Log in

Effective diffusivity for a mixed-matrix membrane

  • Published:
Journal of Engineering Mathematics Aims and scope Submit manuscript

Abstract

The effective diffusivity or permeability of a mixed-matrix membrane composed of spherical fillers of one material distributed within a continuous matrix of another material is computed by a boundary-element method. The boundary value of the concentration of the diffusing species on the surface of the fillers is proportional to that of the adjoining matrix phase according to a linear sorption/desorption kinetics law that is responsible for an interfacial partition coefficient, while the diffusive flux is continuous across the interface. The solution of Laplace’s equation for the concentration field is constructed in terms of the single- and double-layer harmonic potentials involving the boundary values and boundary distribution of the normal derivative of the concentration and the free-space Green’s function of Laplace’s equation in three dimensions according to the standard boundary-integral formulation. The effective permeability of the membrane depends on the ratio of the filler-to-matrix permeabilities and not separately on the diffusivity ratio and partition coefficient. The results are in excellent agreement with the predictions of existing analytical models for random and ordered distributions of spherical particles within an infinite, continuous medium in the entire range of filler volume fractions considered. Extensions to systems governed by an interfacial contact resistance and nonlinear interfacial kinetics are discussed.

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

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price includes VAT (Canada)

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10

Similar content being viewed by others

References

  1. Maxwell JC (1873) Treatise on electricity and magnetism. Oxford University Press, London

    MATH  Google Scholar 

  2. Chung TS, Jiang LY, Li Y, Kulprathipanja S (2007) Mixed matrix membranes (MMMs) comprising organic polymers with dispersed inorganic fillers for gas separation. Prog Polym Sci 32:483–507

    Article  Google Scholar 

  3. Petropoulos JH (1985) A comparative study of approaches applied to the permeability of binary composite polymeric materials. J Polym Sci 23:1309–1324

    Google Scholar 

  4. Minelli M, Doghieri F, Papadokostaki KG, Petropoulos JH (2013) A fundamental study of the extent of meaningful application of Maxwell’s and Wiener’s equations to the permeability of binary composite materials. Part I: a numerical computation approach. Chem Eng Sci 104:630–637

    Article  Google Scholar 

  5. Marand E, Surapathi A (2012) The role of solubility partition coefficient at the mixed matrix interface in the performance of mixed-matrix membranes. J Membr Sci 415–416:871–877

    Article  Google Scholar 

  6. Singh T, Kang D-Y, Nair S (2013) Rigorous calculations of permeation in mixed-matrix membranes: evaluation of interfacial equilibrium effects and permeability-based models. J Membr Sci 448:160–169

    Article  Google Scholar 

  7. Wang T-P, Kang D-Y (2015) Predictions of effective diffusivity of mixed matrix membranes with tubular fillers. J Membr Sci 485:123–131

    Article  Google Scholar 

  8. Petropoulos JH, Papadokostaki KG, Minelli M, Doghieri F (2014) On the role of diffusivity ratio and partition coefficient in diffusional molecular transport in binary composite materials, with special reference to the Maxwell equation. J Membr Sci 456:162–166

    Article  Google Scholar 

  9. Pozrikidis C, Ford DM (2017) Conductive transport through a mixed-matrix membrane. J Eng Math 105:189–202

    Article  MathSciNet  Google Scholar 

  10. Pozrikidis C (2017) Heat conduction across an array of cylinders. Eng Anal Bound Elem 85:105–110

    Article  MathSciNet  Google Scholar 

  11. Benveniste Y (1987) Effective thermal conductivity of composites with a thermal contact resistance between the constituents: nondilue case. J Apply Phys 61:2840–2843

    Article  Google Scholar 

  12. Cheng H, Torquato S (1997) Effective conductivity of periodic arrays of spheres with interfacial resistance. Proc R Soc Lond A 453:145–161

    Article  Google Scholar 

  13. Duschlbauer H, Pettermann E, Böhm HJ (2003) Heat conduction of a spheroidal inhomogeneity with imperfectly bonded interface. J Appl Phys 94:1539–1549

    Article  Google Scholar 

  14. Cussler EL (1997) Diffusion, mass transfer in fluid systems, 2nd edn. Cambridge University Press, New York

    Google Scholar 

  15. Deen WM (1998) Analysis of transport phenomena. Oxford University Press, New York

    Google Scholar 

  16. Pozrikidis C (2002) A practical guide to boundary element methods. Chapman & Hall, Boca Raton

    MATH  Google Scholar 

  17. Dunavant DA (1985) High degree efficient symmetrical Gaussian quadrature rules for the triangle. Int J Numer Methods Eng 21:1129–1148

    Article  MathSciNet  Google Scholar 

  18. Metropolis N, Rosenbluth AW, Rosenbluth NM, Teller AH, Teller E (1953) Equation of state calculations by fast computing machines. J Chem Phys 21:1087–1092

    Article  Google Scholar 

  19. Minelli M, Baschetti MG, Doghieri F (2011) A comprehensive model for mass transport properties in nanocomposites. J Membr Sci 381:10–20

    Article  Google Scholar 

  20. Goldberg M (1971) On the densest packing of equal spheres in a cube. Math Mag 44:199–208

    Article  MathSciNet  Google Scholar 

  21. Specht E (2013) The best known packings of equal spheres in a cube. Accessed 3 Mar 2017

  22. Crank J (1975) The mathematics of diffusion, 2nd edn. Oxford University Press, New York

    MATH  Google Scholar 

  23. Rayleigh, Lord (1892) On the influence of obstacles arranged in rectangular order upon the properties of a medium. Philos Mag 34:481–502

    Article  Google Scholar 

  24. de Vries DA (1952) The thermal conductivity of granular materials. Annex 1952, 1. Bull Inst Intern du Froid: 115–131

  25. Jeffrey DJ (1973) Conduction through a random suspension of spheres. Proc R Soc Lond A 335:355–367

    Article  Google Scholar 

  26. Meredith RE, Tobias CW (1960) Resistance to potential flow through a cubical array of spheres. J Appl Phys 32:1270–1273

    Article  Google Scholar 

  27. Lewis T, Nielsen L (1970) Dynamic mechanical properties of particulate-filled composites. J Appl Poly Sci 14:1449–1471

    Article  Google Scholar 

  28. Pal R (2008) On the Lewis–Nielsen model for thermal/electrical conductivity of composites. Composites A 39:718–726

    Article  Google Scholar 

Download references

Acknowledgements

The author expresses his gratitude to C. Pozrikidis for many insightful discussions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to J. M. Davis.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Effective diffusivity at small volume fractions

For a given arrangement of filler particles, asymptotic and approximate methods provide predictions for the effective diffusivity or permeability of a mixed-matrix membrane containing spherical particles at small filler volume fractions as a function of two parameters,

$$\begin{aligned} \frac{D_{\mathrm{eff}}}{D_m} = f(k \alpha , c_v), \end{aligned}$$
(A.1)

where k is the partition coefficient at the matrix–filler interface, \(\alpha \equiv D_f/D_m\) is the ratio of the diffusivity of the spherical fillers to that of the matrix, and \(c_v\) is the volume fraction of the filler.

The simplest model is an adaptation of Maxwell’s seminal result [1] in electrostatics, yielding

$$\begin{aligned} f(k \alpha , c_v) \simeq \frac{k \alpha + 2 + 2 c_v (k \alpha -1)}{k \alpha + 2 - c_v (k \alpha -1)} + O(c_v^2) = 1 + 3 \, \frac{k \alpha - 1}{k \alpha + 2} \, c_v + O(c_v^2). \end{aligned}$$
(A.2)

In Maxwell’s asymptotic analysis, the scalar distribution in and around an isolated sphere suspended in an infinite matrix subject to a uniform gradient far from the sphere is employed to compute the effective diffusivity of a dilute dispersion, with the presence of other spheres neglected. Although the framework is limited formally to very small volume fractions, Maxwell’s predictions based on the first form of the expression in (A.2) have been shown to be reasonably accurate even for moderate volume fractions of fillers. The \(O(c_v^2)\) contribution resulting from sphere–sphere interactions in random suspensions for \(c_v \ll 1\) was given by Jeffrey [25].

Maxwell’s analysis was extended by Rayleigh [23] to nondilute systems of spherical particles arranged in a simple cubic lattice to yield an infinite series for the effective diffusivity ratio. Meredith and Tobias [26] reformulate the results as a truncated series,

$$\begin{aligned} f(k\alpha , c_v) = 1 + \frac{3}{\Lambda } \, c_v, \end{aligned}$$
(A.3)

where

$$\begin{aligned} \Lambda = \frac{k \alpha + 2}{k \alpha -1}&- c_v - 1.315 \, \frac{c_v^{10/3}}{(k \alpha + 4/3)/(k \alpha -1) - 0.409 c_v^{7/3}} - 0.016 \frac{(k \alpha - 1) c_v^{14/3}}{k \alpha + 6/5} . \end{aligned}$$
(A.4)

Corresponding result for a face-centered cubic lattice of spherical particles was derived by de Vries [24] to a lower order in \(c_v\),

$$\begin{aligned} \Lambda = \frac{k \alpha + 2}{k \alpha -1} - c_v - 0.0752 \, \frac{(k \alpha - 1) c_v^{10/3}}{k \alpha + 4/3} . \end{aligned}$$
(A.5)

For a body-centered cubic lattice, the coefficient 0.0752 is replaced by 0.129 [24].

More complicated models result in implicit expressions for the effective diffusivity and include additional parameters such as the particle shape or size distribution of polydisperse fillers. For example, the Lewis–Nielsen model [27] has been shown to describe a vast amount of available experimental data on the thermal and electrical conductivities of particulate composites [28]. It predicts an effective diffusivity

$$\begin{aligned} \frac{D_\mathrm {eff}}{D_m} = \frac{1+A B c_v }{1 - B F c_v}, \end{aligned}$$
(A.6)

where

$$\begin{aligned} A = k_E - 1, \qquad B = \frac{k \alpha -1}{k \alpha + 2}, \qquad F = 1 + \left( \frac{1 - c_{v, \mathrm{max}}}{c_{v, \mathrm{max}}^2} \right) c_v, \end{aligned}$$
(A.7)

\(k_E\) is the Einstein coefficient, equal to 5/2 for rigid spheres, and \(c_{v, \mathrm{max}}\) is the maximum packing volume fraction of particles: \(c_{v, \mathrm{max}} = 0.637\) for spheres in random close packing, \(\pi /6\) for a simple cubic lattice, and \(\pi /(3 \sqrt{2})\) for a fcc lattice or hexagonal packing. It has been suggested that \(A = 2.0\) (instead of 1.5) for spherical particles so that the Lewis–Nielsen model reduces to the Maxwell model as \(c_{v, \mathrm{max}} \rightarrow 1\) [28]. As \(k \alpha \rightarrow \infty \), the Lewis–Nielsen model predicts that \(D_{\mathrm{eff}} / D_m \rightarrow \infty \) when \(c_v \rightarrow c_{v, \mathrm{max}}\) because touching particles provide pathways with negligible resistance for diffusive transport through the membrane, corresponding to a percolation threshold. Other commonly employed models for composite materials are reviewed by Petropoulos [3].

Appendix B: Effective diffusivity of a dilute suspension of spheres with an interfacial contact resistance

Consider a sphere of radius a composed of filler material that is suspended in a matrix with a linear ambient concentration field of strength G, such that the solute concentration far from the sphere is

$$\begin{aligned} \psi ({\mathbf {x}}) = {\mathbf {x}}\cdot \varvec{}{G} \qquad \mathrm {as} \quad | {\mathbf {x}}| \rightarrow \infty , \end{aligned}$$
(B.1)

where \(\varvec{}{G}\) is a specified gradient with \(| \varvec{}{G} | = G\) and \({\mathbf {x}}\) is the position vector with respect to the center of the sphere. At the interface between the spherical filler and surrounding matrix, the required conditions are continuity of the diffusive flux according to (4) and a jump in the concentration determined by the contact resistance according to (39). These conditions are imposed at \(| {\mathbf {x}}| = a\). The Laplace equation governing the concentration distribution outside and inside the sphere can be solved by constructing scalar functions that are linear in \(\varvec{}{G}\) from vector harmonic functions and applying the above boundary conditions to determine the three unknown constants,

$$\begin{aligned} \psi ({\mathbf {x}}) = \left( 1 - \beta \frac{a^3}{| {\mathbf {x}}|^3} \right) \, {\mathbf {x}}\cdot \varvec{}{G}, \qquad \phi ({\mathbf {x}}) = \frac{1+2 \beta }{\alpha } \, {\mathbf {x}}\cdot \varvec{}{G}, \end{aligned}$$
(B.2)

where

$$\begin{aligned} \beta = \frac{k \alpha - (1+r_c)}{k \alpha + 2(1+r_c)}, \end{aligned}$$
(B.3)

k is the interfacial partition coefficient, \(\alpha = D_f/D_m\) is the filler-to-matrix ratio of the diffusivity, and

$$\begin{aligned} r_c = \frac{D_{f}}{a k_c} = \frac{\alpha D_{m}}{a k_c } \end{aligned}$$
(B.4)

is a dimensionless contact resistance for mass transfer across the matrix–filler interface. The interfacial contact coefficient, \(k_c\), is defined as the ratio of the interfacial molar flux, \(J^*\), to the concentration difference across the interface according to

$$\begin{aligned} J^* = k_c ( k \psi - \phi ) = D_m {\mathbf {n}}^{(1)} \cdot \nabla \psi . \end{aligned}$$
(B.5)

Using Maxwell’s formulation, the scaled effective diffusivity of a dilute dispersion of spherical fillers is then given approximately by

$$\begin{aligned} \frac{D_{\mathrm{eff}}}{D_m} \simeq \frac{1 + 2 \beta c_v}{1 - \beta c_v}. \end{aligned}$$
(B.6)

In the limit of a vanishing contact resistance, \(r_c \rightarrow 0\), (B.6) reduces to (A.2).

Appendix C: One-dimensional layer model for diffusion through a composite medium with nonlinear interfacial kinetics

Consider steady, one-dimensional diffusion through the matrix of a membrane confined between two planes located at \(z = \pm \delta \). A filler layer is contained within the matrix and occupies the space \(-a \le z \le a\), with \(a < \delta \), as illustrated in Fig. 1. The solute concentration fields in the matrix, \(\psi (z)\), and filler, \(\phi (z)\), satisfy Laplace’s equation,

$$\begin{aligned} \frac{\mathrm{{d}}^2 \psi }{\mathrm{{d}} z^2} = 0, \qquad \frac{\mathrm{{d}}^2 \phi }{\mathrm{{d}}z^2}=0. \end{aligned}$$
(C.1)

The solute concentration is specified at the boundaries of the matrix,

$$\begin{aligned} \psi (-\delta ) = \psi _{-}, \qquad \psi (\delta ) = \psi _{+}. \end{aligned}$$
(C.2)

At the interface between the filler and matrix, there is a jump in the solute concentration according to (42),

$$\begin{aligned} \phi = k_F' \psi ^m \quad \mathrm {at} \quad z=\pm a, \end{aligned}$$
(C.3)

and continuity of the diffusive flux,

$$\begin{aligned} D_m \left. \frac{\mathrm{{d}} \psi }{\mathrm{{d}}z} \right| _{z=\pm a} = D_f \left. \frac{\mathrm{{d}} \phi }{\mathrm{{d}} z} \right| _{z=\pm a}. \end{aligned}$$
(C.4)

Using elementary methods, the solution of the boundary-value problem is found to be

$$\begin{aligned} \psi (z)= & {} \psi _{-} + \left[ \psi _+ - \psi (a) \right] \left( \frac{\delta +z}{\delta -a} \right) , \quad -\delta \le z \le -a, \nonumber \\ \phi (z)= & {} \frac{1}{2} \left[ \phi (a) + \phi (-a) \right] + \frac{z}{2a} \left[ \phi (a) - \phi (-a) \right] , \quad -a \le z \le a, \nonumber \\ \psi (z)= & {} \psi _+ - \left[ \psi _+ - \psi (a) \right] \left( \frac{\delta -z}{\delta -a} \right) , \qquad a \le z \le \delta , \end{aligned}$$
(C.5)

where \(\psi (a)\) is the matrix-phase concentration at \(z=a\) and is given implicitly by

$$\begin{aligned} \frac{(\delta - a) k_F' \alpha }{2a} \left[ \psi ^m(a) - \left( \psi _+ + \psi _{-} - \psi (a) \right) ^m \right] = \psi _+ - \psi (a). \end{aligned}$$
(C.6)

The interfacial filler-phase concentrations are

$$\begin{aligned} \phi (a) = k_F' \psi ^m(a), \qquad \phi (-a) = k_F' \left[ \psi _+ + \psi _{-} - \psi (a) \right] ^m, \end{aligned}$$
(C.7)

where \(\alpha = D_f/D_m\).

Using Fick’s law, the diffusive flux across the layer is

$$\begin{aligned} J^* = -D_m \left. \frac{\mathrm{{d}} \psi }{\mathrm{{d}} z} \right| _{z=\delta } = \frac{D_m}{\delta -a} \left[ \psi (a) - \psi _+ \right] . \end{aligned}$$
(C.8)

An effective diffusivity of the composite system can be defined using the integrated form of Fick’s law,

$$\begin{aligned} J^* = -D_{\mathrm{eff}} \left( \frac{\psi _+ - \psi _{-}}{2 \delta } \right) . \end{aligned}$$
(C.9)

Using (C.8) and (C.9), the scaled effective diffusivity of the composite layer is

$$\begin{aligned} \frac{D_{\mathrm{eff}}}{D_m} = \frac{2 \delta }{\delta -a} \left[ \frac{ \psi (a)-\psi _{+}}{\psi _{-} - \psi _+} \right] . \end{aligned}$$
(C.10)

Without loss of generality, the parameter values are chosen to be \(2 \delta =1\), \(\psi _{-}=1\), and \(\psi _+=0\), which gives

$$\begin{aligned} \frac{D_{\mathrm{eff}}}{D_m} = \frac{2 \psi (a)}{1-2a}. \end{aligned}$$
(C.11)

With the other parameters fixed, the scaled effective diffusivity is nonmonotonic in m and is maximized for

$$\begin{aligned} m = m^* \equiv \frac{\ln \left[ \ln \left( 1 - \psi (a) \right) / \ln \left( \psi (a) \right) \right] }{\ln \left[ \psi (a) / \left( 1 - \psi (a) \right) \right] }. \end{aligned}$$
(C.12)

As shown in Fig. 10, \(m^* \approx \sqrt{2}\) for most values of \(\psi (a)\), which must be computed from (C.6).

Using (C.6) and (C.11), the scaled effective diffusivity is seen to be identical for \(m=1\) and \(m=2\) for this one-dimensional layer model with the aforementioned choices of \(\psi _+\) and \(\psi _{-}\),

$$\begin{aligned} \frac{D_{\mathrm{eff}}}{D_m} = \frac{\delta k_F' \alpha }{(\delta -a) k_F' \alpha +a } \quad \mathrm {if} \quad m=1 ~ \mathrm {or} ~ 2. \end{aligned}$$
(C.13)

The volume fraction of filler for this layer model is \(c_v = a/\delta \).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Davis, J.M. Effective diffusivity for a mixed-matrix membrane. J Eng Math 130, 14 (2021). https://doi.org/10.1007/s10665-021-10176-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10665-021-10176-6

Keywords

Navigation