Skip to main content
Log in

A Feasibility Study of the Pore Topology Method (PTM), A Medial Surface-Based Approach to Multi-phase Flow Simulation in Porous Media

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

Abstract

Computationally efficient microscale models designed to simulate multi-phase flow and characterize low-porosity media are challenged in thin, highly porous materials, primarily due to large, irregular pore spaces and inability to satisfy representative elementary volume requirements. In this article, we describe the pore topology method (PTM) and explore its capabilities to characterize a set of isotropic fibrous materials and to simulate multi-phase flow. PTM is a fast, algorithmically simple method that reduces the complexity of the 3-D void space geometry to its topologically consistent medial surface and uses it as a solution domain for single- and multi-phase flow simulations. Our results in permeability calculations, pore size distribution, and quasi-static drainage and imbibition simulations are in very good agreement with other numerical methods and analytical solutions. We expect that incorporating detailed spatial information about the porous media structure into the medial surface will enable a more accurate representation of the void space structure and of physical phenomena involved in multi-phase flow, thus expanding the applicability of PTM to a broader range of porous media, including non-fibrous materials.

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.

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

Similar content being viewed by others

References

  • Ahrenholz, B., Tölke, J., Lehmann, P., Peters, A., Kaestner, A., Krafczyk, M., Durner, W.: Prediction of capillary hysteresis in a porous material using lattice-Boltzmann methods and comparison to experimental data and a morphological pore network model. Adv. Water Resour. 31, 1151–1173 (2008)

    Article  Google Scholar 

  • Bazylak, A., Berejnov, V., Markicevic, B., Sinton, D., Djilali, N.: Numerical and microfluidic pore networks: towards designs for directed water transport in GDLs. Electrochim. Acta 53, 7630–7637 (2008)

    Article  Google Scholar 

  • Blum, H.: A transformation for extracting new descriptors of shape. Models Percept. Speech Vis. Form 19, 362–380 (1967)

    Google Scholar 

  • Bouix, S., Siddiqi, K., Tannenbaum, A.: Flux driven automatic centerline extraction. Med. Image Anal. 9, 209–221 (2005)

    Article  Google Scholar 

  • Corbett, P., Hayashi, F., Alves, M., Jiang, Z., Wang, H., Demyanov, V., Machado, A., Borghi, L., Srivastava, N.: Microbial carbonates: a sampling and measurement challenge for petrophysics addressed by capturing the bioarchitectural components. Geol. Soc. Lond. Spec. Publ. 418, 69–85 (2015)

    Article  Google Scholar 

  • Frette, O.I., Helland, J.O.: A semi-analytical model for computation of capillary entry pressures and fluid configurations in uniformly-wet pore spaces from 2D rock images. Adv. Water Resour. 33, 846–866 (2010)

    Article  Google Scholar 

  • Gervais, P.-C., Bardin-Monnier, N., Thomas, D.: Permeability modeling of fibrous media with bimodal fiber size distribution. Chem. Eng. Sci. 73, 239–248 (2012)

    Article  Google Scholar 

  • Ghassemzadeh, J., Hashemi, M., Sartor, L., Corp, A.D., Dennison, A.: Pore network simulation of imbibition into paper during coating: I. Model development. AIChE J. 47, 519–535 (2001)

    Article  Google Scholar 

  • Ghassemzadeh, J., Sahimi, M.: Pore network simulation of fluid imbibition into paper during coating: II. Characterization of paper’s morphology and computation of its effective permeability tensor. Chem. Eng. Sci. 59, 2265–2280 (2004)

    Article  Google Scholar 

  • Gostick, J.T.: Random pore network modeling of fibrous PEMFC gas diffusion media using Voronoi and Delaunay tessellations. J. Electrochem. Soc. 160, F731–F743 (2013)

    Article  Google Scholar 

  • Gostick, J.T., Ioannidis, M.A., Fowler, M.W., Pritzker, M.D.: Pore network modeling of fibrous gas diffusion layers for polymer electrolyte membrane fuel cells. J. Power Sources 173, 277–290 (2007)

    Article  Google Scholar 

  • Hazlett, R.: Simulation of capillary-dominated displacements in microtomographic images of reservoir rocks. Transp. Porous Media 20, 21–35 (1995)

    Article  Google Scholar 

  • Hilpert, M., Miller, C.: Pore-morphology-based simulation of drainage in totally wetting porous media. Adv. Water Resour. 24, 243–255 (2001)

    Article  Google Scholar 

  • Hinebaugh, J., Bazylak, A.: Condensation in PEM fuel cell gas diffusion layers: a pore network modeling approach. J. Electrochem. Soc. 157, B1382–B1390 (2010)

    Article  Google Scholar 

  • Jiang, Z., van Dijke, R., Geiger, S., Couples, G., Wood, R.: Extraction of fractures from 3D rock images and network modelling of multi-phase flow in fracture-pore systems. In: International Symposium of the Society of Core Analysts, Aberdeen, Scotland (2012)

  • Jiang, Z., Wu, K., Couples, G., van Dijke, M.I.J., Sorbie, K.S., Ma, J.: Efficient extraction of networks from three-dimensional porous media. Water Resour. Res. 43, 1–17 (2007)

    Article  Google Scholar 

  • Joekar-Niasar, V., Prodanović, M., Wildenschild, D., Hassanizadeh, S.M.: Network model investigation of interfacial area, capillary pressure and saturation relationships in granular porous media. Water Resour. Res. 46, W06526 (2010)

  • Koido, T., Furusawa, T., Moriyama, K.: An approach to modeling two-phase transport in the gas diffusion layer of a proton exchange membrane fuel cell. J. Power Sources 175, 127–136 (2008)

    Article  Google Scholar 

  • Lee, K., Nam, J., Kim, C.: Pore-network analysis of two-phase water transport in gas diffusion layers of polymer electrolyte membrane fuel cells. Electrochim. Acta 54, 1166–1176 (2009)

    Article  Google Scholar 

  • Lee, K., Nam, J., Kim, C.: Steady saturation distribution in hydrophobic gas-diffusion layers of polymer electrolyte membrane fuel cells: a pore-network study. J. Power Sources 195, 130–141 (2010)

    Article  Google Scholar 

  • Lee, T., Kashyap, R., Chu, C.: Building skeleton models via 3-D medial surface axis thinning algorithms. CVGIP Graph. Models Image Process. 56, 462–478 (1994)

    Article  Google Scholar 

  • Lindquist, W.: The geometry of primary drainage. J. Colloid Interface Sci. 296, 655–668 (2006)

    Article  Google Scholar 

  • Lindquist, W., Venkatarangan, A.: Investigating 3D geometry of porous media from high resolution images. Phys. Chem. Earth A Solid Earth Geod. 24, 593–599 (1999)

    Article  Google Scholar 

  • Lopez, A., Lloret, D., Serrat, J., Villanueva, J.: Multilocal creaseness based on the level-set extrinsic curvature. Comput. Vis. Image Underst. 77, 111–144 (2000)

    Article  Google Scholar 

  • Markicevic, B., Bazylak, A., Djilali, N.: Determination of transport parameters for multiphase flow in porous gas diffusion electrodes using a capillary network model. J. Power Sources 171, 706–717 (2007)

    Article  Google Scholar 

  • Mason, G., Morrow, N.R.: Capillary behavior of a perfectly wetting liquid in irregular triangular tubes. J. Colloid Interface Sci. 141, 262–274 (1991)

    Article  Google Scholar 

  • Mortensen, N., Okkels, F., Bruus, H.: Reexamination of Hagen-Poiseuille flow: shape dependence of the hydraulic resistance in microchannels. Phys. Rev. E 71, 057301 (2005)

    Article  Google Scholar 

  • Nabovati, A., Llewellin, E., Sousa, A.: A general model for the permeability of fibrous porous media based on fluid flow simulations using the lattice Boltzmann method. Compos. A 40, 860–869 (2009)

    Article  Google Scholar 

  • Palakurthi, N.: Direct Numerical Simulation of Liquid Transport Through Fibrous Porous Media. Ph.D. dissertation, University of Cincinnati (2014)

  • Pan, S., Davis, H., Scriven, L.: Modeling moisture distribution and binder migration in drying paper coatings. Tappi J. 78, 127–143 (1995)

  • Prat, M., Agaësse, T.: Thin porous media. In: Vafai, K. (ed.) Handbook of Porous Media (Chapter 4), 3rd edn.Taylor & Francis, London (2015)

  • Price, M.A., Armstrong, C.G., Sabin, M.A.: Hexahedral mesh generation by medial surface subdivision: Part I. Solids with convex edges. Int. J. Numer. Methods Eng. 38, 3335–3359 (1995)

    Article  Google Scholar 

  • Prodanovic, M., Bryant, S.: Physics-driven interface modeling for drainage and imbibition in fractures. SPE J. 14, 532–542 (2009)

    Article  Google Scholar 

  • Prodanović, M., Bryant, S.: A level set method for determining critical curvatures for drainage and imbibition. J. Colloid Interface Sci. 304, 442–458 (2006)

    Article  Google Scholar 

  • Pudney, C.: Distance-ordered homotopic thinning: a skeletonization algorithm for 3D digital images. Comput. Vis. Image Underst. 72, 404–413 (1998)

    Article  Google Scholar 

  • Qin, C., Hassanizadeh, S.: Multiphase flow through multilayers of thin porous media: General balance equations and constitutive relationships for a solid–gas–liquid three-phase system. Int. J. Heat Mass Transf. 70, 693–708 (2014)

  • Riasi, M.S., Huang, G., Montemagno, C., Yeghiazarian, L.: A General 3-D Methodology for Quasi-Static Simulation of Drainage and Imbibition: Application to Highly Porous Fibrous Materials. AGU Fall Meeting Abstracts, San Francisco (2013)

  • Schulz, V.P., Wargo, E.A., Kumbur, E.C.: Pore-morphology-based simulation of drainage in porous media featuring a locally variable contact angle. Transp. Porous Media 107, 13–25 (2015)

    Article  Google Scholar 

  • Si, C., Wang, X., Yan, W., Wang, T.: A comprehensive review on measurement and correlation development of capillary pressure for two-phase modeling of proton exchange membrane fuel cells. J. Chem. 2015, 1–17 (2015)

  • Styner, M., Lieberman, J., Pantazis, D., Gerig, G.: Boundary and medial shape analysis of the hippocampus in schizophrenia. Med. Image Anal. 8, 197–203 (2004)

    Article  Google Scholar 

  • Sun, H., Frangi, A., Wang, H.: Automatic cardiac MRI segmentation using a biventricular deformable medial model. Med. Image Comput. Comput. Assist. Interv. MICCAI 2010, 468–475 (2010)

    Google Scholar 

  • Tahir, M.A., Vahedi Tafreshi, H.: Influence of fiber orientation on the transverse permeability of fibrous media. Phys. Fluids 21, 083604 (2009)

    Article  Google Scholar 

  • Vennat, E., Attal, J.-P., Aubry, D., Degrange, M.: Three-dimensional pore-scale modelling of dentinal infiltration. Comput. Methods Biomech. Biomed. Eng. 17, 632–642 (2014)

  • Vera, S., Gil, D., Borras, A., Sánchez, X., Párez, F., Linguraru, M., González Ballester, M.A.: Computation and evaluation of medial surfaces for shape representation of abdominal organs. In: Yoshida, H. et al. (eds.) Abdominal Imaging 2011, LNCS 7029, pp. 223–230 (2012a)

  • Vera, S., González, M., Gif, D.: A medial map capturing the essential geometry of organs. In: 2012 International Symposium on Biomedical Imaging (2012b)

  • Vogel, H.J., Tölke, J., Schulz, V.P., Krafczyk, M., Roth, K.: Comparison of a lattice-Boltzmann model, a full-morphology model, and a pore network model for determining capillary pressure–saturation relationships. Vadose Zone J. 4, 380–388 (2005)

    Article  Google Scholar 

  • Wildenschild, D., Sheppard, A.: X-ray imaging and analysis techniques for quantifying pore-scale structure and processes in subsurface porous medium systems. Adv. Water Resour. 51, 217–246 (2013)

    Article  Google Scholar 

  • Wu, R., Liao, Q., Zhu, X., Wang, H.: Pore network modeling of cathode catalyst layer of proton exchange membrane fuel cell. Int. J. Hydrog. Energy 37, 11255–11267 (2012)

    Article  Google Scholar 

Download references

Acknowledgments

The authors gratefully acknowledge the National Science Foundation for supporting this research (CBET-1248385 and CBET-1351361 awards), as well as the University of Cincinnati Research Council Graduate Award. Early stages of this work were sponsored by the Procter & Gamble company through the University of Cincinnati Simulation Center. The authors would like to thank Dr. Rodrigo Rosati for comments on the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Lilit Yeghiazarian.

Appendices

Appendix 1: Medial Surface Extraction Algorithm

In this appendix, we demonstrate the medial surface extraction algorithm on a simple, constant cross-sectional prism shown in Fig. 10. The algorithm outlined here is based on the formulation provided in Vera et al. (2012a). As mentioned earlier, we have modified the last step of the algorithm, medial surface thinning, to ensure full connectivity in the final medial surface.

1.1 Step 1: Obtain a Binary Image of Porous Media

Since the medial surface extraction algorithm described here is a voxel-based approach, obtaining a voxelized binary image of porous media is essential. Binary images for porous materials can be obtained either by using 3-D imaging of actual materials, or by using commercial packages like GeoDict. In the binary image shown in Fig. 10, we labeled voxels of the solid matrix as one, and of void space as zero.

Fig. 10
figure 10

3-D model of the prism used in this example and its 2-D binary representation

1.2 Step 2: Generate Euclidean Distance Map (D)

From the binary image of the porous medium, we generate the distance map inside the void space. Distance map D, also known as distance field or distance transform, is a representation of a digital image based on the distance between each voxel and the closest boundary (solid phase). Assuming solid voxels (1-value voxels) as boundary voxel, the distance map of the void space can be generated by computing the distance between each void voxel and the closest solid voxel. Figure 11 shows the distance map of the sample prism.

Fig. 11
figure 11

2-D slice of the distance map generated in the prism of Fig. 10

1.3 Step 3: Compute Gradient of Distance Map

The gradient of distance map can be computed by convolving the distance map with gradients of Gaussian kernel:

$$\begin{aligned} {\nabla }D= \left( \partial _{x}D_{\sigma } ,{ \partial }_{y}D_{\sigma }, {\partial }_{z}D_{\sigma } \right) =\left( \partial _{x}g_{\sigma }*D ,\partial _{y}g_{\sigma }*D ,\partial _{z}g_{\sigma }*D \right) \end{aligned}$$
(8)

where \(\partial _{x}g_{\sigma }\), \(\partial _{y}g_{\sigma }\), and \(\partial _{z}g_{\sigma }\) are partial derivatives of the Gaussian kernel g with standard deviation of \(\sigma \) and \(*\) is the convolution operator. To avoid over-smoothing, we used \(3\times 3\times 3\) voxels window kernels with \(\sigma =0.5\).

1.4 Step 4: Compute Structure Tensor (ST)

The structure tensor (ST) summarizes the dominant directions of the gradients in a neighborhood of each voxel and is derived from the gradient field of a function. Smoothed ST at each voxel can be computed as

$$\begin{aligned} {ST}_{\rho ,\sigma }\left( D \right) =\left( {\begin{array}{ccc} g_{\rho }*{\partial _{x}D_{\sigma }}^{2} &{}\quad g_{\rho }*\partial _{x}D_{\sigma }\partial _{y}D_{\sigma } &{}\quad g_{\rho }*\partial _{x}D_{\sigma }\partial _{z}D_{\sigma }\\ g_{\rho }*\partial _{y}D_{\sigma }\partial _{x}D_{\sigma } &{}\quad g_{\rho }*{\partial _{y}D_{\sigma }}^{2} &{}\quad g_{\rho }*\partial _{y}D_{\sigma }\partial _{z}D_{\sigma }\\ g_{\rho }*\partial _{z}D_{\sigma }\partial _{x}D_{\sigma } &{}\quad g_{\rho }*\partial _{z}D_{\sigma }\partial _{y}D_{\sigma } &{}\quad g_{\rho }*{\partial _{z}D_{\sigma }}^{2}\\ \end{array} } \right) \end{aligned}$$
(9)

where \(g_{\rho }\) is a \(3\times 3\times 3\) Gaussian kernel with standard deviation of \(\rho \). Here we chose \(\rho =1\).

1.5 Step 5: Find Principal Eigenvectors

ST is a positive definite tensor; therefore, it has three nonnegative eigenvalues for a 3-D problem. For this case, since we have built the structure tensor from the gradients of the distance map, the eigenvectors at each voxel are the directions of principal gradients of distance map at the neighborhood of that voxel. To find the direction of the maximum gradient at each voxel, one needs to find the principal eigenvector of ST (i.e., the eigenvector corresponding to the eigenvalue of largest magnitude) at that voxel. For the sample prism of Fig. 10, the 2-D view of the principal eigenvector field \((\mathbf{V})\) is shown in Fig. 12.

Fig. 12
figure 12

2-D view of the principal eigenvector field \(({\mathbf{V}})\) of the prism

1.6 Step 6: Reorient Principal Eigenvectors in the Direction of Gradient

To prepare the vector field for step 7, we need all the principal eigenvectors to point toward the downward gradient. Reorienting the principal eigenvectors toward the downward gradient can be done by multiplying each vector by the sign of the scalar product of principal eigenvector and the gradient at each voxel:

$$\begin{aligned} {\varvec{V}}=\hbox {sign}\left( {\varvec{V}}\cdot {\nabla } D\right) \cdot {\varvec{V}} \end{aligned}$$
(10)

As shown in Fig. 13, besides reorienting the principal eigenvector field, this assigns zero to zero-gradient voxels.

Fig. 13
figure 13

2-D view of the reoriented principal eigenvector field \(({\mathbf{V}})\) of the prism

1.7 Step 7: Compute Normalized Ridge Map (NRM)

Since the magnitude of all vectors in \({\varvec{V}}\) is less than or equal to one, taking the divergence of \({\varvec{V}}\) will assign a value between N and \({+}{N}\) to each voxel, where N is the dimension of the problem. For a 3-D geometry, this value will be bounded between −3 and \(+3\). We will refer to this value as the normalized ridge map (NRM). Positive NRMs correspond to ridge voxels and negative to valley voxels; therefore, the magnitude of NRM value represents the degree of ridgeness or valleyness. NRM can be calculated as

$$\begin{aligned} \hbox {NRM}=\hbox {div}\left( {\varvec{V}} \right) \end{aligned}$$
(11)

A 2-D view of NRM for the sample prism is shown in Fig. 14. For the purpose of medial surface extraction, we are only interested in ridges (positive NRMs); therefore, non-positive values can be ignored at the end of this step.

Fig. 14
figure 14

2-D slice of the normalized ridge map generated for the sample prism

1.8 Step 8: Medial Surface Thinning: Hessian-Based Non-maxima Suppression

The last step is to apply a thinning algorithm on the positive NRM values to extract a voxel-wide medial surface. One of the most common thinning algorithms is non-maxima suppression (NMS). In NMS, one first needs to find a suppression direction, which is the direction perpendicular to the medial surface. The common practice is to use the direction of principal gradient of the NRM at each voxel as the suppression direction. It should be noted that the direction of principal gradient of NRM is the principal eigenvector of the structure tensor of NRM. However, it has been shown that implementing NMS on NRM map will introduce holes (disconnectivities) in the medial surface (Vera et al. 2012a). Although this drawback can be ignored in many image processing cases, for our application it will change the connectivity of the void space of the porous media, which is undesirable. To overcome this drawback, we propose a modified NMS algorithm based on using the direction of maximum curvature of NRM as the suppression direction. The direction of maximum curvature of NRM is the principal eigenvector of Hessian tensor of NRM. Hessian matrix is a symmetric matrix that summarizes the dominant directions of curvatures at a neighborhood of a voxel. Instead of using the structure tensor of NRM, modified NMS uses the Hessian tensor of NRM map to find the suppression direction. Throughout this paper, we will refer to this non-maxima suppression method as the Hessian-based non-maxima suppression (HNMS).

Fig. 15
figure 15

a Suppression directions for Hessian-based NMS \(({\varvec{V}}_{\mathrm{HNMS}})\) and b enlarged view of the area marked by the red square in Fig. 15a

Fig. 16
figure 16

Extracted medial surface for the prism in Fig. 10

In HNMS, the principal eigenvector of Hessian tensor of NRM at each point is chosen as the suppression direction. We computed the Hessian matrix at each voxel as below:

$$\begin{aligned} H_{\rho ,\sigma }\left( \hbox {NRM} \right) =\left( {\begin{array}{ccc} \partial _{xx}g_{\rho }*\hbox {NRM} &{}\quad \partial _{xy}g_{\rho }*\hbox {NRM} &{}\quad \partial _{xz}g_{\rho }*\hbox {NRM}\\ \partial _{yx}g_{\rho }*\hbox {NRM} &{}\quad \partial _{yy}g_{\rho }*\hbox {NRM} &{}\quad \partial _{yz}g_{\rho }*\hbox {NRM}\\ \partial _{zx}g_{\rho }*\hbox {NRM} &{}\quad \partial _{zy}g_{\rho }*\hbox {NRM} &{}\quad \partial _{zz}g_{\rho }*\hbox {NRM}\\ \end{array} } \right) \end{aligned}$$
(12)

where \(\partial _{x_{i}x_{j}}g_{\rho }\) is the second derivative of the Gaussian kernel with respect to coordinates \(x_{i}\) and \(x_{j}\), and \(\rho =1\). Let \({\varvec{V}}_{\mathrm{HNMS}}\) be the principal eigenvector of the Hessian matrix at each voxel. The non-maxima suppression can be computed as

$$\begin{aligned} \hbox {NMS}\left( x,y,z \right) =\left\{ {\begin{array}{ll} 1 &{} \quad \hbox {if NRM}\left( x,y,z \right) >\hbox {max}\left\{ \hbox {NRM} \left( {\varvec{V}}_{\mathrm{HNMS}}^{+}\right) , \hbox {NRM}\left( {\varvec{V}}_{\mathrm{HNMS}}^{-} \right) \right\} \\ 0&{}\quad \hbox {otherwise} \\ \end{array}} \right. \qquad \end{aligned}$$
(13)

where \(\hbox {NRM}\left( {\varvec{V}}_{\mathrm{HNMS}}^{+} \right) \) and \(\hbox {NRM}\left( {\varvec{V}}_{\mathrm{HNMS}}^{-} \right) \) are NRM values in the immediate neighborhood of a voxel in the direction of \({\varvec{V}}_{\mathrm{HNMS}}\).

Figure 15 shows the suppression directions in the proposed Hessian-based NMS. It should be noted that the ideal suppression direction should be perpendicular to the medial surface. Finally, Fig. 16 shows 2-D and 3-D views of the extracted medial surface for the sample prism.

Appendix 2: Hydraulic Conductivity of Prisms

To test the validity of the planar flow assumption in computing local conductivities in Eq. (2), we computed hydraulic conductivities for a set of prisms with polygonal cross sections using the PTM permeability calculation algorithm and compared the result with exact solutions derived by Mortensen et al. (2005). We followed the procedure outlined in Sect. 2.2, but instead of Eq. (4), we used the following expression for the hydraulic conductivity K of the prism:

$$\begin{aligned} {K} = -\frac{{QL}}{\Delta {P}} \end{aligned}$$
(14)

As in Mortensen et al. (2005), we defined the dimensionless hydraulic resistivity or correction factor \({\alpha }\), and dimensionless compactness C, to represent our results. These parameters can be computed as follows:

$$\begin{aligned} \alpha= & {} \frac{A^{2}}{K\mu } \end{aligned}$$
(15)
$$\begin{aligned} C= & {} \frac{p}{A^{2}} \end{aligned}$$
(16)

where A and p are the cross-sectional area and cross-sectional perimeter of the prism, respectively; K is hydraulic conductivity; and \(\mu \) is fluid viscosity. Figure 17 compares the PTM results with exact values for \(\alpha \) and C computed in Mortensen et al. (2005), for rectangular and triangular prisms. As the compactness increases, the PTM results approaches the exact solution, which is consistent with our assumption of flow between parallel plates for all voxels on medial surface.

Fig. 17
figure 17

Correction factor versus dimensionless compactness for triangular and rectangular prisms

Appendix 3: Discretization of Eq. (1)

As mentioned in Sect. 2.2, to find the pressure field inside the void space, the following equation is discretized and solved on the medial surface of the void space:

$$\begin{aligned} {\nabla }\cdot \left( K_{\mathrm{local}}\nabla \,{P} \right) =0 \end{aligned}$$
(17)

where \(K_{\mathrm{local}}\) is the local hydraulic conductivity at each voxel of the medial surface, and its value is zero on all non-medial surface voxels in the porous media domain.

Since we are aiming to discretize and solve this equation only on the medial surface, we have used a modified version of the well-known central differencing scheme used in finite volume method (FVM).

Following the finite volume approach, we start the discretization by taking the integral of both sides of Eq. (17) over the volume of a voxel (V).

$$\begin{aligned} \int _V {{\nabla }\cdot \left( K_{\mathrm{local}}\nabla {P} \right) \hbox {d}V } =0 \end{aligned}$$
(18)

Then, using the divergence theorem, we can change the above volume integral to a surface integral:

$$\begin{aligned} \int _S {\left( K_{\mathrm{local}}\nabla {P} \right) \cdot {\varvec{n}}\hbox {d}S} =0 \end{aligned}$$
(19)

So far, we have been followed the exact procedure that one would take in FVM. Then, instead of using the six faces of our cubic voxels as our integration surfaces, we assumed an imaginary rhombicuboctahedron with 26 equal-area faces. Using this analogy, instead of three gradient directions, we will have 13 directions:

$$\begin{aligned} \mathop {\sum }\limits _{i=1}^{13} \left[ \left[ K_{\mathrm{local}}\frac{\partial P}{\partial {\varvec{n}}_{{\varvec{i}}}} \right] _{i^{+}}-\left[ K_{\mathrm{local}}\frac{\partial P}{\partial {\varvec{n}}_{{\varvec{i}}}} \right] _{i^{-}} \right] =0 \end{aligned}$$
(20)

where \({\varvec{n}}_{{\varvec{i}}}\) is the gradient direction and \(i^{+}\) and \(i^{-}\) shows the opposite faces along the gradient direction. This assumption is essential in capturing all curvatures and the junctions that are present in the voxel-wide medial surface.

Further simplification of Eq. (20) will result:

$$\begin{aligned}&\mathop {\sum }\limits _{i=1}^{13} \left[ \left[ \left( K_{\mathrm{local},i+1}+K_{\mathrm{local},i} \right) \left( P_{i+1}-P_{i} \right) \right] \right. \hbox {sign}(K_{\mathrm{local},i+1}) \nonumber \\&\quad \left. -\left[ \left( K_{\mathrm{local},i}+K_{\mathrm{local},i-1} \right) \left( P_{i}-P_{i-1} \right) \right] \hbox {sign}(K_{\mathrm{local},i-1}) \right] =0 \end{aligned}$$
(21)

In Eq. (21), sign(\(\cdot \)) is used to automatically zero out the gradient between a medial surface voxel and a non-medial surface voxel.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Riasi, M.S., Palakurthi, N.K., Montemagno, C. et al. A Feasibility Study of the Pore Topology Method (PTM), A Medial Surface-Based Approach to Multi-phase Flow Simulation in Porous Media. Transp Porous Med 115, 519–539 (2016). https://doi.org/10.1007/s11242-016-0720-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11242-016-0720-0

Keywords

Navigation