1 Introduction

Thin, highly porous materials are ubiquitous in nature and widely employed in many products and devices. Examples range from living tissues, filters, membranes, and absorbent materials to fuel cells and microfluidic devices, driving the need to better understand the structure and processes in these materials. Their distinct properties, however, present new challenges in experimental and numerical characterization.

Thin, highly porous materials are characterized by thickness on the order of pore dimension (Prat and Agaësse 2015) and by porosity of larger than 60 % (Gervais et al. 2012; Gostick 2013; Si et al. 2015). Therefore, representative elementary volume (REV) requirements cannot be satisfied, and macroscale theories of transport do not hold (Qin and Hassanizadeh 2014). Furthermore, large, irregular pore spaces preclude easy characterization of the pore space geometry. Microscale models of flow in thin, highly porous media are promising; however, their implementation is not trivial. Some of the challenges include computational cost and convergence in direct numerical simulation methods (DNS) (e.g., Ahrenholz et al. 2008; Palakurthi 2014; Vennat et al. 2014), while computationally fast techniques (e.g., Bazylak et al. 2008; Ghassemzadeh et al. 2001; Ghassemzadeh and Sahimi 2004; Gostick et al. 2007; Hinebaugh and Bazylak 2010; Koido et al. 2008; Lee et al. 2010, 2009; Markicevic et al. 2007; Pan et al. 1995; Wu et al. 2012) need to overcome difficulties in capturing the pore size distribution (Gostick et al. 2007) and computing respective entry capillary pressure (Joekar-Niasar et al. 2010; Lindquist 2006; Prodanović and Bryant 2006), as well as in detecting the correct location of fluid interfaces for a range of contact angles (Frette and Helland 2010; Schulz et al. 2015).

To address these challenges, we have conducted a feasibility study on using medial surfaces as both the representative geometry and the solution domain for multi-phase flow simulations. We refer to this approach as the pore topology method (PTM). Medial surfaces are widely used in biomedical imaging with applications in surface reconstruction and dimension reduction of complex objects (Styner et al. 2004; Sun et al. 2010; Vera et al. 2012b). In porous media research, they have been suggested as a powerful tool to gain insight into complicated geometries (Prodanovic and Bryant 2009) and as the correct approach to pore space representation, likely to improve on pore network modeling and on morphological methods in capturing sheet-like pores (Wildenschild and Sheppard 2013). In this direction, Jiang et al. (2012) extended their skeletonization-based pore network extraction algorithm (Jiang et al. 2007) and developed a modified version of a medial surface to locate and quantify fractures. In their work, the medial surface was used as an intermediate step in generating a pore network model. The medial surface was converted to virtual nodes and bonds and then added to a traditional pore network model, hoping to better represent fractures in a pore network model of a fractured porous media. Later, Corbett et al. (2015) used this approach to characterize microbial carbonates. To authors’ knowledge, a direct use of medial surface has not been reported in porous media literature, neither as a representative geometry nor as the solution domain.

Here, we expand on our recent work (Riasi et al. 2013) and apply PTM to a series of isotropic fibrous materials with different porosities (25–95 %). We have extracted the medial surface of the void space, derived the pore size distribution, computed the absolute permeability, and simulated quasi-static drainage and imbibition using the medial surface as the solution domain. Comparison of results with lattice Boltzmann, full morphology, volume of fluid and finite volume methods, as well as exact solutions, showed a very good fit.

The outline of this article is as follows: We present the definition and properties of a medial surface, followed by a brief description of the extraction algorithm, in Sect. 2.1. Details of the algorithm are presented in “Appendix 1.” We describe absolute permeability calculation in Sect. 2.2 and quasi-static drainage and imbibition simulations in Sect. 2.3. We present and discuss our results in Sects. 3 and 4 and conclude the paper with Sect. 5.

2 Methods

2.1 Medial Surface

2.1.1 Definition and Properties

Medial surfaces of 3-D objects were first proposed by Blum (1967) as shape descriptors and defined as “the loci of center of maximal spheres bi-tangent to the surface boundary points of the shape.” Along the same lines, Price et al. (1995) described them as “the locus of the center of an inscribed sphere of maximal diameter as it rolls around the interior of the object.” For differences between medial axis and medial surface, please refer to classic works of Lee et al. (1994) and Lindquist and Venkatarangan (1999).

An ideal medial surface of an object satisfies three conditions (Vera et al. 2012a): (i) homotopy: preserving the topology of the original object, (ii) thinness and connectivity: generating a fully connected, voxel-wide medial surface; and (iii) medialness: generating a medial surface that is as close as possible to the center of the object in all directions. These properties collectively stipulate that the medial surface of an object offers its compact representation, while retaining sufficient local information to reconstruct the object fully and with high accuracy. If the object in question is the void space of a porous medium, then homotopy and medialness of its medial surface would ensure that all corners and curvatures are captured, while thinness and connectivity would ensure that pore connectivity in void space is preserved.

2.1.2 Medial Surface Extraction Algorithm

Several approaches have been proposed to extract a voxel-wide medial surface of a 3-D object, mostly based on iterative thinning of an energy map; one example is distance maps (Bouix et al. 2005; Lee et al. 1994; Pudney 1998). We used Multilocal Level-Set Extrinsic Curvature based on the Structure Tensor (MLSEC-ST) developed by Lopez et al. (2000) as described in Vera et al. (2012a). MLSEC-ST is a non-iterative analytical approach that is shown to be less sensitive to noise, have higher reconstructive power compared to other available algorithms, and produce surfaces that closely satisfy the conditions of an ideal medial surface (Vera et al. 2012a). We have modified the last step of the algorithm, medial surface thinning, to ensure full connectivity in the final medial surface. A detailed description of the modified MLSEC-ST algorithm applied to a simple prism (Fig. 1) is given in “Appendix 1.”

Fig. 1
figure 1

A 3-D model of a constant cross-sectional prism (a) and its medial surface extracted using the modified MLSEC-ST method (b)

2.2 Absolute Permeability Calculation

To compute absolute transverse permeability, we assumed a steady-state, laminar, single-phase flow through the void space of porous medium. To compute the fluid pressure (P) distribution inside the void space, we solved the following equation on the medial surface using the finite volume method:

$$\begin{aligned} {\nabla }\cdot \left( K_{\mathrm{local}}{\nabla } P \right) =0 \end{aligned}$$
(1)

where \(K_{\mathrm{local}}\) is the local hydraulic conductivity at each voxel of the medial surface. To determine \(K_{\mathrm{local}}\), we assumed that single-phase flow at each voxel was influenced mainly by its nearest solid walls and therefore used the Poiseuille equation for flow between parallel plates. To test the validity of the planar flow assumption, we computed hydraulic conductivities for a set of prisms with polygonal cross sections and compared the result with exact solutions (“Appendix 2”). Note that for each voxel, all 26 neighboring voxels should be included in discretizing Eq. (1) (see detailed discretization in “Appendix 3”).

Solving the Navier–Stokes equation for single-phase, steady-state, fully developed, laminar, pressure-driven flow between two parallel plates, we obtained the following expression for the local hydraulic conductivity at each voxel:

$$\begin{aligned} K_{\mathrm{local}} = \frac{2}{3}({br}^{3}/\mu ) \end{aligned}$$
(2)

where \(\mu \) is the fluid viscosity, b is the voxel size, and r is the distance between the medial surface voxel and the closest solid boundary. The boundary conditions for Eq. (2) are similar to Fig. 2, but with reservoirs replaced with two non-equal wetting phase pressure conditions.

From the pressure field inside the void space, the flow rate can be calculated at the boundary of the porous medium:

$$\begin{aligned} Q=\mathop {\sum }\limits _{\mathrm{boundary\,voxels}} \left( -K_{\mathrm{local}} \frac{\partial P}{\partial x} \right) \end{aligned}$$
(3)

The absolute permeability k can then be calculated as

$$\begin{aligned} k = -\frac{Q\mu L}{A{\Delta } P} \end{aligned}$$
(4)

where L is the length of porous medium, \(A=w_{1}\times w_{2}\) (Fig. 2) is cross-sectional area, and \({\Delta } P\) is the applied pressure difference across the medium.

Fig. 2
figure 2

Schematic of problem setup for quasi-static drainage and imbibition simulation and permeability calculation. Boundary conditions include non-wetting and wetting phase reservoirs at the top and the bottom, and no-flux conditions on sides

2.3 Quasi-Static Drainage and Imbibition

The Young–Laplace equation states that capillary pressure difference across the interface of two static fluids \(P^{c}\) is a function of surface tension \(\sigma \) and the mean curvature of the interface H:

$$\begin{aligned} P^{c}= \sigma H= \sigma \left( \frac{1}{R_{1}}+\frac{1}{R_{2}}\right) \end{aligned}$$
(5)

where \(R_{1}\) and \(R_{2}\) are the principal radii of curvature of the interface. For a finite, positive \(P^{c}\), one can identify two opposite cases: (i) the spherical interface case, where \(R_{1}=R_{2}=R\) and therefore \(P^{c}= \frac{2\sigma }{R};\) (ii) the cylindrical interface case, where \(R_{1}=R\) and \(R_{2}=\infty \) and, therefore, \(P^{c}= \frac{\sigma }{R}\). In reality, quasi-static interfaces span over this range (Vogel et al. 2005). Here, we use the average \(P^{c}\) of these two cases to approximate the entry capillary pressure at each voxel of the medial surface, yielding \(P^{c}= \frac{1.5 \sigma }{R}\). Assuming that the smallest radius of curvature R is related to contact angle \(\theta \) through \(R=\frac{r}{\mathrm {cos}(\theta )}\), where r is the distance between the medial surface voxel and the closest solid boundary, we obtain:

$$\begin{aligned} P^{c}= \frac{1.5\,\sigma \,\mathrm {cos}\,(\theta )}{r} \end{aligned}$$
(6)

To test the validity of this assumption, we compared our drainage and imbibition results for prisms with triangular and rectangular cross sections, with the analytical results of Mason and Morrow (1991), with excellent agreement (not presented). To simulate quasi-static drainage and imbibition, medial surfaces of porous media samples were extracted from binary images as described in “Appendix 1,” and an entry capillary pressure \(P_{\mathrm{entry}}^{c}\) was assigned to each voxel on the medial surface using Eq. (6). For the setup shown in Fig. 2, the percolation of wetting phase (imbibition) or non-wetting phase (drainage) was simulated. For drainage, the wetting phase reservoir pressure was assumed to be zero and initially equal to the pressure of the non-wetting phase reservoir. To initiate drainage, the non-wetting phase pressure was increased until it exceeded the entry capillary pressure of voxels connected to non-wetting phase reservoir. At this pressure, the non-wetting phase can occupy voxels connected to the non-wetting phase reservoir, if \({P}_{\mathrm {entry}}^{{c}}<({P}^{\mathrm {nw}}-{P}^{{w}})\), where \({P}^{\mathrm {nw}}\) is the non-wetting phase reservoir pressure, and \({P}^{\mathrm {w}}\) is the wetting phase reservoir pressure. The drainage simulation proceeded by increasing the non-wetting phase pressure in increments, checking for the above-mentioned condition and propagating the non-wetting phase accordingly.

Imbibition simulation started from a large pressure difference \(({P}^{\mathrm {nw}}-{P}^{{w}})\) corresponding to \({\sim }0\,\%\) wetting phase saturation and continued by decreasing the non-wetting phase pressure in increments. At each pressure difference, the wetting phase can occupy voxels connected to the wetting reservoir, if \({P}_{\mathrm {entry}}^{{c}}>({P}^{\mathrm {nw}}-{P}^{{w}})\). As evident from both drainage and imbibition algorithms, we assumed that no trapping occurred. We also ignored merging and splitting of interfaces.

For both drainage and imbibition, wetting phase saturation \(S^{w}\) was computed at each pressure difference:

$$\begin{aligned} S^{w}=\frac{\sum _i r_{i}}{\sum _j r_{j} } \end{aligned}$$
(7)

where r is the minimum distance between each voxel of the medial surface and nearby solid clusters, \(i=1,\ldots ,I\) where I is the number of medial surface voxels occupied by wetting phase, and \(j=1,\ldots ,J\) where J is the total number of voxels on medial surface. Alternatively, one can compute the saturation values by reconstructing the saturated 3-D void space from the medial surface voxels that are occupied by wetting phase during drainage and imbibition, albeit at a much higher computational time. Our numerical investigations (not presented) showed that despite its simplicity, Eq. (7) produced negligible error compared to the 3-D volume reconstruction approach.

3 Results

We implemented PTM on six isotropic homogenous fibrous materials with different porosities. Fibrous is referred to materials made from natural or synthetic fibers. All structures were generated using the GeoDict software package. Table 1 below summarizes the geometrical and physical parameters of virtual porous media used in this study. Figure 3 shows a sample of fibrous material and its medial surface, and Fig. 4 shows cross sections of the virtual fibrous materials used in this study.

Table 1 Geometrical and physical parameters of fibrous samples used in this study
Fig. 3
figure 3

a Virtual 80 % porous fibrous material generated with GeoDict; b medial surface; c sliced image showing fibers (gray), medial surface (white) and void space (black); d medial surface colored according to its distance map values; e, f sliced image for 25 % porous fibrous material colored similar to c, d

Fig. 4
figure 4

2-D cross sections of the virtual porous media used in this study

3.1 Pore Size Distribution

Pore size distributions were computed in GeoDict and PTM for all material samples, using porosimetric and geometric methods (Fig. 5). Porosimetric pore size distribution analysis is based on numerical modeling of the intrusion of a non-wetting fluid (mercury) into the void space of an air-occupied porous medium. GeoDict uses the full morphology (FM) method and PTM—the drainage algorithm described in Sect. 2.3.

The geometric pore size distribution in GeoDict is obtained by fitting spheres into the void space of the porous structure. In PTM, each voxel of the medial surface is associated with a voxel-wide square prism with height of 2r, where r is the minimum distance between each voxel and nearby solid clusters. Each prism represents a pore. The volume fraction is then calculated similarly to Eq. (7).

Fig. 5
figure 5

Pore size distributions computed with GeoDict and PTM, using porosimetric and geometric methods, for materials ranging in porosity from \({\varphi } =0.25\) to \({\varphi } =0.95\)

Figure 5 shows a rather critical problem associated with current pore-based methodologies such as pore network modeling and full morphology: the definition of pore. Although characterizing the pores through sphere fitting is an appropriate method for low-porosity, non-fractured porous media, applying the same concept to highly porous materials, where high-aspect-ratio pores are abundant, is not reasonable. As shown in this figure, PTM pore size distributions closely match the trend of the GeoDict pore size distribution curves for low-porosity media. But as the porosity increases, the discrepancy increases.

Another observation in Fig. 5 is the oscillations in the results of both GeoDict and PTM for higher porosity materials, which are mostly due to the wider pore size distribution in these materials. Note that same resolution has been used for all cases. If we decrease the resolution, some of these oscillations will vanish.

3.2 Absolute Permeability

Figure 6 compares permeability computed in PTM with three other studies on isotropic fibrous materials. These works include that of Nabovati et al. (2009), in which they derived an empirical equation for absolute permeability of isotropic fibrous materials based on their lattice Boltzmann simulation results and works of Tahir and Vahedi Tafreshi (2009) and Palakurthi (2014), who performed finite volume method simulations with FLUENT and OpenFOAM packages, respectively. PTM is in excellent agreement with these numerical results.

Fig. 6
figure 6

Absolute transverse permeability versus porosity

3.3 Quasi-Static Drainage and Imbibition

Figure 7 shows the capillary pressure–saturation curves for primary drainage and imbibition for all six fibrous material samples using PTM and FM. The FM simulations have been performed using the GeoDict software package. FM uses morphological operations to determine the flow (or open) paths inside the pore space (Hazlett 1995; Hilpert and Miller 2001). The results are in very good agreement; errors range from 1.14 to 16.51 %. The FM method produces drainage results comparable to the Lattice–Boltzmann method and was chosen for comparison due to its low computational cost and ease of implementation (Vogel et al. 2005). To capture small pores in samples of 25 and 50 % porosity, we increased the resolution from 4 to 2 \({\upmu }\hbox {m}\).

Fig. 7
figure 7

Capillary pressure–saturation curve for a primary drainage and b primary imbibition in six GeoDict samples

The values of capillary pressure versus porosity at 50 % saturation are shown in Fig. 8. This plot clearly shows that PTM results are in excellent agreement with FM, particularly for drainage. Note that full morphology method is known for its accuracy in drainage simulations, but not in imbibition (Vogel et al. 2005). Figure 8 also shows that PTM can capture the geometrical hysteresis that is present in capillary pressure–saturation curves.

Fig. 8
figure 8

Capillary pressure–saturation curve for primary drainage in an 80  % porosity isotropic fibrous GeoDict sample using volume of fluid (VOF), full morphology (FM), and pore topology method (PTM)

Finally, for a material of 80 % porosity, we compared primary drainage curves generated with volume of fluid (VOF) (Palakurthi 2014), FM and PTM (Fig. 9), with excellent agreement.

Fig. 9
figure 9

Capillary pressure–saturation curve for primary drainage using volume of fluid (VOF), full morphology (FM), and pore topology method (PTM)

4 Discussion

This work was inspired by challenges in characterization and flow simulations in thin, highly porous media. Our goal was to develop a fast method for microscale modeling in thin, highly porous materials, but potentially applicable to porous materials with any porosity and void space geometry. In this study, we probed the suitability of PTM as a possible candidate. Through numerical experiments presented here, we sought to answer two questions: (i) does the medial surface of the void space extracted with PTM’s modified MLSEC-ST algorithm possess high reconstructive power and, therefore, serve as a meaningful representative geometry for porous media? If yes, (ii) can this geometry be directly used as the solution domain for multi-phase flow simulations? Our results to date provide positive answers to both inquiries.

One way to assess how well a medial surface represents the structure of the void space is to estimate the statistical properties of the void space based on the information contained in the medial surface. Figure 5 shows the geometric and porosimetric pore size distributions for all six materials, computed with GeoDict and PTM. Overall, both geometric and porosimetric pore size distribution curves derived via PTM follow the same trend and spectrum as those derived using GeoDict software package. The shift toward the lower pore sizes observed in PTM with respect to GeoDict results is due to differences in how pores are defined. In GeoDict, an inscribed sphere in the void space represents a pore. In PTM, on the other hand, a pore is defined as a voxel-wide square prism centered at each voxel on the medial surface and perpendicular to the medial surface at that voxel. PTM, therefore, would count more pores of smaller sizes. Despite these algorithmic differences, pore size distributions for all tested materials from low to high porosity yielded curves in the same range and of comparable values, lending confidence in the reconstructive power of the medial surfaces used in PTM.

The degree of fidelity of a medial surface to the structure of the void space can be further tested indirectly, through numerical simulations of multi-phase flow. We designed simple numerical experiments to implement percolation and to solve partial differential equations on medial surface. In particular, permeability calculations demonstrate that PTM accurately captures pore connectivity. Our results presented in Figs. 5678, and 9 demonstrate a good agreement with state-of-the-art methods and confirm that PTM’s medial surfaces (1) have adequate reconstructive power and (2) can be used directly as the solution domain for multi-phase flow simulations.

Numerical experiments presented here were based on several simplifying assumptions and did not utilize all information that can potentially be contained in the medial surface. We searched for the interfaces using the contact angle and the distance to the closest solid wall. This strategy does not include interface stability conditions or trapping assumptions and therefore may result in possible false detection of interfaces and may miss merging and splitting of interfaces. To find the exact location of quasi-static interfaces in capillary dominated flows, detailed geometry of the local cross sections, the 3-D orientation of surrounding solid walls, and the location of other interfaces will be included in the algorithm in the future.

Although the assumption of laminar, fully developed flow between two parallel plates works well for sheet-like, high-aspect-ratio cross sections, it may cause errors when low-aspect-ratio pores are dominant, typically for very low-porosity media. Removing these limitations and relaxing the simplifying assumptions will likely result in higher reconstructive power of the medial surfaces. Again, more geometrical and topological information should be incorporated into the algorithms. Such information is inherent to the medial surface and its attributes such as the distance map. These steps will lead to more realistic assumptions in PTM formulation and help extend its applicability to a wider range of porous media, including non-fibrous materials

5 Conclusions

In this article, we introduced the pore topology method (PTM), a fast, microscale modeling approach based on representing the void space of porous medium by its medial surface and using it as the solution domain for multi-phase flow computations. To validate PTM, we simulated quasi-static drainage and imbibition and computed pore size distributions and absolute permeability for a set of isotropic fibrous materials of varying porosity. Our results are in very good agreement with several state-of-the-art methods such as lattice Boltzmann, full morphology, volume of fluid, and finite volume methods, as well as analytical solutions. Future work will focus on increasing the reconstructive power of the medial surface through incorporating more detailed spatial information about the porous media structure into the medial surface.