Skip to main content
Log in

Impact of Non-uniform Properties on Governing Equations for Fluid Flows in Porous Media

  • Published:
Transport in Porous Media Aims and scope Submit manuscript

Abstract

The macroscopic governing equations of a compressible multicomponents flow with non-uniform viscosity and with mass withdrawal (due to heterogeneous reactions) in a porous medium are developed. The method of volume averaging was used to transform local (or microscopic) governing equations into averaged (or macroscopic) governing equations. The impacts of compressibility, non-uniform viscosity, and mass withdrawal on the form of the averaged equations and on the value of the macroscopic transport coefficients were investigated. The results showed that the averaged mass conservation equation is significantly affected by mass withdrawal when a specific criterion on the size of the domain is respected. The results also showed that the form of the averaged momentum equations is not affected by mass withdrawal, by compressibility effects or by non-uniform viscosity, provided that the Reynolds number at the pore level is small. Nonetheless, the velocity field is affected by the heterogeneous reaction via the averaged mass conservation equation, and also by viscosity variations due to the presence of the volume-averaged viscosity (which value changes with position) in the averaged momentum equations. A new closure variable definition was proposed to formulate the closure problem, which avoided the need to solve an integro-differential equation in the closure problem. This formulation was used to show that the permeability tensor only depends on the geometry of the porous medium. In other words, that tensor is independent on whether the fluid is compressible/incompressible, has uniform/non-uniform viscosities, and whether mass withdrawal due to heterogeneous reactions is present/absent.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

Abbreviations

\(a\) :

Arbitrary scalar field

\(a_\mathrm{V}\) :

Surface area per unit of volume, \(\hbox {m}^{-1}\)

a, b, c :

Arbitrary vector fields

A :

Index of the reacting species in the fluid

B :

Index of a species in the fluid

B :

Closure variable

C :

Arbitrary tensor field

d\(A\) :

Differential area, \(\hbox {m}^{2}\)

\(e^{-}\) :

Electrons

d :

Unknown vector, \(\hbox {s}^{-1}\)

f :

Closure variable, \(\hbox {m}^{-1}\)

\(E_\mathrm{a},\;E_\mathrm{{r,a}}\) :

Potential and reversible potential in anode, V

\(F\) :

Faraday constant, \(\hbox {C mol}^{-1}\)

F :

Closure variable

\({\mathbb {F}}_\rho , {\mathbb {F}}_\mu , {\mathbb {F}}_\kappa \) :

Density and viscosity functional dependences

g :

Closure variable, m

G :

Closure variable, \(\hbox {m}^{2}\)

H\(_2\) :

Hydrogen species

\(\mathrm{H}_{(\mathrm{d})}^+\) :

Dissolved ion species

\(i\) :

Fuel cell current density, \(\hbox {A m}^{-2}\)

I :

Identity tensor

\(K\) :

Reaction rate coefficient, \(\hbox {m s}^{-1}\)

K :

Permeability tensor, \(\hbox {m}^{2}\)

\({{\varvec{K}}}_\mathrm{p}\) :

Integrated tensor for pressure

\({{\varvec{K}}}_\mathrm{u}\) :

Integrated tensor for velocity, \(\hbox {m}^{2}\)

\(l_\alpha \) :

Typical size of pores, m

\(L\) :

Typical size of macroscopic domain, m

\(L_\mathrm{III}\) :

Anode catalyst layer thickness, m

\(L_{\eta _i}\) :

Typical length of variation for \(\eta _i\), m

\(M_\mathrm{{H}_2}\) :

Hydrogen molar mass, \(\hbox {kg mol}^{-1}\)

\({{\varvec{n}}}_{\alpha \beta }\) :

Normal unit vector at \(\Gamma _{\alpha \beta }\)

N :

Integrated dyadic product (defined in Eq. (41)), \(\hbox {m}^{-1}\)

\(N\) :

Number of species in fluid

\(p\) :

Fluid static pressure, \(\hbox {N m}^{-2}\)

P :

Projection tensor

q :

Closure variable, \(\hbox {m s}^{-1}\)

\(r\) :

Closure variable, \(\hbox {N m}^{-2}\)

\(r_0\) :

Typical size of representative elementary volume, m

\(\mathrm{Re}_\alpha \) :

Reynolds number at the pore level

\(S\) :

Number of spatial fields

\(T\) :

Temperature, K

T :

Arbitrary tensor field

u :

Fluid velocity, \(\hbox {m s}^{-1}\)

\({{\varvec{u}}}_A ,\;{{\varvec{u}}'}_A\) :

Velocity and diffusion velocity of Species \(A,\, \hbox {m s}^{-1}\)

\({{\varvec{u}}}_B ,\;{{\varvec{u}}'}_B\) :

Velocity and diffusion velocity of Species \(B,\, \hbox {m s}^{-1}\)

\(V\) :

Volume of representative elementary volume, \(\hbox {m}^{3}\)

\(V_\alpha \) :

Volume of \(\alpha \) phase in representative elementary volume, \(\hbox {m}^{3}\)

w :

Arbitrary uniform vector field

x :

Position vector (3 components) in porous medium, m

\(\Delta {{\varvec{x}}}_i\) :

Lattice vector \(i\) of periodic porous media, m

\(y_B\) :

Species \(B\) mass fraction

\(\alpha \) :

Fluid phase

\(\alpha _\mathrm{Ox,a}\) :

Transfer coefficient

\(\beta \) :

Solid phase

\(\varepsilon \) :

Porosity

\(\Gamma _\alpha \) :

Entry and exit boundaries for representative elementary volume

\(\Gamma _{\alpha \beta }\) :

Interface between \(\alpha \) and \(\beta \) phases

\(\Gamma _\mathrm{e}\) :

Porous media entry and exit boundaries in \(\alpha \) phase

\(\kappa \) :

Fluid bulk viscosity, \(\hbox {N s m}^{-2}\)

\(\rho \) :

Density, \(\hbox {kg m}^{-3}\)

\(\Lambda \) :

Porous medium macroscopic domain

\(\Omega \) :

Representative elementary volume domain

\(\Omega _\alpha \) :

Representative elementary volume domain (\(\alpha \) phase only)

\(\eta _i\) :

Generic field

\(\theta \) :

Generic field

\(\varvec{\tau }\) :

Viscous stress tensor, \(\hbox {N m}^{-2}\)

\(\mu \) :

Fluid dynamic viscosity, \(\hbox {N s m}^{-2}\)

\(\psi \) :

Scalar, vector or tensor field in \(\alpha \) phase

\(\tilde{\psi }\) :

Spatial deviation of a field \(\psi \)

\(\left\langle \psi \right\rangle \) :

Superficial average of a field \(\psi \)

\(\left\langle \psi \right\rangle _\alpha \) :

Intrinsic average of a field \(\psi \)

References

  • Ahmadi, A., Quintard, M., Whitaker, S.: Transport in chemically and mechanically heterogeneous porous media: V. Two-equation model for solute transport with adsorption. Adv. Water Resour. 22(1), 59–86 (1998)

    Article  Google Scholar 

  • Barbir, F.: PEM Fuel Cells: Theory and Practice. Elsevier Academic Press, Amsterdam (2013)

    Google Scholar 

  • Barrere, J., Gipouloux, O., Whitaker, S.: On the closure problem for Darcy’s law. Transp. Porous Media 7(3), 209–222 (1992)

    Article  Google Scholar 

  • Bear, J.: Dynamics of Fluids in Porous Media. Dover Publications, New York (1988)

    Google Scholar 

  • Bear, J., Bachmat, Y.: Introduction to Modeling of Transport Phenomena in Porous Media. Springer, Dordrecht (1990)

    Book  Google Scholar 

  • Bird, R.B., Lightfoot, E.N., Steward, W.E.: Transport Phenomena. Wiley, New York (2007)

    Google Scholar 

  • Capozza, A., De Carli, M., Zarrella, A.: Investigations on the influence of aquifers on the ground temperature in ground-source heat pump operation. Appl. Energy 107, 350–363 (2013)

    Article  Google Scholar 

  • Carbonell, R.G., Whitaker, S.: Dispersion in pulsed systems II: Theoretical developments for passive dispersion in porous media. Chem. Eng. Sci. 38(11), 1795–1802 (1983)

    Article  Google Scholar 

  • Carbonell, R.G., Whitaker, S.: Heat and mass transfer in porous media. In: Bear, J., Corapcioglu, M.Y. (eds.) Fundamentals of Transport Phenomena in Porous Media, 121–198 (1984)

  • Çengel, Y.A., Boles, M.: Thermodynamics: An Engineering Approach, 4th edn. McGraw-Hill, New York (2001)

    Google Scholar 

  • Chen, X., Dai, J., Luo, Z.: CFD modeling using heterogeneous reaction kinetics for catalytic dehydrogenation syngas reactions in a fixed bed reactor. Particuology 11, 703–714 (2013)

    Article  Google Scholar 

  • Del Rio, J.A., Whitaker, S.: Maxwell’s equations in two-phase systems II: Two-equation model. Transp. Porous Media 39(3), 259–287 (2000)

    Article  Google Scholar 

  • Ferng, Y.M., Su, A.: A three-dimensional full-cell CFD model used to investigate the effects of different flow channel designs on PEMFC performance. Int. J. Hydrogen Energy 32(17), 4466–4476 (2007)

    Article  Google Scholar 

  • Gray, W.G.: A derivation of the equations for multi-phase transport. Chem. Eng. Sci. 30(2), 229–233 (1975)

    Article  Google Scholar 

  • Gray, W.G., O’Neill, K.: On the general equations for flow in porous media and their reduction to Darcy’s Law. Water Resour. Res. 12(2), 148–154 (1976)

    Article  Google Scholar 

  • Greenkorn, R.A.: Flow Phenomena in Porous Media. Marcel Dekker Inc., New York (1983)

    Google Scholar 

  • Hager, J., Whitaker, S.: The thermodynamic significance of the local volume averaged temperature. Transp. Porous Media 46(1), 19–35 (2002)

    Article  Google Scholar 

  • Harvey, D., Pharoah, J.G., Karan, K.: A comparison of different approaches to modelling the PEMFC catalyst layer. J. Power Sources 179(1), 209–219 (2008)

    Article  Google Scholar 

  • Howes, F.A., Whitaker, S.: The spatial averaging theorem revisited. Chem. Eng. Sci. 40(8), 1387–1392 (1985)

    Article  Google Scholar 

  • Hwang, J.J.: Thermal-electrochemical modeling of a proton exchange membrane fuel cell. J. Electrochem. Soc. 153(2), A216–A224 (2006)

    Article  Google Scholar 

  • Ju, H., Wang, C.-Y.: Experimental validation of a PEM fuel cell model by current distribution data. J. Electrochem. Soc. 151(11), A1954–A1960 (2004)

    Article  Google Scholar 

  • Kee, R.J., Coltrin, M.E., Glarborg, P.: Chemically Reacting Flow: Theory and Practice. Wiley-Interscience, Hoboken (2003)

    Book  Google Scholar 

  • Maheshwari, P.H., Mathur, R.B., Dhami, T.L.: The influence of the pore size and its distribution in a carbon paper electrode on the performance of a PEM fuel cell. Electrochim. Acta 54(2), 655–659 (2008)

    Article  Google Scholar 

  • Masliyah, J.H., Bhattacharjee, S.: Electrokinetic and Colloid Transport Phenomena. Wiley, Hoboken (2006)

    Book  Google Scholar 

  • Nikolaevskii, V.: Mechanics of porous and fractured media. In: Hsieh, R.K.T. (ed.) Series in Theoretical and Applied Mechanics 8. World Scientific Publishing, Singapore (1990)

    Google Scholar 

  • Nield, D.A., Bejan, A.: Convection in Porous Media, 4th edn. Springer, New York (2013)

    Book  Google Scholar 

  • Nozad, I., Carbonell, R.G., Whitaker, S.: Heat conduction in multiphase systems I: Theory and experiment for two-phase systems. Chem. Eng. Sci. 40(5), 843–855 (1985)

    Article  Google Scholar 

  • Ochoa-Tapia, J., Delrio, J., Whitaker, S.: Bulk and surface-diffusion in porous media—an application of the surface-averaging theorems. Chem. Eng. Sci. 48(11), 2061–2082 (1993)

    Article  Google Scholar 

  • Ochoa-Tapia, J.A., Valdes-Parada, F.J., Alvarez-Ramirez, J.: A fractional-order Darcy’s law. Physica A: Stat. Mech. Appl. 374(1), 1–14 (2007)

    Article  Google Scholar 

  • Quintard, M., Bletzacker, L., Chenu, D., Whitaker, S.: Nonlinear, multicomponent, mass transport in porous media. Chem. Eng. Sci. 61(8), 2643–2669 (2006)

    Article  Google Scholar 

  • Quintard, M., Whitaker, S.: One- and two-equation models for transient diffusion processes in two-phase systems. Adv. Heat Transfer 23, 369–464 (1993)

    Article  Google Scholar 

  • Ryan, D., Carbonell, R.G., Whitaker, S.: Effective diffusivities for catalyst pellets under reactive conditions. Chem. Eng. Sci. 35(1–2), 10–16 (1980)

    Article  Google Scholar 

  • Shimpalee, S., Beuscher, U., Van Zee, J.W.: Investigation of gas diffusion media inside PEMFC using CFD modeling. J. Power Sources 163(1), 480–489 (2006)

    Article  Google Scholar 

  • Sorek, S., Levi-Hevroni, D., Levy, A., Ben-Dor, G.: Extensions to the macroscopic Navier–Stokes Equation. Transp. Porous Media 61(2), 215–233 (2005)

    Article  Google Scholar 

  • Sorek, S., Ronen, D., Gitis, V.: Scale-dependant macroscopic balance equations governing transport through porous media: theory and observations. Transp. Porous Media 81(1), 61–72 (2010)

    Article  Google Scholar 

  • Soulaine, C., Davit, Y., Quintard, M.: A two-pressure model for slightly compressible single phase flow in bi-structured porous media. Chem. Eng. Sci. 96, 55–70 (2013)

    Article  Google Scholar 

  • Su, A., Ferng, Y.M., Shih, J.C.: CFD investigating the effects of different operating conditions on the performance and the characteristics of a high-temperature PEMFC. Energy 35(1), 16–27 (2010)

    Article  Google Scholar 

  • Valdes-Parada, F.J., Ochoa-Tapia, J.A., Alvarez-Ramirez, J.: Validity of the permeability Carman–Kozeny equation: a volume averaging approach. Physica A: Stat. Mech. Appl. 388(6), 789–798 (2009)

    Article  Google Scholar 

  • Verda, V., Baccino, G., Sciacovelli, A., Russo, S.L.: Impact of district heating and groundwater heat pump systems on the primary energy needs in urban areas. Appl. Therm. Eng. 40, 18–26 (2012)

    Article  Google Scholar 

  • Whitaker, S.: Diffusion and dispersion in porous media. AIChE J. 13(3), 420–427 (1967)

    Article  Google Scholar 

  • Whitaker, S.: Advances in theory of fluid motion in porous media. Ind. Eng. Chem. 61(12), 14–28 (1969)

    Article  Google Scholar 

  • Whitaker, S.: Simultaneous heat, mass, and momentum transfer in porous media: a theory of drying. Adv. Heat Transfer 13, 119–203 (1977)

    Article  Google Scholar 

  • Whitaker, S.: Flow in porous media I: A theoretical derivation of Darcy’s law. Transp. Porous Media 1(1), 3–25 (1986)

    Article  Google Scholar 

  • Whitaker, S.: The Forchheimer equation: a theoretical development. Transp. Porous Media 25(1), 27–61 (1996)

    Article  Google Scholar 

  • Whitaker, S.: Volume averaging of transport equations. In: du Plessis, P. (eds.) Fluid Transport in Porous Media. Advances in Fluid Mechanics, vol. 13, pp. 1–60 (1997).

  • Whitaker, S.: The Method of Volume Averaging. Springer, London (1999)

    Book  Google Scholar 

  • Whitaker, S.: Mechanics of composite solids. J. Eng. Mech. 128(8), 823–828 (2002)

    Article  Google Scholar 

  • Whitaker, S.: Derivation and application of the Stefan–Maxwell equations. Revista Mexicana de Ingenieria Quimica 8(3), 213–243 (2009)

    Google Scholar 

  • Wood, B.D., Valdés-Parada, F.J.: Volume averaging: local and nonlocal closures using a Green’s function approach. Adv. Water Resour. 51, 139–167 (2013)

    Article  Google Scholar 

  • Yu, L., Ren, G., Qin, M., Jiang, X.: Transport mechanisms and performance simulations of a PEM fuel cell with interdigitated flow field. Renew. Energy 34(3), 530–543 (2009)

    Article  Google Scholar 

Download references

Acknowledgments

The authors’ work was supported by the Natural Sciences and Engineering Research Council of Canada (NSERC).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to François Mathieu-Potvin.

Appendices

Appendix 1: Boundary Condition at \(\Gamma _{\alpha \beta }\)

In this Appendix, the boundary conditions associated to the fluid velocity \({\varvec{u}}\) is developed, the fluid being made of species \(B=1,2,\ldots ,N\). In this paper, one of these species (the species \(A)\) reacts at the surface \(\Gamma _{\alpha \beta }\), which may be written as

$$\begin{aligned} {\varvec{n}}_{\alpha \beta } \cdot \rho _A {\varvec{u}}_A =\rho \textit{Ky}_A\;\;\;\; \hbox { at }\Gamma _{\alpha \beta }, \end{aligned}$$
(99)

where \(\rho _A\) and \({\varvec{u}}_A\) are the density and velocity of species \(A\), and other variables have been defined in the text. The other species do not react at the surface \(\Gamma _{\alpha \beta }\), which yields

$$\begin{aligned} {\varvec{n}}_{\alpha \beta } \cdot \rho _B {\varvec{u}}_B =0\;\;\;\; \hbox { at } \Gamma _{\alpha \beta },\, \forall B\ne A. \end{aligned}$$
(100)

Notice that the species velocities \({\varvec{u}}_A\) or \({\varvec{u}}_B\) are defined as being the sum of the fluid velocity \({\varvec{u}}\) and of the diffusion velocity \({\varvec{{u}}}{^{\prime }}_A\) or \({\varvec{{u}}}{^{\prime }}_B\) (e.g., Eq. (12.154) in Kee et al. 2003; Eq. (12) in Whitaker 1999). In other words, they are expressed as \({\varvec{u}}_A ={\varvec{u}}+{\varvec{{u}}}{^{\prime }}_A\) and \({\varvec{u}}_B ={\varvec{u}}+{\varvec{{u}}}{^{\prime }}_B \). Hence, diffusional mass transfer at surface \(\Gamma _{\alpha \beta }\) is implicitly taken into account by \({\varvec{u}}_A\) and \({\varvec{u}}_B\), in Eqs. (99) and (100). Summation of Eq. (99) with the \(N-1\) boundary conditions given by Eq. (100) yields

$$\begin{aligned} \sum _{B=1}^N {{\varvec{n}}_{\alpha \beta } \cdot \rho _B {\varvec{u}}_B} =\rho \textit{Ky}_A \;\;\;\;\hbox { at } \Gamma _{\alpha \beta } \end{aligned}$$
(101)

or more concisely,

$$\begin{aligned} {\varvec{n}}_{\alpha \beta } \cdot {\varvec{u}}=\textit{Ky}_A\;\;\;\; \hbox { at }\Gamma _{\alpha \beta }, \end{aligned}$$
(102)

where the variable \(\rho \) has been canceled out using the following definitions (Whitaker 2009):

$$\begin{aligned} \rho \equiv \sum _{B=1}^N {\rho _B} \;\;\;\;\;\; {\varvec{u}}\equiv \frac{1}{\rho }\sum _{B=1}^N \rho _B {\varvec{u}}_B. \end{aligned}$$
(103)

Hence, the normal component of the fluid velocity u at the surface \(\Gamma _{\alpha \beta }\) is determined by Eq. (102). Finally, it is assumed that the no-slip condition is applied at the boundaries \(\Gamma _{\alpha \beta } \). In other words, the tangential component of u is zero, so that the boundary condition for \({\varvec{u}}\) may finally be written as \({\varvec{u}}=\textit{Ky}_A {\varvec{n}}_{\alpha \beta }\) at \(\Gamma _{\alpha \beta }\).

Appendix 2: Spatial Averaging Theorems

In this section, six different versions of the spatial averaging theorem are presented. To do so, we start with the fundamental spatial averaging theorem (Whitaker 1967) for a scalar field \(a\), i.e.,

$$\begin{aligned} \frac{1}{V}\int \limits _{\Omega _\alpha } \nabla a\; {\text{ d }}V =\nabla \left( {\frac{1}{V}\int \limits _{\Omega _\alpha } a\; {\text{ d }}V} \right) +\frac{1}{V}\int \limits _{\Gamma _{\alpha \beta }} {{\varvec{n}}_{\alpha \beta } a\; {\text{ d }}A}, \end{aligned}$$
(104)

which may be written in a more compact form as

1.1 Theorem 1

$$\begin{aligned} \left\langle {\nabla a} \right\rangle =\nabla \left\langle a \right\rangle +\frac{1}{V}\int \limits _{\Gamma _{\alpha \beta }} {{\varvec{n}}_{\alpha \beta } a\; {\text{ d }}A}. \end{aligned}$$
(105)

This theorem is valid for representative elementary volumes (REV) of constant size and orientation. As one can see in Eq. (105), this theorem allows an interchange between the averaging operator and the gradient operator \(\nabla \) applied to a scalar field \(a\) in a porous medium.

To obtain an equivalent theorem for a vector field \({\varvec{a}}\), the expression \(\left\langle {\nabla {\varvec{a}}} \right\rangle \) is multiplied by an arbitrary uniform vector \({\varvec{w}}\), which yields

$$\begin{aligned} \left\langle {\nabla {\varvec{a}}} \right\rangle \cdot {\varvec{w}}=\left\langle {\nabla \left( {{\varvec{a}}\cdot {\varvec{w}}} \right) }\right\rangle . \end{aligned}$$
(106)

The equality given in Eq. (106) may be demonstrated with tensor analysis. The expression \(\left( {{\varvec{a}}\cdot {\varvec{w}}} \right) \) in the r.h.s. of Eq. (106) is a scalar field; hence, it may be inserted in the fundamental spatial averaging theorem Eq. (105), which yields

$$\begin{aligned} \left\langle {\nabla {\varvec{a}}} \right\rangle \cdot {\varvec{w}}=\nabla \left\langle {\left( {{\varvec{a}}\cdot {\varvec{w}}} \right) } \right\rangle +\frac{1}{V}\int \limits _{\Gamma _{\alpha \beta }} {{\varvec{n}}_{\alpha \beta } \left( {{\varvec{a}}\cdot {\varvec{w}}} \right) {\text{ d }}A}. \end{aligned}$$
(107)

Using the fact that w is uniform, Eq. (107) may be written as

$$\begin{aligned} \left\langle {\nabla {\varvec{a}}} \right\rangle \cdot {\varvec{w}}=\nabla \left\langle {\varvec{a}} \right\rangle \cdot {\varvec{w}}+\left( \frac{1}{V}\int \limits _{\Gamma _{\alpha \beta }} {\varvec{n}}_{\alpha \beta } {\varvec{a}}\, {\text{ d }}A \right) \cdot {\varvec{w}}. \end{aligned}$$
(108)

Since Eq. (108) is valid for any uniform vector \({\varvec{w}}\), it may be reduced to

1.2 Theorem 2

$$\begin{aligned} \left\langle {\nabla {\varvec{a}}} \right\rangle =\nabla \left\langle {\varvec{a}} \right\rangle +\frac{1}{V}\int \limits _{\Gamma _{\alpha \beta }} {\varvec{n}}_{\alpha \beta } {\varvec{a}}\; {\text{ d }}A. \end{aligned}$$
(109)

To obtain an equivalent theorem for the divergence of a vector field \({\varvec{a}}\), the procedure proposed at Problem 1.7 in Whitaker (1999) is used, i.e., the double contraction of Eq. (109) is made with the identity tensor \({\varvec{I}}\), which yields

$$\begin{aligned} \left\langle {\nabla {\varvec{a}}} \right\rangle :{\varvec{I}}=\nabla \left\langle {\varvec{a}} \right\rangle :{\varvec{I}}+\left( {\frac{1}{V}\int \limits _{\Gamma _{\alpha \beta }} {\varvec{n}}_{\alpha \beta } {\varvec{a}}\; {\text{ d }}A} \right) :{\varvec{I}}. \end{aligned}$$
(110)

It can be shown with elementary tensor algebra that Eq. (110) may be written as

1.3 Theorem 3

$$\begin{aligned} \left\langle {\nabla \cdot {\varvec{a}}} \right\rangle =\nabla \cdot \left\langle {\varvec{a}} \right\rangle +\frac{1}{V}\int \limits _{\Gamma _{\alpha \beta }} {\varvec{n}}_{\alpha \beta } \cdot {\varvec{a}}\; {{\text{ d }}A}. \end{aligned}$$
(111)

To obtain an equivalent theorem for the divergence of a tensor field \({\varvec{C}}\), the expression \(\left\langle {\nabla \cdot {\varvec{C}}} \right\rangle \) is multiplied by an arbitrary uniform vector w, which yields

$$\begin{aligned} \left\langle {\nabla \cdot {\varvec{C}}} \right\rangle \cdot {\varvec{w}}=\left\langle {\nabla \cdot \left( {\varvec{C}}\cdot {\varvec{w}} \right) } \right\rangle . \end{aligned}$$
(112)

The equality given in Eq. (112) may be demonstrated with tensor analysis. The expression \(\left( {{\varvec{C}}\cdot {\varvec{w}}} \right) \) in the r.h.s. of Eq. (112) is a vector field; hence, it may be inserted in the spatial averaging theorem Eq. (111), which yields

$$\begin{aligned} \left\langle {\nabla \cdot {\varvec{C}}} \right\rangle \cdot {\varvec{w}}=\nabla \cdot \left\langle {{\varvec{C}}\cdot {\varvec{w}}} \right\rangle +\frac{1}{V}\int \limits _{\Gamma _{\alpha \beta }} {\varvec{n}}_{\alpha \beta } \cdot \left( {{\varvec{C}}\cdot {\varvec{w}}} \right) \;{{\text{ d }}A}. \end{aligned}$$
(113)

Using the fact that w is uniform, Eq. (113) may be written as

$$\begin{aligned} \left\langle {\nabla \cdot {\varvec{C}}} \right\rangle \cdot {\varvec{w}}=\left( {\nabla \cdot \left\langle {\varvec{C}} \right\rangle } \right) \cdot {\varvec{w}}+\left( \frac{1}{V}\int \limits _{\Gamma _{\alpha \beta }} {\varvec{n}}_{\alpha \beta } \cdot {\varvec{C}}\, {{\text{ d }}A} \right) \cdot {\varvec{w}}. \end{aligned}$$
(114)

Since Eq. (114) is valid for any uniform vector w, it may be reduced to

1.4 Theorem 4

$$\begin{aligned} \left\langle {\nabla \cdot {\varvec{C}}} \right\rangle =\nabla \cdot \left\langle {\varvec{C}} \right\rangle +\frac{1}{V}\int \limits _{\Gamma _{\alpha \beta }} {\varvec{n}}_{\alpha \beta } \cdot {\varvec{C}}\; {{\text{ d }}A}. \end{aligned}$$
(115)

This theorem has already been used by Gray and O’Neill (1976). Another expression of interest is the transpose of a vector gradient, i.e., \(\left( {\nabla {\varvec{a}}} \right) ^{\dagger }\). This term appears, for example, in the viscous stress tensor expression for Newtonian fluid. The first step is to apply the transpose operator to Eq. (109), i.e.,

$$\begin{aligned} \left\langle {\nabla {\varvec{a}}} \right\rangle ^{\dagger }=\left( {\nabla \left\langle {\varvec{a}} \right\rangle } \right) ^{\dagger }+\left( {\frac{1}{V}\int \limits _{\Gamma _{\alpha \beta }} {{\varvec{n}}_{\alpha \beta } {\varvec{a}}\, {{\text{ d }}A}}} \right) ^{\dagger }. \end{aligned}$$
(116)

The transpose and integral operator being commutable, Eq. (116) may be written as

$$\begin{aligned} \left\langle {\left( {\nabla {\varvec{a}}} \right) ^{\dagger }} \right\rangle =\left( {\nabla \left\langle {\varvec{a}} \right\rangle } \right) ^{\dagger }+\frac{1}{V}\int \limits _{\Gamma _{\alpha \beta }} {\left( {{\varvec{n}}_{\alpha \beta } {\varvec{a}}} \right) ^{\dagger } {{\text{ d }}A}}. \end{aligned}$$
(117)

and with tensor algebra one may prove that the transpose of a dyadic product \(\left( {{\varvec{n}}_{\alpha \beta } {\varvec{a}}} \right) ^{\dagger }\) becomes \({\varvec{an}}_{\alpha \beta }\), which leads to

1.5 Theorem 5

$$\begin{aligned} \left\langle {\left( {\nabla {\varvec{a}}} \right) ^{\dagger }} \right\rangle =\left( {\nabla \left\langle {\varvec{a}} \right\rangle } \right) ^{\dagger }+\frac{1}{V}\int \limits _{\Gamma _{\alpha \beta }} {\varvec{an}}_{\alpha \beta } \; {{\text{ d }}A}. \end{aligned}$$
(118)

Another expression of interest has the form \(\left( {\nabla \cdot {\varvec{a}}} \right) {\varvec{I}}\). This term appears, for example, in the viscous stress tensor expression for Newtonian fluid. To obtain a spatial averaging theorem for this expression, the identity tensor \({\varvec{I}}\) is extracted from the averaging operator , i.e.,

$$\begin{aligned} \left\langle {\left( {\nabla \cdot {\varvec{a}}} \right) {\varvec{I}}} \right\rangle =\left\langle {\nabla \cdot {\varvec{a}}} \right\rangle {\varvec{I}}. \end{aligned}$$
(119)

The expression \(\left\langle {\nabla \cdot {\varvec{a}}} \right\rangle \) in Eq. (119) may be developed by Eq. (111). Hence, Eq. (119) may be written as

1.6 Theorem 6

$$\begin{aligned} \left\langle {\left( {\nabla \cdot {\varvec{a}}} \right) {\varvec{I}}} \right\rangle =\left( {\nabla \cdot \left\langle {\varvec{a}} \right\rangle } \right) {\varvec{I}}+\left( \frac{1}{V}\int \limits _{\Gamma _{\alpha \beta }} {\varvec{n}}_{\alpha \beta } \cdot {\varvec{a}}\; {\text{ d }}A \right) {\varvec{I}}. \end{aligned}$$
(120)

Theorems 5 and 6 have already been used by Gray and O’Neill (1976), and also by Whitaker (2002). Other forms of the spatial averaging theorems have also been developed by Del Rio and Whitaker (2000) for electrodynamics, and by Ochoa-Tapia et al. (1993) for surface diffusion, but they are not used in this text.

Appendix 3: Estimation of \(K\left\langle {y_A} \right\rangle _\alpha \)

The value of the expression of \(K\left\langle {y_A} \right\rangle _\alpha \) in the anode catalyst layer of a polymer electrolyte fuel cell (where species \(A\) is hydrogen) is to be estimated in this section.

First, the current density \(i\) of a fuel cell is converted into a volumetric mass source in the anode catalyst layer by the relation

$$\begin{aligned} \hbox {volumetric mass source}={iM_{\mathrm{H}_2}}/{2L_{\text {III}} F}, \end{aligned}$$
(121)

where the current density \(i\) is of the order of \(1\times 10^{4}\hbox { A/m}^{2}\) (Barbir 2013), the thickness \(L_\mathrm{III}\) of catalyst layers (see Layer III in Fig. 3) is of the order of \(3\times \hbox {10}^{-4}\hbox {m}\) (Ju and Wang 2004), the molar mass of hydrogen \(\mathrm{H}_2\) is \(2\times 10^{-3}\hbox { kg/mol}\) (Çengel and Boles 2001), and the Faraday constant \(F\) is equal to \(96487\hbox { C/mol}\) (Ju and Wang 2004).

Moreover, one has to recall that the volumetric mass source expression given in Eq. (21) is \(\left\langle \rho \right\rangle _\alpha a_\mathrm{V} K\left\langle {y_A} \right\rangle _\alpha \). The surface density \(a_V\) is of the order of \(1/{l_\alpha }\), where \(l_\alpha \approx 3\times 10^{-5}\hbox { m}\) (Maheshwari et al. 2008), and the density of a mixture of hydrogen and steam at standard pressure and temperature (for a relative humidity of 1) is on the order of \(0.35\hbox { kg/m}^{3}\). Equating the volumetric mass source expression for a PEMFC anode catalyst given in Eq. (121) to the volumetric mass source expression given in Eq. (21) leads to

$$\begin{aligned} \left\langle \rho \right\rangle _\alpha a_\mathrm{V} K\left\langle {y_A} \right\rangle _\alpha ={iM_{\mathrm{H}_2}}/{2L_{\text {III}} F}, \end{aligned}$$
(122)

and isolating the expression \(K\left\langle {y_A} \right\rangle _\alpha \) yields

$$\begin{aligned} K\left\langle {y_A} \right\rangle _\alpha ={iM_{\mathrm{H}_2}}/2L_\mathrm{{III}} F\left\langle \rho \right\rangle _\alpha a_\mathrm{V}. \end{aligned}$$
(123)

Using the values given above, the order of magnitude of \(K\left\langle {y_A} \right\rangle _\alpha \) is \(3\times 10^{-5}\hbox {m}/\hbox {s}\).

Appendix 4: Constraints for \(\tilde{{\varvec{u}}}\) at \(\Gamma _{\alpha \beta }\)

In the course of the macroscopic momentum equation development performed in Sect. 4.2, it was asserted (without demonstration) that the expression \({\varvec{n}}_{\alpha \beta } \cdot \nabla _\dagger \tilde{{\varvec{u}}}\) present in Eq. (47) was equal to zero at the surface \(\Gamma _{\alpha \beta }\) (the reader should recall that in this paper, the notation \(\nabla _\dagger \tilde{{\varvec{u}}}\) is used for \(\left( {\nabla \tilde{{\varvec{u}}}} \right) ^{\dagger })\). This assertion is demonstrated in the following paragraphs.

The first step is to obtain a practical expression equivalent to the constraint on \(\tilde{{\varvec{u}}}\) given in Eq. (77), i.e.,

$$\begin{aligned} \tilde{{\varvec{u}}}=-\left\langle {\varvec{u}} \right\rangle _\alpha \;\;\; \hbox { at } \Gamma _{\alpha \beta }. \end{aligned}$$
(124)

Using the approximation that the value of \(\left\langle {\varvec{u}} \right\rangle _\alpha \) undergoes negligible change with position within a REV, the constraint Eq. (124) means that the value of \(\tilde{{\varvec{u}}}\) does not change when the position considered is on the surface \(\Gamma _{\alpha \beta }\). This can be mathematically expressed as

$$\begin{aligned} \nabla _S \tilde{{\varvec{u}}}=0\;\;\;\; \hbox { at } \Gamma _{\alpha \beta }. \end{aligned}$$
(125)

The surface gradient operator \(\nabla _S\) present in Eq. (125) is defined as

$$\begin{aligned} \nabla _S =\left( {{\varvec{P}}\cdot \nabla } \right) \;\;\;\hbox { at } \Gamma _{\alpha \beta }, \end{aligned}$$
(126)

where the projection tensor \({\varvec{P}}\) is defined as \({\varvec{P}}\equiv {\varvec{I}}-{\varvec{n}}_{\alpha \beta } {\varvec{n}}_{\alpha \beta }\) (Ochoa-Tapia et al. 1993). Using these definitions of \({\varvec{P}}\) and \(\nabla _S\) into Eq. (125) leads to

$$\begin{aligned} \nabla \tilde{{\varvec{u}}}-{\varvec{n}}_{\alpha \beta } {\varvec{n}}_{\alpha \beta } \cdot \nabla \tilde{{\varvec{u}}}=0 \;\;\;\hbox { at }\Gamma _{\alpha \beta }. \end{aligned}$$
(127)

The general relations \(\left( {{\varvec{T}}\cdot \nabla } \right) {\varvec{a}}\equiv {\varvec{T}}\cdot \nabla {\varvec{a}}\) and \({\varvec{I}}\cdot \nabla {\varvec{a}}\equiv \nabla {\varvec{a}}\), which hold for any vector \({\varvec{a}}\) and tensor \({\varvec{T}}\), were used to obtain Eq. (127) from Eq. (125). Finally, Eq. (127) is rearranged to obtain the expression

$$\begin{aligned} \text {Constraint }\, \textit{1}\;\;\;\;\;\;\; \nabla \tilde{{\varvec{u}}}={\varvec{n}}_{\alpha \beta } \left( {{\varvec{n}}_{\alpha \beta } \cdot \nabla \tilde{{\varvec{u}}}} \right) \;\;\;\;\hbox { at }\Gamma _{\alpha \beta }, \end{aligned}$$
(128)

where the general relation \({\varvec{aa}}\cdot \nabla {\varvec{b}}\equiv {\varvec{a}}\left( {{\varvec{a}}\cdot \nabla {\varvec{b}}} \right) \), which holds for any vector fields \({\varvec{a}}\) and \({\varvec{b}}\), was used. Notice that Eq. (128) is just another way to write the constraint given by Eq. (124).

The second step is to write the other constraint on \(\tilde{{\varvec{u}}}\) that was given in Sect. 5.1 (see Eq. (61)), i.e.,

$$\begin{aligned} \nabla \cdot \tilde{{\varvec{u}}}=0. \end{aligned}$$
(129)

Since this expression is equal to the scalar zero, one may multiply Eq. (129) by the vector \({\varvec{n}}_{\alpha \beta }\), so that the r.h.s. of this equation becomes a vector equal to zero, i.e., \({\varvec{n}}_{\alpha \beta } \left( {\nabla \cdot \tilde{{\varvec{u}}}} \right) =0\). Moreover, it is easy to show with tensor algebra that, for any vector field \({\varvec{a}}\), the relation \(\nabla \cdot {\varvec{a}}\equiv \nabla {\varvec{a}}:{\varvec{I}}\) is true; hence, Eq. (129) becomes

$$\begin{aligned} \text {Constraint }\, \textit{2}\;\;\;\; \underbrace{{\varvec{n}}_{\alpha \beta }}_{\hbox {vector}}\underbrace{\left( {\nabla \tilde{{\varvec{u}}}:{\varvec{I}}} \right) }_{\hbox {scalar}}=\underbrace{0}_{\begin{array}{c} \hbox {nul} \\ \hbox {vector} \\ \end{array}} \end{aligned}$$
(130)

Notice that the constraint given by Eq. (130) is valid everywhere in the \(\alpha \) phase, including the surface \(\Gamma _{\alpha \beta }\) where the boundary conditions for \(\tilde{{\varvec{u}}}\) applies.

The objective is to prove, with the help of the two constraints given by Eq. (128) and Eq. (130), that \({\varvec{n}}_{\alpha \beta } \cdot \nabla _\dagger \tilde{{\varvec{u}}}=0\) at \(\Gamma _{\alpha \beta }\). Hence, the next step consists in writing the expression

$$\begin{aligned} {\varvec{n}}_{\alpha \beta } \cdot \nabla _\dagger \tilde{{\varvec{u}}}={\varvec{d}}\;\;\;\hbox { at }\Gamma _{\alpha \beta }, \end{aligned}$$
(131)

where \({\varvec{d}}\) is the unknown vector to find. Using the fact that \({\varvec{a}}\cdot \nabla _\dagger {\varvec{b}}\equiv \nabla {\varvec{b}}\cdot {\varvec{a}}\) for any vector fields \({\varvec{a}}\) and \({\varvec{b}}\), Eq. (131) is rewritten as

$$\begin{aligned} \nabla \tilde{{\varvec{u}}}\cdot {\varvec{n}}_{\alpha \beta } ={\varvec{d}} \end{aligned}$$
(132)

Since the expression given by Eq. (130) is equal to zero, one can simply subtract Eq. (130) from Eq. (132) so as to obtain

$$\begin{aligned} \nabla \tilde{{\varvec{u}}}\cdot {\varvec{n}}_{\alpha \beta } -\underbrace{{\varvec{n}}_{\alpha \beta } \left( {\nabla \tilde{{\varvec{u}}}:{\varvec{I}}} \right) }_{\begin{array}{c} \hbox {zero} \\ \hbox {by Eq. (130)} \\ \end{array}}={\varvec{d}} \end{aligned}$$
(133)

Then, the constraint given in Eq. (128), (i.e., \(\nabla \tilde{{\varvec{u}}}={\varvec{n}}_{\alpha \beta } \left( {{\varvec{n}}_{\alpha \beta } \cdot \nabla \tilde{{\varvec{u}}}} \right) )\) is used by inserting directly that expression where \(\nabla \tilde{{\varvec{u}}}\) is present in Eq. (133). Hence, Eq. (133) becomes

$$\begin{aligned} \left( {{\varvec{n}}_{\alpha \beta } \left( {{\varvec{n}}_{\alpha \beta } \cdot \nabla \tilde{{\varvec{u}}}} \right) } \right) \cdot {\varvec{n}}_{\alpha \beta } -{\varvec{n}}_{\alpha \beta } \left( {\left[ {{\varvec{n}}_{\alpha \beta } \left( {{\varvec{n}}_{\alpha \beta } \cdot \nabla \tilde{{\varvec{u}}}} \right) } \right] :{\varvec{I}}} \right) ={\varvec{d}}. \end{aligned}$$
(134)

The strategy is to prove that the second term in the l.h.s. of Eq. (134) is equivalent to the first, which would prove that \({\varvec{d}}\) is equal to zero. To do so, the term \(\left[ {{\varvec{n}}_{\alpha \beta } \left( {{\varvec{n}}_{\alpha \beta } \cdot \nabla \tilde{{\varvec{u}}}} \right) } \right] :{\varvec{I}}\) in Eq. (134) is transformed into the expression \({\varvec{n}}_{\alpha \beta } \cdot \left( {{\varvec{n}}_{\alpha \beta } \cdot \nabla \tilde{{\varvec{u}}}} \right) \), using a relation that hold for any vector field \({\varvec{a}}\) and \({\varvec{b}}\), i.e., \({\varvec{ab}}:{\varvec{I}}\equiv {\varvec{a}}\cdot {\varvec{b}}\). In this process, the expression \({\varvec{n}}_{\alpha \beta }\) was associated to \({\varvec{a}}\) and \(\left( {{\varvec{n}}_{\alpha \beta } \cdot \nabla \tilde{{\varvec{u}}}} \right) \) was associated to \({\varvec{b}}\). Therefore, Eq. (134) becomes

$$\begin{aligned} \left( {{\varvec{n}}_{\alpha \beta } \left( {{\varvec{n}}_{\alpha \beta } \cdot \nabla \tilde{{\varvec{u}}}} \right) } \right) \cdot {\varvec{n}}_{\alpha \beta } -{\varvec{n}}_{\alpha \beta } \left( {{\varvec{n}}_{\alpha \beta } \cdot \left( {{\varvec{n}}_{\alpha \beta } \cdot \nabla \tilde{{\varvec{u}}}} \right) } \right) ={\varvec{d}}. \end{aligned}$$
(135)

The general relation \({\varvec{a}}\left( {{\varvec{b}}\cdot {\varvec{c}}} \right) \equiv {\varvec{bc}}\cdot {\varvec{a}}\), which holds for any vector fields \({\varvec{a}},\, {\varvec{b}}\), and \({\varvec{c}}\), is used to transform the expression \({\varvec{n}}_{\alpha \beta } \left( {{\varvec{n}}_{\alpha \beta } \cdot \left( {{\varvec{n}}_{\alpha \beta } \cdot \nabla \tilde{{\varvec{u}}}} \right) } \right) \) into \(\left( {{\varvec{n}}_{\alpha \beta } \left( {{\varvec{n}}_{\alpha \beta } \cdot \nabla \tilde{{\varvec{u}}}} \right) } \right) \cdot {\varvec{n}}_{\alpha \beta } \). In this process, the expression \({\varvec{n}}_{\alpha \beta }\) was associated to \({\varvec{a}}\) and \({\varvec{b}}\), and the expression \(\left( {{\varvec{n}}_{\alpha \beta } \cdot \nabla \tilde{{\varvec{u}}}} \right) \) was associated to \({\varvec{c}}\) . Therefore, Eq. (135) becomes

$$\begin{aligned} \underbrace{\left( {{\varvec{n}}_{\alpha \beta } \left( {{\varvec{n}}_{\alpha \beta } \cdot \nabla \tilde{{\varvec{u}}}} \right) } \right) \cdot {\varvec{n}}_{\alpha \beta } -\left( {{\varvec{n}}_{\alpha \beta } \left( {{\varvec{n}}_{\alpha \beta } \cdot \nabla \tilde{{\varvec{u}}}} \right) } \right) \cdot {\varvec{n}}_{\alpha \beta }}_{\hbox {zero}}={\varvec{d}}, \end{aligned}$$
(136)

where the l.h.s. is identically equal to zero.

To summarize, by transforming the two constraints given by Eqs. (124) and (129) into Eqs. (128) and (130), respectively, it was possible to prove that the value of \({\varvec{d}}\) in Eq. (131) is equal to zero. In other words, it was proved that \({\varvec{n}}_{\alpha \beta }\cdot \nabla _\dagger \tilde{{\varvec{u}}}=0\) at \(\Gamma _{\alpha \beta }\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mathieu-Potvin, F., Gosselin, L. Impact of Non-uniform Properties on Governing Equations for Fluid Flows in Porous Media. Transp Porous Med 105, 277–314 (2014). https://doi.org/10.1007/s11242-014-0370-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11242-014-0370-z

Keywords

Navigation