You are on page 1of 13

Advances in Water Resources 90 (2016) 70–82

Contents lists available at ScienceDirect

Advances in Water Resources


journal homepage: www.elsevier.com/locate/advwatres

A new formulation of the dispersion tensor in homogeneous porous


media
Francisco J. Valdés-Parada a, Didier Lasseux b,∗, Fabien Bellet c
a
Departamento de I.P.H., Universidad Autónoma Metropolitana-Iztapalapa. Av. San Rafael Atlixco 186, Col. Vicentina, México, D.F., 09340, Mexico
b
CNRS, UMR 5295, Univ. Bordeaux, Esplanade des Arts et Métiers, 33405 Talence, Cedex, France
c
Laboratoire EM2C, CNRS, CentraleSupélec, UniversitéParis-Saclay, Grande Voie des Vignes, Chatenay-Malabry Cedex 92295, France

a r t i c l e i n f o a b s t r a c t

Article history: Dispersion is the result of two mass transport processes, namely molecular diffusion, which is a pure
Received 22 February 2015 mixing effect and hydrodynamic dispersion, which combines mixing and spreading. The identification of
Revised 17 February 2016
each contribution is crucial and is often misinterpreted. Traditionally, under a volume averaging frame-
Accepted 19 February 2016
work, a single closure problem is solved and the resulting fields are substituted into diffusive and disper-
Available online 27 February 2016
sive filters. However the diffusive filter (that leads to the effective diffusivity) allows passing information
Keywords: from convection, which leads to an incorrect definition of the effective medium coefficients composing
Dispersion the total dispersion tensor. In this work, we revisit the definitions of the effective diffusivity and hydrody-
Effective diffusion namic dispersion tensors using the method of volume averaging. Our analysis shows that, in the context
Volume averaging of laminar flow with or without inertial effects, two closure problems need to be computed in order to
Closure problem correctly define the corresponding effective medium coefficients. The first closure problem is associated
to momentum transport and needs to be solved for a prescribed Reynolds number and flow orientation.
The second closure problem is related to mass transport and it is solved first with a zero Péclet number
and second with the required Péclet number and flow orientation. All the closure problems are written
using closure variables only as required by the upscaling method. The total dispersion tensor is shown
to depend on the microstructure, macroscopic flow angles, the cell (or pore) Péclet number and the cell
(or pore) Reynolds number. It is non-symmetric in the general case. The condition for quasi-symmetry is
highlighted. The functionality of the longitudinal and transverse components of this tensor with the flow
angle is investigated for a 2D model porous structure obtaining consistent results with previous studies.
© 2016 Elsevier Ltd. All rights reserved.

1. Pore-scale model ered as a constant in this work, whereas vβ is the fluid velocity,
which satisfies the total mass and momentum conservation equa-
Dispersion of a solute (species A) in porous media is a funda- tions at the pore-scale:
mental subject that has been largely studied over the past century ∇ ·vβ = 0, in the β − phase (2a)
and it remains as an interesting study field due to the wide range
of applications that it encompasses. This transport mechanism is ρβ vβ · ∇ vβ = −∇ pβ +μβ ∇ 2 vβ , in the β − phase (2b)
the result of diffusion, which is related to a pure mixing effect, Note that flow has been assumed to be steady, incompressible
and variations in the convective fluxes within the pores leading to and Newtonian. Without loosing generality, gravity has been omit-
both mixing and spreading. The governing pore-scale mass balance ted in the momentum equation. In Eq. (2b), ρ β and μβ denote
equation for a solute (species A) transported within the fluid phase the fluid density and dynamic viscosity, respectively, which are as-
(the β -phase) that saturates the porous medium is sumed to be constants. Furthermore, for the sake of simplicity in
∂ c Aβ the analysis, the porous medium is assumed to be rigid and homo-
+ ∇ · (cAβ vβ ) = ∇ · (Dβ ∇ cAβ ), in the β − phase (1)
∂t geneous, so that intrinsic average properties (e.g. the porosity, the
permeability, etc.) are position-invariant. In addition, the non-slip
Here cAβ and Dβ respectively denote the pointwise species A con-
boundary condition is imposed at the solid–fluid interface Aβσ
centration and the molecular diffusivity, the latter being consid-
vβ = 0, at Aβσ (3)

Corresponding author. Tel.: +33 556 845 403; fax: +33 556 845 436. and the solid phase (i.e., the σ -phase) is assumed impermeable to
E-mail address: didier.lasseux@ensam.eu (D. Lasseux). mass transfer, so that

http://dx.doi.org/10.1016/j.advwatres.2016.02.012
0309-1708/© 2016 Elsevier Ltd. All rights reserved.
F.J. Valdés-Parada et al. / Advances in Water Resources 90 (2016) 70–82 71

Nomenclature V averaging domain


V norm of the averaging domain, m3
Aβσ solid–fluid interface Vβ domain occupied by the β -phase within V
Aβσ norm of the solid–fluid interface contained in the Vβ norm of Vβ , m3
averaging domain, m2 Vβ∗ (=Vβ /3 ) dimensionless form of Vβ
A∗βσ (= Aβσ /2 ) dimensionless area of the solid–fluid in-
terface Greek symbols
cAβ molar concentration of species A in the β -phase, εβ volume fraction of the fluid phase contained in the
averaging domain
mol/m3
λ (= ∇ pβ β /∇ pβ β ) unit vector in the direction
cAβ β intrinsic average of cAβ , mol/m3
of the macroscopic pressure gradient
c˜Aβ spatial deviations of cAβ with respect to cAβ β ,
λv (= vβ /vβ ) unit vector in the direction of the
mol/m3 macroscopic velocity
Dβ species mixture diffusivity in the β -phase, m2 /s μβ viscosity of the β -phase, Pa · s
D∗β total dispersion tensor, m2 /s θ macroscopic pressure gradient inclination on ex rad
Dβ hydrodynamic dispersion tensor, m2 /s θv macroscopic flow inclination on ex , rad
De f f intrinsic effective diffusion tensor, m2 /s ρβ density of the β -phase, kg/m3
D∗e f f effective diffusion tensor that depends of the flow
rate, m2 /s
eβ closure variable that maps ∇pβ β onto p˜ β , m −n · Dβ ∇ cAβ = 0, at Aβσ (4)
e∗β (= eβ /) dimensionless version of eβ
Despite the simple form of the governing equations at the pore-
Eβ closure variable associated to the spatial variations
scale, all the essential elements are contained to give rise to the
of the velocity, m2
well-known convection-dispersion equation for mass transport af-
E∗β (= Eβ /2 ) dimensionless version of Eβ
ter an upscaling process is applied. The study of dispersion from
fβ closure variable that maps ∇cAβ β onto c˜Aβ , m different theoretical points of view is available in several refer-
f∗β (= fβ /) dimensionless form of fβ ences, a couple of examples being the works by Cushman et al. [1]
f0 (= fβ |vβ =0 ) closure variable that maps ∇cAβ β onto and Chapter 11 of Sahimi [2]. In particular, the study of dispersion
c˜Aβ under purely diffusive conditions, m using the volume averaging method dates back to the classic pa-
f∗0 (= f0 /) dimensionless form of f0 per of Whitaker in 1967 [3], in which the macroscopic model was
Fβ Forchheimer correction tensor derived from the pore-scale equations by means of an averaging
process. However, no closure procedure was provided at that mo-
Hβ apparent permeability tensor, m2
ment and thus the effective-medium coefficients involved in the
H∗β (=Hβ /2 ) dimensionless version of Hβ
model had to be estimated experimentally. The model derivation
I identity tensor was later improved by correctly defining the spatial decomposi-
Kβ intrinsic permeability tensor, m2 tion of pore-scale variables as detailed by Gray [4]. The closure
li lattice vectors associates to the unit cell, i = 1, 2, 3, process, under a volume averaging approach, was presented in a
m series of works dealing with dispersion in pulsed systems. Firstly,
 characteristic length of the unit cell, m transport of a solute in capillaries undergoing heterogeneous reac-
β characteristic length associated to the fluid phase, tion and adsorption was presented by Paine et al. [5]. Their study
m showed the time dependence of the dispersion coefficient under
L characteristic length associated to the macroscopic different reactive conditions. As expected, under quasi-steady con-
scale, m ditions, the well-known Taylor [6] and Aris [7] results were recov-
nβσ unit normal vector pointing from the fluid phase to- ered. Following this approach, the closure problem for dispersion
wards the solid phase in porous media was derived by Carbonell and Whitaker [8] and it
pβ pressure of the β -phase, Pa was subsequently solved in periodic unit cells by Eidsath et al. [9]
Pe cell Péclet number obtaining good agreement with experimental data for particle Pé-
Pep pore-scale Péclet number clet number values below 10 0 0 [see Fig. 13 in 9]. This type of anal-
r position vector, m ysis was soon extended to study heterogeneous porous materials
r0 characteristic size of the averaging domain, m by Plumb and Whitaker [10] and the corresponding closure prob-
Re∗ Reynolds number based on the unit cell length and lem was solved in spatially periodic unit cells for stratified porous
macroscopic pressure gradient media [11]. In a later work by Quintard et al. [12], a one-equation
Re Reynolds number based on the unit cell length and model, without involving large-scale mass equilibrium, was derived
the macroscopic velocity magnitude for studying mass dispersion in heterogeneous porous media. In
Rep Reynolds number based on the pore-size and inter- addition, solutions to the closure problems have also been carried
stitial velocity out for two-equation models in fissured media using the volume
Sc Schmidt number element method as shown by Caillabet, et al. [13]. It is worth men-
t time, s tioning that Wood et al. [14] proved that the closure problem so-
t0 characteristic time-scale associated to diffusion at lution for dispersion in heterogeneous porous media is equivalent
the pore-scale, s to the one resulting from the ensemble averaging method.
vβ velocity vector of the fluid phase, m/s The solution of the closure problem for homogeneous porous
vβ β intrinsic average of vβ , m/s media was further investigated by Amaral-Souto and Moyne [15]
v˜ β spatial variations of the velocity with respect to considering ordered and disordered geometries for the solid phase
vβ β , m/s in the unit cell as well as the flow orientation for particle Reynolds
number values ranging in the laminar-inertial regime. The
72 F.J. Valdés-Parada et al. / Advances in Water Resources 90 (2016) 70–82

pertinence of using these geometries for the closure problem so-


lution was validated by comparison with experiments in a subse-
quent work by Didierjean, et al. [16]. Later on, Wood [17] studied
the role of inertial effects over dispersion in homogeneous porous
media, finding that longitudinal dispersion was not dramatically
affected by inertia, whereas transverse dispersion was enhanced by
a factor of 40 with respect to creeping flow conditions. Recently,
Aguilar-Madera et al. [18] reported that the flow direction is the
main cause of anisotropy of the dispersion tensor, especially for
the transverse component of the tensor.
Application of the volume averaging method has not been re-
stricted to passive dispersion under one-phase flow condition;
Quintard and Whitaker [19] studied active dispersion, which cor-
responds to the dissolution of trapped non-aqueous phase liquids
(NAPL) in the water phase that saturates homogeneous porous me-
dia. These authors solved the corresponding closure problems in
simple unit cells and later on Ahmadi et al. [20] carried out the
computations in complex unit cells involving thousands of pores.
A more general problem of multiphase-multicomponent fluid flow
in homogeneous and heterogeneous porous media was discussed Fig. 1. Sketch of the system, including characteristic lengths and a sample of the
in [31]. averaging domain.
In all the above applications of the volume averaging method,
the closure problem formulation follows essentially the same phi-
losophy outlined by Carbonell and Whitaker [8]. In this approach, where ε β ≡ Vβ /V is the volume fraction of the fluid phase con-
the closure problem requires knowledge of the pore-scale veloc- tained in the averaging domain, i.e. the porosity in the present de-
ity fields and their deviations. In addition, the total dispersion ten- velopment.
sor has been defined as the sum of an effective diffusivity tensor The averaging process, as reported in [21], commences with the
and a hydrodynamic dispersion tensor, with both contributions be- application of the superficial averaging operator to the pore-scale
ing dependent on the flow rate. This is inconsistent with the def- equations. While doing this, it is necessary to interchange spatial
inition of the effective diffusivity, which is an intrinsic effective differentiation and integration and this is performed by means of
medium coefficient that only depends of the porous medium ge- the spatial averaging theorem [22]

ometry. Within this context, the purpose of the present work is 1
twofold: firstly, we reformulate the closure problems in a way that ∇ψβ  = ∇ψβ  + nβσ ψβ dA (6)
V Aβσ
they are only written in terms of closure variables, hence providing
a fully upscaled closed form, and secondly, a new formulation of where nβσ is the unit normal vector at Aβσ pointing towards
the total dispersion tensor, involving an effective diffusivity tensor the σ -phase. The introduction of an interfacial integral resulting
and two hydrodynamic dispersion tensors is presented. The paper from the application of the spatial averaging theorem allows the
is organized as follows. In Section 2, we provide some generalities interfacial boundary conditions of the pore-scale problem to be
about the upscaling process using the method of volume averag- substituted when appropriate. However, at this stage, the result-
ing and directly show the structure of the averaged model and the ing averaged equations contain both microscopic (ψ β ) and macro-
ancillary closure problems for mass and momentum transport, the scopic (ψ β ) quantities. To progress towards a model involving
latter corresponding to the inertial regime. In Section 3, we refor- macroscale variables only, it is necessary to express pointwise
mulate the effective medium coefficients and the closure problems quantities in terms of their intrinsic averages and spatial deviations
to meet our goals. Symmetry of the total dispersion tensor is then as proposed in [4]
investigated. Results of the closure problems, which were solved in ψβ = ψβ β + ψ˜ β (7)
unit cells of two-dimensional model porous media for several flow
rates and orientations are reported in Section 4. Finally, the ensu- To carry out the development up to this stage where macro-
ing conclusions are presented in Section 5. scopic equations are yet unclosed, it is not necessary to impose
any length-scale constraint, as explained by Wood and Valdés-
2. Averaging Parada [23]. Thus, the average model contains practically the same
amount of information as its pore-scale counterpart. The system-
In order to upscale the pore-scale governing equations by atic use of a set of time and length-scale constraints and assump-
means of the volume averaging method, it is necessary to define tions in the form of scaling postulates is the essence of the upscal-
an averaging domain V of norm V that contains both solid and ing procedure and allows filtering out unnecessary information at
fluid phases such as the one sketched in Fig. 1. In terms of this av- the macroscopic level. A careful and detailed explanation of the
eraging domain, let us define the superficial and intrinsic averaging imposition of these postulates is available in ref. [23]. For the pur-
operators as [21] poses of this work, it suffices to summarize that the characteristic

1 size of the averaging domain, r0 , is usually taken to be much larger
ψβ  = ψβ dV (5a) than the characteristic size of the pores (β ) and, simultaneously,
V Vβ
 it must be much smaller than the characteristic length associated
1
ψβ β = ψβ dV (5b) to the macroscale (L), i.e. [21]
Vβ Vβ
β  r0  L (8)
with ψ β being a piece-wise smooth function defined in the β -
phase occupying the region Vβ of norm Vβ within V . These two In addition, in many transport processes, there is a disparity be-
averaging operators are related by tween the characteristic time scales associated to pore-scale trans-
∗ ) and to macroscopic transport (say t ∗
ψβ  = εβ ψβ β (5c) port (say tψ
β β ) so that
ψβ 
F.J. Valdés-Parada et al. / Advances in Water Resources 90 (2016) 70–82 73

∗  t∗
tψ . Under these circumstances, averaged properties can As explained above, in order to arrive at the form of the closure
β ψβ β
problem reported in Eqs. (13), one also has to satisfy the time-scale
be treated as if they were uniform, both in space and time, within
constraint expressed by t ≥ t0 with
the integration domains, equal to their values at the centroid. This
is supported by the fact that non-local terms resulting from Taylor Dβ t 0
expansions of the averaged quantities away from the centroid can 1 (14)
2β
be neglected as far as the distances over which their successive
gradients experience significant variations are much larger than r0 as reported in [21], section 3.3.4.
(see [21], section 1.4.3 for instance). A direct corollary of this is the For momentum transport, the resulting equation from the up-
following average constraint for the deviations fields scaling process is the Darcy–Forchheimer equation [24]

ψ˜ β β = 0 (9) vβ  = −



· ∇ pβ β − Fβ · vβ  (15)
μβ
The final step of the volume averaging method is to derive ex-
pressions that relate ψ˜ β with ψ β β and/or its derivatives, i.e., to with Kβ and Fβ being the intrinsic permeability tensor and the
close the macroscale model. The closure process can be summa- Forchheimer correction tensor, respectively. Note that this expres-
rized by the three following main steps: sion does not include the Brinkman correction term and it is thus
constrained to the homogeneous portions of the system where this
• derive the governing equations for the deviations by subtracting term can be shown to be negligible. Certainly, the above expres-
the unclosed averaged equations from the pore-scale equations; sion can be rearranged in a more compact form that resembles to
• simplify the problem by imposing reasonable constraints and Darcy’s law:
assumptions on the basis of orders of magnitude analyzes (i.e.,

scaling postulates). In particular, when the space and time con- vβ  = − · ∇ pβ β (16)
straints mentioned above are satisfied, a quasi-stationary clo- μβ
sure problem can be obtained; where the tensor Hβ may be regarded as an apparent permeability
• formally solve the closure problem in terms of average quanti- tensor, which is defined as
ties in simplified but still representative domains that represent
β = Kβ · (I + Fβ )
H−1 −1
the essential pore-scale geometry such as a periodic unit cell. (17)
This last step can be performed by using Green’s functions as
In order to compute this tensor, it is necessary to solve the follow-
detailed by Wood and Valdés-Parada [23].
ing closure problem [25]:
In this way, the spatial deviations can be related to average ∇ · Eβ = 0, in Vβ (18a)
quantities and their derivatives by means of a linear superposition
in terms of closure variables that can be shown to be integrals of  
ρβ v β
the associated Green’s functions. It is worth mentioning that with- · ∇ Eβ = −∇ eβ + ∇ 2 Eβ + I, in Vβ (18b)
out the separation of time and length scale constraints indicated μβ
above, supporting the assumption that average quantities can be
Eβ = 0, at Aβσ (18c)
regarded as constants within the integration domain, it would not
be possible to express the spatial deviations in the form of the su-
perposition mentioned above. Eβ (r + li ) = Eβ (r ), i = 1, 2, 3 (18d)
These are the essential elements of the application of the vol-
ume averaging method to the pore-scale equations. A detailed de-
eβ ( r + li ) = eβ ( r ), i = 1, 2, 3 (18e)
scription of the derivation of the upscaled model for mass and mo-
mentum transport is available elsewhere [cf. 8,24,21]. The resulting
macroscopic mass transport model, for ε β not necessarily constant, eβ β = 0 (18f)
can be recalled as
∂cAβ β Eβ  = Hβ (18g)
εβ + ∇ · (εβ vβ β cAβ β ) = ∇ · (εβ D∗β · ∇cAβ β ) (10)
∂t where the vector eβ and the tensor Eβ are the closure variables
In Eq. (10), the total dispersion tensor D∗β is defined by that are defined by [24]
   p˜ β =εβ μβ eβ · (Hβ )−1 · vβ β = −eβ · ∇ pβ β (19a)
1
D∗β = D∗e f f + Dβ = Dβ I + nβσ fβ dA − v˜ β fβ β (11)
Vβ Aβσ   
   −Dβ v˜ β = (εβ Eβ · (Hβ )−1 − I ) · vβ β =
1
(−Eβ + Eβ β ) ·
D∗e f f μβ
Here D∗e f f and Dβ denote an effective diffusivity tensor and a hy- 1
∇ pβ β = − ˜ β · ∇ pβ β
E (19b)
drodynamic dispersion tensor, respectively, while fβ is a closure μβ
variable vector defined by ˜ β = Eβ − Eβ β .
with the obvious notation E
c˜Aβ = fβ · ∇cAβ β (12) As mentioned above, the domain in which the closure prob-
lems are solved is usually a periodic unit cell (of period li in the
that solves the following boundary-value problem
direction i = 1, 2, 3) that captures the essential features of the
v˜ β + vβ · ∇ fβ =Dβ ∇ 2 fβ , in Vβ (13a) microscopic geometry, flow and mass transport, such as the one
sketched in Fig. 2 [15].
−nβσ · ∇ fβ =nβσ , at Aβσ (13b) Directing the attention to the closure problem for mass trans-
port, it is worth noting that fβ depends on Dβ , the fluid velocity
fβ ( r + li ) = fβ ( r ), i = 1, 2, 3 (13c) within the pores, vβ , its deviations, and of the unit cell geometry.
Similarly, the closure problem for momentum in Eqs. (18) depends
fβ β =0 (13d) on the cell geometry, vβ and the fluid properties. Once the closure
74 F.J. Valdés-Parada et al. / Advances in Water Resources 90 (2016) 70–82

−nβσ · ∇ f0 = nβσ , at Aβσ (20b)

f0 ( r + li ) = f0 ( r ), i = 1, 2, 3 (20c)

f0 β = 0 (20d)

Here, f0 = fβ is intrinsic, i.e. is only a function of the unit
v β =0
cell geometry since it does not depend on Dβ nor vβ . This closure
problem corresponds to the one given by Eqs. (1.4–58) in [21], the
solution of which allows computing the intrinsic effective diffusiv-
ity tensor, De f f given by [see Eq. (1.4–62) in 21]
  
1
De f f = Dβ I + nβσ f0 dA (21)
Vβ Aβσ

Fig. 2. Sketch of a periodic representation of the porous medium geometry consist- In this way, after a simple algebraic manipulation, we may write
ing of an in-line cylinders array with square cross-section and periodic unit cell of
size .
Eq. (11) as follows


problem for fβ is solved, the fields are substituted into the integral D∗β = De f f + Dh = De f f + nβσ (fβ − f0 )dA − v˜ β fβ β
Vβ Aβσ
of Eq. (11) to compute the total dispersion tensor. These integrals
(22)
play the role of filters of redundant information coming from the
closure problem (see [21]). In this way, the surface integral term The last two terms in the above expression define the true hy-
in Eq. (11) is regarded as a diffusive filter and it is contained in the drodynamic dispersion tensor, Dh . Certainly, for highly convective
effective diffusivity definition, whereas the volumetric integral in D

conditions, i.e., for conditions in which V β A nβσ fβ − f0 dA 
the third term is a convective filter and it constitutes the hydrody- β βσ
namic dispersion term. fβ v˜ β β , one expects Dh

D∗β . Alternatively, when convec-
The structure of Eq. (11) nicely indicates that total dispersion is tive effects are insignificant, D∗β
D∗e f f
De f f Dh
Dβ . How-
the sum of the effective diffusivity and hydrodynamic dispersion. ever, for cases in which convective mass transfer is comparable
However, as noted recently [26, see Section 6 therein], the diffu- to or greater than diffusive transfer, one may expect different pre-
sive filter in Eq. (11) allows passing some convective information, dictions for Dh and Dβ (as well as D∗e f f being significantly differ-
(fβ is, in general, a function of the fluid velocity), so that D∗e f f ent from De f f ). This has been discussed in [26, see section 6]. As
changes with the fluid velocity, which is unexpected for a diffu- will be shown below, under these conditions, transverse dispersion
sion coefficient. Furthermore, the fact that the closure problems’ may be larger than longitudinal dispersion.
solutions require knowledge of the pointwise velocity field, seems At this point, it may appear that once the flow problem is
inconsistent with the philosophy of the volume averaging method, solved over the unit cell, two closure problems need to be solved
in which the computation of effective medium coefficients requires in order to compute D∗β . Actually, it is only necessary to solve the
solving closure problems and not problems for the pointwise phys- closure problem given by Eqs. (13), firstly under non-convective
ical variables. Indeed, under their current forms, boundary value conditions and secondly for the desired flow condition. Let us now
problems in Eqs. (13) and (18) appear to remain unclosed. From a rewrite the closure problems and effective medium coefficients for
practical point of view, this means that the determination of D∗β both momentum and mass transport in such a way that they do
requires a prior computation of the pore scale flow problem over not require prior solution of the pore-scale flow problem.
the unit cell (see Section 2 in [8]) for any combination of flow
and/or mass transport regimes. Therefore, an alternative formula-
3.2. Dispersion tensor determination from closure problems
tion that does not involve the pore scale velocity is highly desir-
able. In the following, these issues are carefully addressed so that
We start with the reformulation of the closure problem for mo-
the effective diffusivity is defined as an intrinsic coefficient and the
mentum given in Eqs. (18). Using the expressions for v˜ β , vβ 
closure problems are written only in terms of closure variables.
and Hβ given in Eqs. (19b), (16) and (18g), yields vβ = − μ1 Eβ ·
β
3. Effective coefficients and closure problems ∇ pβ β . As a consequence, the momentum-like Eq. (18b) can be
rewritten as
In this section, a new formulation of the dispersion tensor in ρβ
terms of the intrinsic effective diffusivity tensor and the hydro- − ∇  pβ β · ETβ · ∇ Eβ = −∇ eβ + ∇ 2 Eβ + I (23)
dynamic dispersion is presented. Furthermore, in Section 3.2, the
μ2β
associated closure problems for mass and momentum transport, It is worth noting that, in the classical formulation, the pore-scale
including both viscous and inertial effects, are reformulated using velocity must be solved every time the macroscopic pressure gra-
closure variables only. A symmetry analysis on D∗β is provided in dient changes, and in the current formulation, the link between
Section 3.3. Finally, simplifications in the case of creeping flow are vβ and ∇pβ β is the closure variable, Eβ which is finally inte-
presented in Section 3.4. grated in the closure problem. In this way, one needs to specify
the macroscopic pressure gradient applied on the unit cell instead
3.1. Reformulation of the dispersion tensor
of providing the microscale velocity field.
Turning our attention to the mass transport closure problem,
Let us start our derivations by noting that, in the absence of
Eq. (13a) can be similarly reformulated as
convection, Eqs. (13) reduce to
−μ−1 β β
β (Eβ · ∇ pβ  + Eβ · ∇ pβ  · ∇ fβ ) = Dβ ∇ fβ
˜ 2
(24)
∇ 2 f0 = 0, in Vβ (20a)
F.J. Valdés-Parada et al. / Advances in Water Resources 90 (2016) 70–82 75

  
In the same way, the hydrodynamic dispersion tensor takes the 1
form D∗β = Dβ I + nβσ f∗0 dA∗ De f f
Vβ∗ Aβσ
   
Dβ β β
Dh = nβσ (fβ − f0 )dA + μ−1
β ∇ pβ  · Eβ fβ 
T
(25) effective diffusivity
Vβ Aβσ   
1 Pe ∗ β
+ Dβ nβσ f∗β − f∗0 dA∗ + λ · E∗T
β fβ  Dh
While deriving this last expression, we have made use of the Vβ∗ Aβσ E∗β  · λ
fact that, due to the constraint in (8), ∇pβ β can be assumed con-   
stant within the averaging domain and we have also considered hydrodynamic dispersion

Eq. (13d) which implies E ˜ Tβ fβ β = ET fβ β . (29)


β
A convenient dimensionless form, well adapted for numerical
where f∗0
is the solution of Eqs. (27) when Pe = 0.
solution, may now be used based on the following dimensionless
At this point of the analysis, the following comments are in or-
quantities: Vβ∗ = Vβ /3 , A∗βσ = Aβσ /2 , f∗β = fβ /, f∗0 = f0 /, e∗β =
der:
eβ / and E∗β = Eβ /2 , where  is the characteristic size of the unit
• The effective diffusion coefficient, De f f , is only a function of the
cell (see Fig. 2). In addition, we introduce λ = ∇ pβ β /∇ pβ β 
porous medium microstructure and it is no longer affected by
the unit vector in the direction of the average pressure gradient.
convective effects. Although the diagonal terms of this tensor
Using the same symbol ∇ for the nabla operator with or without
remain always positive, the surface integral part is a negative
dimension (the latter being the product of the former by ) yields
definite function that reflects the influence of the pore scale ge-
the following dimensionless closure problem for momentum trans-
ometry.
port • The components of the hydrodynamic dispersion tensor, Dh ,
∇ · Eβ = 0, in

Vβ (26a) only depend on: (1) The porous medium microstructure, (2)
The cell Péclet number, Pe, (3) The Reynolds number, Re∗ and
(4) The macroscopic pressure gradient orientation defined by
−Re∗ λ · E∗T
β · ∇ Eβ = −∇ eβ + ∇ Eβ + I,
∗ ∗ 2 ∗
in Vβ (26b) λ.
• Closure problems, as expressed in Eqs. (26) and (27), are cer-
tainly different from their original formulations since they no
E∗β = 0, at Aβσ (26c) longer require solving the pore scale flow field. In fact, for a
given Re∗ and λ, the closure problem for total mass and mo-
mentum transport can be solved once in order to obtain the
E∗β (r + li ) = E∗β (r ), i = 1, 2, 3 (26d) fields of E∗β so that Dh can then be computed for any required
value of the Péclet number after solving the closure problem
e∗β (r + li ) = e∗β (r ), i = 1, 2, 3 (26e) for the species mass transport.

The forms of the closure problems given by Eqs. (26) and (27)
are convenient if one is willing to compute D∗β for a prescribed
e∗β β = 0 (26f)
macroscopic pressure gradient applied to the structure, as reflected
from which the dimensionless apparent permeability can be de- in the expression of Eq. (29). In some circumstances, it might also
duced as H∗β = Hβ /2 = E∗β . In Eq. (26b), Re∗ is the Reynolds be of interest to determine D∗β for a specific mean flow direction
number defined as Re∗ = ρβ 3 ∇ pβ β /μ2β . It must be noted that that does not necessarily coincide with the corresponding pressure
gradient. To do so, it is convenient to rewrite Eqs. (26b), (27a) and
λ and Re∗ are the two parameters allowing to specify the pressure
(29) in terms of the unit vector λv in the direction of vβ , which
gradient direction and intensity, respectively.
is defined as
Accordingly, the dimensionless closure problem for mass trans-
port may be expressed as vβ  H∗β · λ
λv = =− ∗ (30)
Pe
 vβ  Hβ · λ
∗T
− λ · E˜ β + E∗T
β · ∇ fβ = ∇ fβ
∗ 2 ∗
(27a)
E∗β  · λ When this is done, the momentum-like equation in the closure
problem (26) takes the form
−nβσ · ∇ f∗β = nβσ , λv · E∗T
β 
at Aβσ (27b) −1
Re∗ · E∗T · ∇ E∗β = −∇ e∗β + ∇ 2 E∗β + I (31)
Eβ  · λv  β
∗ −1

f∗β (r + li ) = f∗β (r ), i = 1, 2, 3 (27c) while the transport equation in the closure problem (27) is given
by

f∗β β = 0 (27d) Peλv · E∗T ˜ ∗T
β  · Eβ + Eβ · ∇ fβ = ∇ 2 f∗β
−1 ∗T ∗
(32)
Note that in Eq. (27a), the fields of Eβ are now required in or-
This yields the following expression for D∗β
der to solve the closure problem for mass transport. They are ob-
tained by solving Eqs. (26), which, contrary to the original formula- D∗β 
1
= I+ n f∗ dA∗
tion [24], do not require the solution of the pore scale velocity and Dβ Vβ∗ Aβσ βσ 0
can now be solved as an independent boundary-value problem. In 
Eq. (27a), Pe is the cell Péclet number defined as 1 ∗T ∗ β
+ ∗ n (f∗ − f∗0 )dA∗ − Peλv · E∗T
β  · Eβ fβ  (33)
−1
Vβ Aβσ βσ β
εβ vβ β 
Pe = (28) Eqs. (31) and (32) hold a stronger degree of non linearity than

the versions written in terms of λ and certainly feature closure
In this way, the total dispersion tensor can be expressed as problems that are more complex to solve, although in Eq. (31),
76 F.J. Valdés-Parada et al. / Advances in Water Resources 90 (2016) 70–82

the prefactor 1/E∗β −1 · λv  might be lumped together with Re∗ Dh . More specifically, a comparison between the expressions of D∗β
to form a cell Reynolds number given by Re = Re∗ /E∗ −1
β · λv  = in Eqs. (29) and (38), taking Eq. (37) into account, indicates that
ρ v  asymmetry may originate from the volume average and/or the area
Re∗ E∗ β  · λ = β μβ . integral parts of Dh in the rhs of Eq. (29).
β
While Re∗ and Re conveniently measure the inertial to vis- A quick order of magnitude analysis to the formal solution of
cous forces ratio at the scale, , of the unit cell, for a pre- the closure problem given by Eqs. (13) indicates that fβ = O β
scribed pressure gradient and flow rate respectively, other def- and hence that the second term in the rhs of Eq. (38) is O(1). Con-
initions may be used when a different scaling is employed.
√ In sequently, a sufficient condition for D∗β to be quasi-symmetric is
fact, one might use a Reynolds number based on k, k being when
the magnitude of the intrinsic permeability tensor for an isotropic  
Pe ∗ ∗ β
medium [25,27], or, more commonly, a pore-scale Reynolds num- O λ · E∗T
β · ∇ f β + 2 I f β  1 (39)
ρβ vβ β β β E∗β ·λ E∗β  · λ
ber, Re p = μβ = Re∗ εβ  . Accordingly, a pore Péclet
β
vβ  β  Since ∇ f∗β can be estimated to be O(1), the above constraint for
number Pe p = Dβ = Pe ε β is often employed [15,28] and, for D∗β to be quasi-symmetric is finally
β
the sake of consistency with data reported in the literature, our  
results in Section 4 will be presented in terms of Rep and Pep . εβ 
Pe  O (40)

3.3. Symmetry properties of the dispersion tensor
or, equivalently, in terms of the pore Péclet number, Pe p =
vβ β β
Our purpose in this section is to investigate the symmetry prop- Dβ
erties of D∗β . The starting point of the development is to form the
dyadic product of Eq. (27a) with f∗β which gives Pe p  1 (41)

∗ ∗ 2 ∗ ∗ in agreement with some investigations reported in [29]. The suffi-


Pe β ∗
λ · E∗T
β  f β − Eβ · ∇ f β + I f β = ∇ f β f β
∗T
(34) cient constraint given above is independent of the Reynolds num-
E∗β  · λ ber, and this is an important feature of the symmetry properties
The right hand side (rhs) of Eq. (34) can be rewritten in an equiv- of the dispersion tensor that requires a closer attention regard-
alent form as ing the compatibility of symmetry with the existence of inertial
T effects. Since inertia becomes significant when Rep ࣡ 10, the suffi-
∇ 2 f∗β f∗β = ∇ · ∇ f∗β f∗β − ∇ f∗β · ∇ f∗β (35) cient condition (41) for symmetry would require, in that case, that
the Schmidt number, Sc, is such that
and while taking the dimensionless superficial average of Eq. (34)
in which the rhs is replaced by its expression in Eq. (35), we ob- Pe p μβ
Sc = =  0.1 (42)
tain Re p ρβ D β
Pe β ∗
∗ ∗ ∗ ∗ This condition is never met for conventional fluids as Sc = O(1 ) for
λ · E∗T
β   f β  −  Eβ · ∇ f β + I f β  = ∇ ·  ∇ f β f β 
∗T
E∗β  · λ gases while, for liquids, it is rather O(103 ). For porous structures in
 T which the unit cell does not possess any specific geometrical sym-
1
+ nβσ · ∇ f∗β f∗β dA∗ −  ∇ f∗β · ∇ f∗β  (36) metry or when the mean flow is not along a symmetry axis, this
V∗ Aβσ
suggests that inertia is a potential mechanism that can certainly
Due to periodicity, the first term in the rhs of this last expression is trigger asymmetry of D∗β .
such that ∇ · (∇ f∗β )f∗β  = 0 while the second term in the rhs can Finally, one should note that the condition given in (41) is inde-

be re-written as V1∗ A nβσ · (∇ f∗β )f∗β dA∗ = − V1∗ A nβσ f∗β dA∗ pendent from (but compatible with) the time-scale constraint ex-
βσ βσ
pressed in (14) that is required to treat the mass transport closure
upon making use of the boundary condition of Eq. (27b). More-
problem as quasi-steady.
over, because of the zero average constraint of f∗β (see Eq (27d)),
the first term in the left hand side of Eq. (36) is also zero, so that 3.4. Creeping flow regime
this equation can be written as

1 Pe ∗ β Special attention should be dedicated to situations in which the
nβσ f∗β dA∗ − λ · E∗T
β fβ 
Vβ∗ Aβσ E∗β  · λ Reynolds number is such that the flow remains in the creeping
regime, i.e. typically when Rep  10. In such circumstances, the
Pe ∗ ∗ β ∗ T
= λ · E∗T
β · ∇ fβ fβ  −  ∇ fβ · ∇ f∗β β (37) closure problem for momentum transport does not depend on the
E∗β  · λ Reynolds number nor on the macroscopic pressure gradient orien-
Dβ∗ tation and is hence intrinsic, yielding Hβ = Kβ . As a consequence,
When this last result is introduced back into the expression of Dβ the computation of D∗β requires the momentum closure problem
given by Eq. (29), we have to be solved only once and the solution can be used for any Péclet

number value and λ.
Dβ Pe
= I − (∇ f∗β )T · ∇ f∗β β + λ · E∗T ∗ β
β · ( ∇ fβ + 2I )fβ 
∗ In the following sections, some computational results are pro-
Dβ E∗β  · λ vided in both the inertial and creeping flow regimes.
(38)
4. Results
While the first two tensors in the rhs of Eq. (38) are symmetric, the
last term is not symmetric, except when the unit cell is symmet- The dimensionless closure problems given by Eqs. (26) and
ric and λ is along one of its symmetry axis. This features a non- (27) were solved on periodic two-dimensional unit cells of model
symmetric D∗β in general. Moreover, by taking Pe = 0, it straight- porous media using the finite element software Comsol Multi-
forwardly follows from Eq. (38) that De f f is a symmetric tensor. physics 4.4 involving sufficient mesh elements to guarantee con-
As a consequence, it is clear that asymmetry may only occur in sistency in the numerical results. A mesh-convergence analysis
F.J. Valdés-Parada et al. / Advances in Water Resources 90 (2016) 70–82 77

showed that a dimensionless grid-block area of about 4 × 10−5 = 0 (see [25]) and we denote by θ v the macroscopic flow inclina-
along with a dimensionless boundary element size of 0.01 were tion on the horizontal x-direction. As a matter of fact, θ = θv for
appropriate to achieve convergent numerical results. The solid θ = 0 and π /4; however, for θ = π /8 we obtained that θv = π /7
phase was considered as parallel cylinders of square cross sections for εβ = 0.4 and θ v ≈ 0.1521 for εβ = 0.8, in both cases Re p = 200.
arranged on a square regular periodic pattern (see Fig. 2). Two dif- The longitudinal (along λv ) and transverse (along τ v ) compo-
ferent values of the porosity, namely εβ = 0.8 and εβ = 0.4, were nents of the total dispersion tensor can be obtained from the com-
considered. ponents of D∗β in (ex , ey ) as follows,
Dispersion is studied in the plane orthogonal to cylinders axes
D∗β ,λv λv = cos2 θv D∗β ,xx + sin θv D∗β ,yy + sin θv cos θv (D∗β ,yx + D∗β ,xy )
2
as a function of Pep , without inertia (Re p = 0) and for Re p = 200
while considering θ = 0, π /8 and π /4. For the pore-length β in (43a)
the definition of Rep and Pep a quantitative estimate, based on the
hydraulic diameter and already proposed in [9] (see also [21] p.

Dβ ,τv τv = sin
2
θv Dβ ,xx + cos θv Dβ ,yy − sin θv cos θv (Dβ ,yx + Dβ ,xy )
∗ 2 ∗ ∗ ∗

ε (43b)
144), is employed as β ∼ σ 1−βε , σ being the cylinder size (see
β
E∗β ·λσ When mass transport is strongly convection-dominated (i.e.
Fig. 2). This yields Re p = Re∗ (1−ε ) and Pe p = Pe (1−εσ ) when the Péclet number is large compared to unity), D∗β essen-
β β
Although closure problems are solved for a prescribed pressure tially depends on the volume average part of the hydrodynamic
gradient direction defined by λ, inclined of an angle θ on ex , it dispersion represented by the last term in Eq. (29) (or (33)). Ex-
is more physically appealing to present the results in terms of amples of the fields of the λv λv component of the corresponding
the orientation of the average velocity, i.e. in the longitudinal and λ·E∗T f∗
tensor E∗β·λ
β
are reported over a unit cell in Figs. 3 and 4 for
transverse directions λv (see Eq. (30)) and τ v , τ v being the unit β
vector directly orthogonal to λv . Note that, even for porous mate- Pe p = 103 , the three flow angles and the two Reynolds numbers
rials such that Kβ is a spherical tensor, as for the structure under Re p = 0 and 200. Clearly, the flow orientation and the inertial ef-
consideration here, λ and λv are not necessarily aligned when Rep fects have a determinant effect on these fields and consequently

Fig. 3. Examples of the fields of the λv λv component of the tensor λ · E∗T β fβ /Hβ · λ and velocity streamlines for three flow angles taking εβ = 0.4 and Pe p = 10 0 0 for (a)

θ = 0, Re p = 0; (b) θ = π /8, Re p = 0; (c) θ = π /4, Re p = 0; (i) θ = 0, Re p = 200; (ii) θ = π /8, Re p = 200; (iii) θ = π /4, Re p = 200.
78 F.J. Valdés-Parada et al. / Advances in Water Resources 90 (2016) 70–82

Fig. 4. Examples of the fields of the λv λv component of the tensor λ · E∗T β fβ /Hβ · λ and velocity streamlines for three flow angles taking εβ = 0.8 and Pe p = 10 0 0 for (a)

θ = 0, Re p = 0; (b) θ = π /8, Re p = 0; (c) θ = π /4, Re p = 0; (i) θ = 0, Re p = 200; (ii) θ = π /8, Re p = 200; (iii) θ = π /4, Re p = 200.

on the components of the total dispersion tensor as will be shown D∗β ,λv λv increases in the presence of inertia when the flow is
below. Although strongly correlated to the flow structure as in- not aligned with the principal axes (see Fig. 5b and c) but de-
dicated by the flow streamlines that are superimposed in Figs. 3 creases with Rep when θ = 0 (Fig. 5a). This surprising behavior
and 4, the dependence of the fields on Rep and θ is complex as contrasts with the case εβ = 0.8 for the same range of Pep and
their magnitude varies non-monotonically with θ and increases or the same flow angle (see Fig. 5i). A physical explanation can be
decreases with Rep . This is reflected on the graphs of Figs. 5 and given from the fields in Figs. 3a, 3i and 4a, 4i). On the one hand,
6 where the dimensionless longitudinal and transverse dispersion for θ = 0, Re p = 0 and εβ = 0.4, the vertical gaps between two
coefficients are represented versus Pep for the different values of adjacent cylinders in the direction of the flow are occupied by
θ and Rep . Regarding these results, the following comments are in vortices taking place at small velocities. On the other hand, for
order: εβ = 0.8 and the same flow conditions, only two small vortices
are present in these regions that are mainly filled with tortuous
1. For Pep ≤ 1, all the predictions of D∗β converge to the same flow streamlines connected from the entrance to the exit of the
value that is given by the effective diffusivity, which is in- unit cell. When Re p = 200 and the same flow angle, the vor-
sensitive to inertial effects and to the flow angle. As a mat- tices pattern is not significantly modified in the unit cell with
ter of fact, under these conditions, D∗β reduces to De f f , which εβ = 0.4 (only two additional vortices are produced and veloc-
only requires solving the closure problem for f0 , given by ities remain extremely small) whereas streamlines are signifi-
Eqs. (20). cantly straightened in the horizontal channels in the vicinity of
2. For Pep > 1, the longitudinal dispersion increases when Rep in- the vertical gaps. The conjunction of these two coupled mecha-
creases, whatever the flow angle, for a sufficiently large poros- nisms yields a less efficient dispersion and a decrease of D∗β ,λv λv
ity (see Fig. 5i–iii for εβ = 0.8). However, for a smaller poros- in the presence of inertia for θ = 0 and εβ = 0.4. Conversely, for
ity (εβ = 0.4, see (Fig. 5a–c), the longitudinal dispersion may εβ = 0.8, when Re p = 200 (θ = 0), large eddies with significant
decrease or increase with Rep . In particular, for Pep ࣡ 100, velocities are taking place in the vertical gaps. Although they
F.J. Valdés-Parada et al. / Advances in Water Resources 90 (2016) 70–82 79

εβ = 0.4 εβ = 0.8
105 105
Rep = 0 Rep = 0
Rep = 200 Rep = 200
104 104

103 103
∗ ∗
Dβ,λv λv
Dβ,λv λv
102 102
Dβ Dβ
101 101

100 100
a) θ = 0 i) θ = 0
10−1 10−1
10−1 100 101 102 103 10−1 100 101 102 103
P ep P ep
105 105
Rep = 0 Rep = 0
Rep = 200 Rep = 200
104 104

103 103
∗ ∗
Dβ,λv λv
Dβ,λv λv
102 102
Dβ Dβ
101 101

100 π 100 π
b) θ = 8
ii) θ = 8
10−1 10−1
10−1 100 101 102 103 10−1 100 101 102 103
P ep P ep
105 105
Rep = 0 Rep = 0
Rep = 200 Rep = 200
104 104

103 103
∗ ∗
Dβ,λv λv
Dβ,λv λv
102 102
Dβ Dβ
101 101

100 π 100 π
c) θ = 4
iii) θ = 4
−1 −1
10 10
10−1 100 101 102 103 10−1 100 101 102 103
P ep P ep
Fig. 5. Longitudinal component of the total dispersion tensor vs. Pep taking three different values of θ . Results are obtained from solving the associated closure problems in
a 2D unit cell with the solid phase modeled as a square having porosities of 0.4 and 0.8. Black curves correspond to Re p = 0 and blue curves correspond to Re p = 200. (For
interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

induce straighter streamlines in the horizontal channels where In Fig. 7, we have reported the ratio D∗β ,λv λv /D∗β ,τv τv versus Pep
spreading is however favored, the net result is a significant im- for the values of ε β , θ and Rep under consideration. As expected,
provement of dispersion and hence a larger value of D∗β ,λv λv . for Pep ≤ 1, mass transport is mainly driven by diffusion and
These last two mechanisms, i.e. development of eddies with sig- consequently D∗β ,λv λv = D∗β ,τv τv . Interestingly, for 1 ࣠ Pep ࣠ 100
nificant velocities allowing the development of spreading, are and θ = π /4, the transverse component is larger than its longitu-
the main ones explaining the increase of the longitudinal dis- dinal counterpart, except for εβ = 0.8, Re p = 200. This observation
persion with Rep for all other flow angles and the two values of is consistent with numerical results from Salles et al. [28] (see
εβ . Table 5 therein). However, this effect is no longer present when
3. Keeping all other parameters the same, the longitudinal disper- flow is along the principal axes of the structure and for a large
sion coefficient increases when porosity decreases. Moreover, value of the porosity when inertia is significant. Certainly, for
the exponent of the power-law dependence of D∗β ,λv λv on Pep , Pep larger than ∼ 100, the longitudinal component is larger than
occurring at sufficiently large values of the Péclet number, is the transverse component, regardless of θ and the flow regime.
not modified by the presence of inertia, which however favors Under these conditions, the longitudinal to transverse dispersion
its emergence at smaller Pep . coefficients ratio corresponding to θ = π /4 is larger than the
4. The transverse dispersion coefficient also exhibits a complex values corresponding to θ = π /8 for cases involving (or not)
dependence upon Pep , θ and Rep . As for longitudinal dispersion, inertial effects and for the two values of ε β . These observations
the transverse coefficient increases when porosity decreases for are in agreement with a previous study [see, Figs. 2 and 3 in ref.
a given Pep , θ and Rep . 15].
80 F.J. Valdés-Parada et al. / Advances in Water Resources 90 (2016) 70–82

εβ = 0.4 εβ = 0.8
Rep = 0 Rep = 0
Rep = 200 Rep = 200
100.05
100

∗ ∗
Dβ,τv τv
Dβ,τv τv 10
0

D β 10−0.1 Dβ
10−0.05

10−0.2 a) θ = 0 i) θ = 0
10−0.1
10−1 100 101 102 103 10−1 100 101 102 103
P ep P ep
Rep = 0 Rep = 0
Rep = 200
100.4 Rep = 200
101

∗ ∗
Dβ,τv τv
Dβ,τv τv
100.2
Dβ Dβ

100 100
π π
b) θ = 8
ii) θ = 8

10−1 100 101 102 103 10−1 100 101 102 103
P ep P ep
Rep = 0 Rep = 0
101 Rep = 200 Rep = 200

∗ ∗
100.2
Dβ,τv τv
Dβ,τv τv

Dβ Dβ

100 100
π π
c) θ = 4
iii) θ = 4

10−1 100 101 102 103 10−1 100 101 102 103
P ep P ep
Fig. 6. Transverse component of the total dispersion tensor vs. Pep taking three different values of θ . Results are obtained from solving the associated closure problems in
a 2D unit cell with the solid phase modeled as a square having porosities of 0.4 and 0.8. Black curves correspond to Re p = 0 and blue curves correspond to Re p = 200. (For
interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

5. Conclusions flow regime, the closure problem for momentum is intrinsic and
needs to be solved only once. This is a clear difference compared
In this work, the effective diffusivity and hydrodynamic disper- to the existing formulation in which the pore scale flow problem
sion tensors that compose the total dispersion tensor have been was a necessary input for both the momentum and mass trans-
redefined in a volume averaging context. In these reformulations, port closure problems. The benefits of this new formulation are at
the effective diffusivity is no longer a function of the flow and the least twofold. Firstly, it provides a completely closed set of bound-
ancillary closure problem corresponds to the one typically found ary value problems as intended in the volume averaging method.
when studying passive diffusion in porous media. The hydrody- Secondly, since the closure problem for mass transport is now cou-
namic dispersion tensor was shown to depend on the Péclet num- pled to the one for momentum transport, for conditions in which
ber, the Reynolds number, the macroscopic flow angle and the ge- Rep 1, the numerical solution of the latter is significantly less
ometry of the porous medium represented by the unit cell. It was demanding in terms of computational resources in the current for-
shown that the dispersion tensor can be obtained from the so- mulation.
lution of two closure problems, namely, one for momentum and Symmetry of the total dispersion tensor was analyzed and it
one for mass transport. Both closure problems were formulated in was shown that D∗β is not symmetric in the general case. The con-
terms of closure variables only, showing that the pore scale flow dition for D∗β to be quasi-symmetric was investigated and a suf-
fields are not necessary. In this formulation, the mass transport ficient condition is when the pore Péclet number remains small
closure problem needs to be solved twice, firstly with a zero Pé- compared to unity, a situation that is however not compatible with
clet number yielding the effective diffusivity and secondly for the the existence of significant inertial effects for conventional fluids.
desired Péclet number and flow angle. In the latter case, the solu- The ancillary closure problems were solved on periodic unit
tion of the closure problem for momentum, obtained with the re- cells of 2D model structures made of parallel cylinders of square
quired Reynolds number and flow angle, is needed. In the creeping cross section arranged on a square regular pattern. The closure
F.J. Valdés-Parada et al. / Advances in Water Resources 90 (2016) 70–82 81

εβ = 0.4 εβ = 0.8
105 105

104 a) Rep = 0 θ=0 104 i) Rep = 0


θ=0
3 3
10 10
∗ ∗
Dβ,λv λv
Dβ,λv λv
∗ 102 ∗ 102
Dβ,τv τv
Dβ,τv τv
101 π/8 101
π/4

100 100
π/4 π/8
10−1 10−1
10−1 100 101 102 103 10−1 100 101 102 103
P ep P ep
105 105

104 b) Rep = 200 θ=0 104 ii) Rep = 200


θ=0
103 103
∗ ∗
Dβ,λv λv
Dβ,λv λv
∗ 102 ∗ 102
Dβ,τv τv
Dβ,τv τv
101 π/8 101 π/4
π/8
0 0
10 10
π/4
10−1 10−1
10−1 100 101 102 103 10−1 100 101 102 103
P ep P ep
Fig. 7. Dependence of the ratio of the longitudinal and transverse components of the total dispersion tensor with Pep for three flow orientations taking εβ = 0.4 and εβ = 0.8.
The results were obtained from solving the closure problems in unit cells with the solid phase represented by a square obstacle.

problem solutions in the plane orthogonal to the cylinders’ axes number: 12511908; Arrangement number: 112087) for the financial
were used to compute the longitudinal and transverse components aid provided.
of the total dispersion tensor. This was carried out considering
three flow angles, with and without inertial effects. For the mass References
transport and flow conditions under consideration, the dispersion
coefficients are highly sensitive to the flow rate and to the flow [1] Cushman JH, Bennethum LS, Hu BX. A primer on upscaling tools for porous
media. Adv Water Resour 2002;25(8–12):1043–67. http://dx.doi.org/10.1016/
angle. Furthermore, it was shown that, for small enough values of
S0309-1708(02)0 0 047-7.
the porosity, inertial effects may lead to a smaller longitudinal dis- [2] Sahimi M. Flow and transport in porous media and fractured rock. Weinheim,
persion coefficient compared to that in creeping flow conditions, Germany: Wiley; 2011. http://dx.doi.org/10.1002/9783527636693.2nd ed.
specifically when flow is along a principal axis of the structure. For [3] Whitaker S. Diffusion and dispersion in porous media. AIChE J 1967;13(3):420–
7. http://dx.doi.org/10.1002/aic.690130308.
some particular flow orientations within the structure under con- [4] Gray W. A derivation of the equations for multiphase transport. Chem Eng Sci
sideration (namely θ = π /4), and except for a large value of the 1975;30:229–33. http://dx.doi.org/10.1016/0 0 09-2509(75)80 010-8.
porosity when inertia is significant, the transverse component may [5] Paine M, Carbonell R, Whitaker S. Dispersion in pulsed systems I. Hetero-
geneous reaction and reversible adsorption in capillary tubes. Chem Eng Sci
be larger than the longitudinal dispersion for 1 ࣠ Pep ࣠ 100. For 1983;38:1781–93. http://dx.doi.org/10.1016/0 0 09- 2509(83)85035- 0.
Pep ࣡ 100, the longitudinal component is larger than the trans- [6] Taylor G. Dispersion of soluble matter in solvent flowing slowly through
verse component for any flow condition and porosity. a tube. Proc Roy Soc 1953;219:186–203. http://dx.doi.org/10.1098/rspa.1953.
0139.
In essence, the methodology developed in the present work [7] Aris R. On the dispersion of a solute in a fluid flowing through a tube. Proc
could be extended to broader situations such as heat transfer Roy Soc 1956;235:67–77. http://dx.doi.org/10.1098/rspa.1956.0065.
in homogeneous porous media involving convection in the fluid [8] Carbonell R, Whitaker S. Dispersion in pulsed systems. II: Theoretical
developments for passive dispersion in porous media. Chem Eng Sci
phase as studied in [30] or dispersion in heterogeneous porous 1983;38(11):1795–802. http://dx.doi.org/10.1016/0 0 09- 2509(83)85036- 2.
media investigated in [12], among many others. In both cases, the [9] Eidsath A, Carbonell R, Whitaker S, Herrmann L. Dispersion in pulsed systems.
transport equations that involve diffusion and convective mecha- III: Comparison between theory and experiments for packed beds. Chem Eng
Sci 1983;38(11):1803–16. http://dx.doi.org/10.1016/0 0 09- 2509(83)85037- 4.
nisms are formally the same as Eq. (1). Therefore, it is not hard to
[10] Plumb O, Whitaker S. Dispersion in heterogeneous porous media: 1. Local vol-
realize that the formulation of closure problems provided here can ume averaging and large-scale averaging. Water Resour Res 1988;24(7):913–26.
be straightforwardly applied to these and other similar situations. http://dx.doi.org/10.1029/WR024i0 07p0 0913.
[11] Plumb O, Whitaker S. Dispersion in heterogeneous porous media: 2.
Predictions for stratified and two-dimensional spatially periodic sys-
tems. Water Resour Res 1988;24(7):927–38. http://dx.doi.org/10.1029/
Acknowledgments WR024i0 07p0 0927.
[12] Quintard M, Cherblanc F, Whitaker S. Dispersion in heterogeneous
FJVP is grateful to École CentraleSupélec, Châtenay-Malabry and porous media: One-equation non-equilibrium model. Trans Porous Media
2001;44(1):181–203. http://dx.doi.org/10.1023/A:1010746011296.
to the Ecole Nationale Supérieure des Arts et Métiers for the vis-
[13] Caillabet Y, Fabrie P, Lasseux D, Quintard M. Computation of large-scale param-
iting professor positions. The same author is thankful to Fondo eters for dispersion in fissured porous medium using finite-volume method.
Sectorial de Investigación para la Educación from CONACyT (Project Comput Geosci 2001;5(2):121–50. http://dx.doi.org/10.1023/A:1013140922402.
82 F.J. Valdés-Parada et al. / Advances in Water Resources 90 (2016) 70–82

[14] Wood B, Cherblanc F, Quintard M, Whitaker S. Volume averaging for de- [23] Wood B, Valdés-Parada F. Volume averaging: local and nonlocal closures using
termining the effective dispersion tensor: Closure using periodic unit cells a Green’s function approach. Adv Water Resour 2013;51:139–67.
and comparison with ensemble averaging. Water Resour Res 2003;38(8):1210. [24] Whitaker S. The Forchheimer equation: a theoretical development. Trans
http://dx.doi.org/10.1029/20 02WR0 01723. Porous Media 1996;25:27–61. http://dx.doi.org/10.10 07/BF0 0141261.
[15] Amaral-Souto H, Moyne C. Dispersion in two-dimensional periodic porous me- [25] Lasseux D, Abbasian-Arani AA, Ahmadi A. On the stationary macroscopic iner-
dia. Part II Dispersion tensor. Phys Fluids 1997;9(8):2253–63. http://dx.doi.org/ tial effects for one phase flow in ordered and disordered porous media. Phys
10.1063/1.869365. Fluids 2011;23:073103. http://dx.doi.org/10.1063/1.3615514.
[16] Didierjean S, Souto HPA, Delannay R, Moyne C. Dispersion in periodic porous [26] Valdés-Parada F, Porter M, Wood B. The role of tortuosity in upscaling. Trans
media. Experience versus theory for two-dimensional systems. Chem Eng Sci Porous Media 2011;88:1–30. http://dx.doi.org/10.1007/s11242-010-9613-9.
1997;52(12):1861–74. http://dx.doi.org/10.1016/S0 0 09-2509(96)0 0518-0. [27] Papathanasiou TD, Markicevic B, Dendy ED. A computational evaluation of
[17] Wood B. Inertial effects in dispersion in porous media. Water Resour Res the Ergun and Forchheimer equations for fibrous porous media. Phys Fluids
2007;43. http://dx.doi.org/10.1029/2006WR005790.W12S16– 2001;13(10):2795–804.
[18] Aguilar-Madera C, Baz-Rodríguez S, Ocampo-Pérez R. Effective mass diffusion [28] Salles J, Thovert J-F, Delannay R, Prevors L, Auriault J-L, Adler P. Taylor dis-
and dispersion in random porous media. Can J Chem Eng 2015;93:756–65. persion in porous media. determination of the dispersion tensor. Phys Fluids
http://dx.doi.org/10.1002/cjce.22156. 1992;5:2348. http://dx.doi.org/10.1063/1.858751.
[19] Quintard M, Whitaker S. Convection, dispersion, and interfacial trans- [29] Auriault JL, Moyne C, Souto HPA. On the asymmetry of the dispersion tensor in
port of contaminants: Homogeneous porous media. Adv Water Resour porous media. Transp Porous Media 2010;85:771–83. http://dx.doi.org/10.1007/
1994;17(4):221–39. http://dx.doi.org/10.1016/0309-1708(94)90 0 02-7. s11242- 010- 9591- y.
[20] Ahmadi A, Aigueperse A, Quintard M. Calculation of the effective proper- [30] Quintard M, Kaviany M, Whitaker S. Two-medium treatment of heat transfer
ties describing active dispersion in porous media: from simple to complex in porous media: Numerical results for effective properties. Adv Water Resour
unit cells. Adv Water Resour 2001;24(3–4):423–38. http://dx.doi.org/10.1016/ 1997;20:77–94. http://dx.doi.org/10.1016/S0309-1708(96)0 0 024-3.
S0309-1708(0 0)0 0 065-8. [31] Chella R, Lasseux D, Quintard M. Multiphase, multicomponent fluid flow in
[21] Whitaker S. The method of volume averaging. Kluwer Academic Publishers; homogeneous and heterogeneous porous media. Oil & Gas Science and Tech-
1999. http://dx.doi.org/10.1007/978- 94- 017- 3389- 2. nology - Rev. IFP 1998;53(3):335–46. http://dx.doi.org/10.2516/ogst:1998029.
[22] Howes F, Whitaker S. The spatial averaging theorem revisited. Chem Eng Sci
1985;40:1387–92.

You might also like