acadmie Royale 52 W. Lai, D. Rubin, and E. Krempl, Introduction to Continuum Mechanics,

des Sciences de l’Institut de France, 1822). 4th ed. (Elsevier, 2009).

Você também pode gostar