Skip to main content
Log in

Stochastic 3D Modeling of Three-Phase Microstructures for Predicting Transport Properties: A Case Study

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

Abstract

We compare two conceptually different stochastic microstructure models, i.e., a graph-based model and a pluri-Gaussian model, that have been introduced to model the transport properties of three-phase microstructures occurring, e.g., in solid oxide fuel cell electrodes. Besides comparing both models, we present new results regarding the relationship between model parameters and certain microstructure characteristics. In particular, an analytical expression is obtained for the expected length of triple phase boundary per unit volume in the pluri-Gaussian model. As a case study, we consider 3D image data which show a representative cutout of a solid oxide fuel cell anode obtained by FIB-SEM tomography. The two models are fitted to image data and compared in terms of morphological characteristics (like mean geodesic tortuosity and constrictivity) as well as in terms of effective transport properties. The Stokes flow in the pore phase and effective conductivities in the solid phases are computed numerically for realizations of the two models as well as for the 3D image data using Fourier methods. The local and effective physical responses of the model realizations are compared to those obtained from 3D image data. Finally, we assess the accuracy of the two methods to predict permeability as well as electronic and ionic conductivities of the anode.

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
Fig. 5

Similar content being viewed by others

Notes

  1. A different notation has been used in Neumann et al. (2016), where \(\varXi _1\) denoted the pore space and \(\varXi _3\) the nickel phase. We change the notation here to be consistent with the notation of the pluri-Gaussian model considered in Abdallah et al. (2016). In the graph-based model, the index of \(\varXi \) is not meaningful in contrast to the pluri-Gaussian model.

Abbreviations

\(\beta _1, \beta _2, \beta _3\) :

Constrictivities of the three phases

\(\varepsilon _1, \varepsilon _2, \varepsilon _3\) :

Volume fractions of the three phases

\(\widehat{\varepsilon }\) :

Estimator for the volume fraction of a stationary random closed sets

\(\widehat{\varepsilon }^{\star }\) :

Estimator for the volume fraction in the graph-based microstructure model

\(\gamma _1, \gamma _2, \gamma _3\) :

Parameters of the distance measure used for the graph-based microstructure model

\(\varGamma \) :

Pore–solid interface

\(\kappa \,{(}\mathrm {m}^2{)}\) :

Permeability

\(\widehat{\kappa }~{(}\mathrm {m}^2{)}\) :

Geometrical predictor of permeability

\(\lambda _1, \lambda _2, \lambda _3~{(}\mathrm {m}^{-3}{)}\) :

Intensities of the Poisson point processes

\(\mu _f~{(}{{\mathrm {kg}}}\cdot \mathrm {m}^{-1} \cdot \mathrm {s}^{-1}{)}\) :

Viscosity of an incompressible Newtonian fluid

\(\nu _3\) :

3-dimensional Lebesgue measure

\(\varPhi \) :

probability distribution function of the standard normal distribution

\(\phi ~{(}{\mathrm {kg}}\cdot \mathrm {m}^2 \cdot \mathrm {s}^{-3} \cdot \mathrm {A}^{-1}{)}\) :

Electrical potential (or ionic concentration)

\(\rho _Y, \rho _Z\) :

Covariance functions of the Gaussian random fields Y and Z

\(\sigma ~{(}{\mathrm {kg}}^{-1}\cdot \mathrm {m}^{-2} \cdot \mathrm {s}^{3} \cdot \mathrm {A}^{2}{)}\) :

Effective conductivity

\(\sigma _{\mathrm {sol}}~{(}{\mathrm {kg}}^{-1}\cdot \mathrm {m}^{-2} \cdot \mathrm {s}^{3} \cdot \mathrm {A}^{2}{)}\) :

Intrinsic conductivity

\(\tau _1, \tau _2, \tau _3\) :

Mean geodesic tortuosities of the three phases

\(\varTheta \) :

Parameter space

\(\theta _{ij}~{(}\mathrm {m}^{-1}{)}\) :

Parameters for modeling two-point coverage probability functions, \(i,j \in \lbrace 1,2 \rbrace \)

\(\vartheta _0~{(}\mathrm {m}^{-2}{)}, \vartheta _1~{(}\mathrm {m}^{-1}{)}\) :

Intensities of point processes related to the triple phase boundary

\(\varXi _1, \varXi _2, \varXi _3\) :

Random closed sets denoting the three different phases

\(b_1, b_2, b_3\) :

Parameters of the beta-skeletons

\(C_1, C_2, C_3\) :

Two-point coverage probability functions of the three phases

d(xA):

Euclidean distance between a point \(x \in \mathbb {R}^3\) and a set \(A \subset \mathbb {R}^3\)

\(d_\gamma (x,A)\) :

Distance measure with parameter \(\gamma \) between a point \(x \in \mathbb {R}^3\) and a set \(A \subset \mathbb {R}^3\)

\(\mathbf {E}~{(}{\mathrm {kg}}\cdot \mathrm {m} \cdot \mathrm {s}^{-3} \cdot \mathrm {A}^{-1}{)}\) :

Electrical vector field (or opposite gradient of ionic concentration)

\(\mathbf {G}~{(}{\mathrm {kg}}\cdot \mathrm {m}^{-2} \cdot \mathrm {s}^{-2}{)}\) :

Macroscopic pressure gradient

\(\mathcal {G}_1, \mathcal {G}_2, \mathcal {G}_3\) :

Beta-skeletons of the three phases

h :

Function used to estimate the volume fraction in the graph-based model

\(\mathcal {H}_{k}\) :

k-dimensional Hausdorff measure for \(k \in \lbrace 1,2,3 \rbrace \)

\(\mathbf {J}~{(}A{)}\) :

Electrical current (or particle current)

\(L_{\mathrm {TPB}}~{(}{\mathrm {m}}^{-2}{)}\) :

Expected length of the triple phase boundary per unit volume

M :

M-factor, i.e., the ratio of effective and intrinsic conductivity

\(\widehat{M}\) :

Geometrical predictor of the M-factor

o :

Origin in the 3-dimensional Euclidean space

\(p~{(}{\mathrm {kg}}\cdot \mathrm {m}^{-1} \cdot \mathrm {s}^{-2}{)}\) :

Pressure field

\(\mathbb {R}^3\) :

3-dimensional Euclidean space

\(R^2\) :

Coefficient of determination

\(r_{\mathrm {max}}~{(}\mathrm {m}{)}\) :

Median of the volume equivalent particle radius distribution

\(r_{\mathrm {min}}~{(}\mathrm {m}{)}\) :

Median radius of the characteristic bottleneck in a microstructure

\({\mathcal {S}}\) :

Conductive phase

\(S_1, S_2, S_3\) :

Specific surface area of the three phases

\(s_{\mathrm {GBM}}~{(}\mathrm {m}{)}\) :

Smoothing parameter of the graph-based microstructure model

\(s_{\mathrm {PGM}}~{(}\mathrm {m}{)}\) :

Smoothing parameter of the pluri-Gaussian microstructure model

\(u_Y, u_Z\) :

Thresholds defining the excursion sets of the Gaussian random fields Y and Z

\(\mathbf {v}~{(}\mathrm {m} \cdot \mathrm {s}^{-1}{)}\) :

Velocity of an incompressible Newtonian fluid

\(X_1, X_2, X_3\) :

Homogeneous Poisson point processes

YZ :

Gaussian random fields

\(\varDelta \) :

Laplacian operator

\(\nabla \) :

Gradient operator

\(\partial A\) :

Boundary of a set \(A \subset \mathbb {R}^3\)

References

  • Abdallah, B., Willot, F., Jeulin, D.: Stokes flow through a Boolean model of spheres: representative volume element. Transp. Porous Media 109(3), 711–726 (2015)

    Article  Google Scholar 

  • Abdallah, B., Willot, F., Jeulin, D.: Morphological modelling of three-phase microstructures of anode layers using SEM images. J. Microsc. 263(1), 51–63 (2016)

    Article  Google Scholar 

  • Adler, R.J.: The Geometry of Random Fields. Wiley, Chichester (1981)

    Google Scholar 

  • Ballani, F., Kabluchko, Z., Schlather, M.: Random marked sets. Adv. Appl. Probab. 44(3), 603–616 (2012)

    Article  Google Scholar 

  • Bertei, A., Chueh, C.C., Pharoah, J.G., Nicolella, C.: Modified collective rearrangement sphere-assembly algorithm for random packings of nonspherical particles: towards engineering applications. Powder Technol. 253, 311–324 (2014)

    Article  Google Scholar 

  • Brabec, C.J., Heeney, M., McCulloch, I., Nelson, J.: Influence of blend microstructure on bulk heterojunction organic photovoltaic performance. Chem. Soc. Rev. 40(3), 1185–1199 (2011)

    Article  Google Scholar 

  • Cai, Q., Adjiman, C.S., Brandon, N.P.: Modelling the 3D microstructure and performance of solid oxide fuel cell electrodes: computational parameters. Electrochim. Acta 56(16), 5804–5814 (2011)

    Article  Google Scholar 

  • Chiu, S.N., Stoyan, D., Kendall, W.S., Mecke, J.: Stochastic Geometry and Its Applications, 3rd edn. Wiley, Chichester (2013)

    Book  Google Scholar 

  • Hirsch, C., Neuhäuser, D., Schmidt, V.: Connectivity of random geometric graphs related to minimal spanning forests. Adv. Appl. Probab. 45(1), 20–36 (2013)

    Article  Google Scholar 

  • Holzer, L., Pecho, O., Schumacher, J., Marmet, P., Stenzel, O., Büchi, F.N., Lamibrac, A., Münch, B.: Microstructure-property relationships in a gas diffusion layer (GDL) for polymer electrolyte fuel cells, part I: effect of compression and anisotropy of dry GDL. Electrochim. Acta 227, 419–434 (2017)

    Article  Google Scholar 

  • Holzer, L., Wiedenmann, D., Münch, B., Keller, L., Prestat, M., Gasser, P., Robertson, I., Grobéty, B.: The influence of constrictivity on the effective transport properties of porous layers in electrolysis and fuel cells. J. Mater. Sci. 48, 2934–2952 (2013)

    Article  Google Scholar 

  • Kanit, T., Forest, S., Galliet, I., Mounoury, V., Jeulin, D.: Determination of the size of the representative volume element for random composites: statistical and numerical approach. Int. J. Solids Struct. 40, 3647–3679 (2003)

    Article  Google Scholar 

  • Kenney, B., Valdmanis, M., Baker, C., Pharoah, J.G., Karan, K.: Computation of tpb length, surface area and pore size from numerical reconstruction of composite solid oxide fuel cell electrodes. J. Power Sour. 189(2), 1051–1059 (2009)

    Article  Google Scholar 

  • Kirkpatrick, D.R., Radke, J.D.: A framework for computational morphology. In: Toussaint, G.T. (ed.) Computational Geometry, pp. 217–248. North-Holland, Amsterdam (1985)

    Chapter  Google Scholar 

  • Lantuéjoul, C.: Geostatistical Simulation: Models and Algorithms. Springer, Berlin (2013)

    Google Scholar 

  • Matheron, G.: Random Sets and Integral Geometry. Wiley, New York (1975)

    Google Scholar 

  • MATLAB 2015b, The MathWorks. www.matlab.com (2015)

  • Mayer, J., Schmidt, V., Schweiggert, F.: A unified simulation framework for spatial stochastic models. Simulation Modelling Practice and Theory 12(5), 307–326 (2004)

    Article  Google Scholar 

  • Molchanov, I.: Statistics of the Boolean Model for Practitioners and Mathematicians. Wiley, Chichester (1997)

    Google Scholar 

  • Møller, J., Waagepetersen, R.P.: Statistical Inference and Simulation for Spatial Point Processes. Chapman & Hall/CRC, Boca Raton (2004)

    Google Scholar 

  • Moulinec, H., Suquet, P.: A fast numerical method for computing the linear and non linear mechanical properties of the composites. Comptes Rendus de l’Académie des Sciences Série II(318), 1417–1423 (1994)

    Google Scholar 

  • Moussaoui, H., Laurencin, J., Gavet, Y., Delette, G., Hubert, M., Cloetens, P., Le Bihan, T., Debayle, J.: Stochastic geometrical modeling of solid oxide cells electrodes validated on 3D reconstructions. Comput. Mater. Sci. 143, 262–276 (2018)

    Article  Google Scholar 

  • Nelder, J.A., Mead, R.: A simplex method for function minimization. Comput. J. 7, 308–313 (1965)

    Article  Google Scholar 

  • Neumann, M., Hirsch, C., Staněk, J., Beneš, V., Schmidt, V.: Estimation of geodesic tortuosity and constrictivity in stationary random closed sets. Scand. J. Stat. (2019). https://doi.org/10.1111/sjos.12375

  • Neumann, M., Staněk, J., Pecho, O.M., Holzer, L., Beneš, V., Schmidt, V.: Stochastic 3D modeling of complex three-phase microstructures in SOFC-electrodes with completely connected phases. Comput. Mater. Sci. 118, 353–364 (2016)

    Article  Google Scholar 

  • Ohser, J., Schladitz, K.: 3D Images of Materials Structures: Processing and Analysis. Wiley, Weinheim (2009)

    Book  Google Scholar 

  • Pecho, O.M., Stenzel, O., Gasser, P., Neumann, M., Schmidt, V., Hocker, T., Flatt, R.J., Holzer, L.: 3D microstructure effects in Ni-YSZ anodes: prediction of effective transport properties and optimization of redox-stability. Materials 8(9), 5554–5585 (2015)

    Article  Google Scholar 

  • Prakash, B.S., Kumar, S.S., Aruna, S.T.: Properties and development of Ni/YSZ as an anode material in solid oxide fuel cell: a review. Renew. Sustain. Energy Rev. 36, 149–179 (2014)

    Article  Google Scholar 

  • Scholz, C., Wirner, F., Klatt, M.A., Hirneise, D., Schröder-Turk, G.E., Mecke, K., Bechinger, C.: Direct relations between morphology and transport in Boolean models. Phys. Rev. E 92(4), 043023 (2015)

    Article  Google Scholar 

  • Stenzel, O., Neumann, M., Pecho, O.M., Holzer, L., Schmidt, V.: Big data for microstructure-property relationships: a case study of predicting effective conductivities. AIChE J. 63(9), 4224–4232 (2017)

    Article  Google Scholar 

  • Stenzel, O., Pecho, O.M., Holzer, L., Neumann, M., Schmidt, V.: Predicting effective conductivities based on geometric microstructure characteristics. AIChE J. 62, 1834–1843 (2016)

    Article  Google Scholar 

  • Suzuki, T., Hasan, Z., Funahashi, Y., Yamaguchi, T., Fujishiro, Y., Awano, M.: Impact of anode microstructure on solid oxide fuel cells. Science 325(5942), 852–855 (2009)

    Article  Google Scholar 

  • Tjaden, B., Brett, D.J.L., Shearing, P.R.: Tortuosity in electrochemical devices: a review of calculation approaches. Int. Mater. Rev. 63(2), 47–67 (2018)

    Article  Google Scholar 

  • Torquato, S.: Random Heterogeneous Materials: Microstructure and Macroscopic Properties. Springer, New York (2013)

    Google Scholar 

  • Westhoff, D., Van Franeker, J.J., Brereton, T., Kroese, D.P., Janssen, R.A.J., Schmidt, V.: Stochastic modeling and predictive simulations for the microstructure of organic semiconductor films processed with different spin coating velocities. Modell. Simul. Mater. Sci. Eng. 23(4), 045003 (2015)

    Article  Google Scholar 

  • Wiegmann, A.: Computation of the permeability of porous materials from their microstructure by FFF-Stokes (2007). http://kluedo.ub.uni-kl.de/files/1984/bericht129.pdf. Accessed 22 July 2015

  • Willot, F., Abdallah, B., Pellegrini, Y.P.: Fourier-based schemes with modified green operator for computing the electrical response of heterogeneous media with accurate local fields. Int. J. Numer. Methods Eng. 98, 518–533 (2014)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to M. Neumann.

Additional information

Publisher's Note

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

Appendices

Appendix

A Proof of Proposition 1

To prove Proposition 1, we introduce a further stationary random set \(\widetilde{\varXi }_2\) defined by \(\widetilde{\varXi }_2 = \lbrace t \in \mathbb {R}^3: Y(t) \ge \lambda _Y \rbrace .\) Note that the specific surface areas \(S_1\) and \(\widetilde{S}_2\) of \(\varXi _1\) and \(\widetilde{\varXi }_2\), respectively, can be computed by

$$\begin{aligned} S_1 = \frac{2}{\pi } \, e^{-u_{Z}^{2} / 2} \sqrt{-\rho ''_{Z}(0+)} \end{aligned}$$
(20)

and

$$\begin{aligned} \widetilde{S}_2 = \frac{2}{\pi } \, e^{-u_{Y}^{2} / 2} \sqrt{-\rho ''_{Y}(0+)}, \end{aligned}$$
(21)

if Z and Y are mean-square differentiable, see Eq. (6.165) in Chiu et al. (2013) and Remark 7 in Ballani et al. (2012). In that case, \(\rho _Z\) and \(\rho _Y\) are twice differentiable from the right and \(\rho ''_{Z}(0+), \rho ''_{Y}(0+)<0\), see Adler (1981). At first, we derive a formula which allows us to express \( L_{\mathrm {TPB}}\) in terms of \(S_1\) and \(\widetilde{S}_2.\) We show that

$$\begin{aligned} L_{\mathrm {TPB}} =\pi S_1 \widetilde{S}_2 / 4. \end{aligned}$$
(22)

Then, the assertion follows directly, when plugging Eqs. (20) and (21) into Eq. (22). To prove Eq. (22), note that the intersection \(\varXi _0 = \varXi _1 \cap \varXi _2 \cap \varXi _3\) is a motion-invariant random closed set and can be considered as a spatial fiber process in the sense of Sect. 8.4 in Chiu et al. (2013). Then, \( \varXi _0 \cap [0,1]^2 \times \lbrace o \rbrace \) forms a motion-invariant point process with intensity \(\vartheta _0 \ge 0.\) According to Eq. (8.63) in Chiu et al. (2013) we obtain \( L_{\mathrm {TPB}} = 2 \vartheta _0.\) Furthermore, note that the intersection of \(\partial \varXi _1\) with an arbitrary one-dimensional subset of \(\mathbb {R}^3\) forms a motion-invariant point process with intensity \(\vartheta _1 \ge 0\) satisfying \(S_1 = 2 \vartheta _1,\) see Eq. (8.84) in Chiu et al. (2013). In order to compute \(\vartheta _0\), we use the independence of Z and Y. This gives

$$\begin{aligned} \vartheta _0&= \mathbb {E}\mathcal {H}_0(\varXi _0 \cap [0,1]^2 \times \lbrace o \rbrace ) \nonumber \\&= \int \mathbb {E}\left( \mathcal {H}_0(\partial \varXi _1 \cap \partial \widetilde{\varXi }_2 \cap ([0,1]^2 \times \lbrace o \rbrace )) \mid Y\right) \, \mathrm {d}P_Y \nonumber \\&= \vartheta _1 \int \mathcal {H}_1(\partial \widetilde{\varXi }_2 \cap ([0,1]^2 \times \lbrace o \rbrace ) \, \mathrm {d}P_Y \nonumber \\&= \frac{\pi }{8} S_1 \widetilde{S}_2, \end{aligned}$$
(23)

where the last equality is obtained by Eq. (8.83) in Chiu et al. (2013).\(\square \)

Fig. 6
figure 6

Current field \(J_1\) (b, d) in 2D microstructures (a, c, flow occurs in the complement of the white phase) obtained as the union of the YSZ and pore phases: FIB-SEM image (a, b) and PGM model (c, d). The applied electrical field is \(\langle E_1\rangle =1\), \(\langle E_2\rangle =0\) and axis 1 is oriented left to right on the maps. The color maps (b, d) indicate the current flow in the direction of axis 1 (lowest value in black, highest in white, values in-between in red and yellow). The color bar is restricted to current values between 0 and 2. Zones of low current values \(|J_1|<0.026\) are shown in red in maps (a) and (c)

B Visualization of Dead-End Volume in 2D

The dead-end volume turned out to serve as a reasonable interpretation of the difference observed between models (e.g., the PGM model) and the tomographic image data regarding the M-factor of the nickel phase. In order to visualize the dead-end volume, we have computed the current flow in a hypothetical structure. For this purpose, we considered YSZ and the pore space in a random 2D slice of tomographic image data and of realizations of the PGM model as conducting phases, while nickel is insulating. An electrical field \(\langle E_1\rangle =1\) is applied for these 2D structures. This approach is useful for several reasons. First, it is easier to compare the field patterns between different structures in a purely 2D problem. This is however possible as long as discrepancies between the M-factors of model realizations and FIB-SEM images observed in 3D are also reflected in the results of 2D computations. Second, the nickel phase does not percolate in a 2D cut; hence, we consider the flow in the complementary phase of the nickel (YSZ and pores). Indeed if the nickel phase of the models was representative of that observed in the FIB-SEM image, this would be the case also for its complementary. We observe a very significant difference between the two effective conductivities in 2D, equal to \(\sigma _{{\mathrm {YSZ,pores}}}=0.21\) for the realization of the PGM model and \(\sigma _{{\mathrm {YSZ,pores}}}=0.31\) for the FIB-SEM image. This suggests that the reason for the discrepancy is also present in the simpler 2D problem. In Fig. 6, regions of low current values \(|J_1|<0.026\) are highlighted in red. These regions are considered as dead-end volume of the union of pores and YSZ. We observe that there is a significantly larger amount of such regions in realization of the PGM model than in tomographic image data. This can be related to much larger clusters in the model realization for the union of YSZ and of the pores than in tomographic image data (Fig. 6a and b), which act as barriers. The presence of such barriers is consistent with a higher value of effective conductivity \(\sigma _{{\mathrm {YSZ,pores}}}=0.31\) of tomographic image compared to the model realization (\(\sigma _{{\mathrm {YSZ,pores}}}=0.21\)).

C Description of Computer Implementation of the Calculations

In the following, we provide some technical details of the implementations used to simulate the virtual microstructures and their effective transport properties. The simulation of virtual microstructures by the GBM is implemented using Java in the framework of the software library Geostoch (Mayer et al. 2004). Drawing one model realization with the parameters given in Table 1 takes about 25 min on a desktop computer. The code for generating virtual microstructures with the PGM is written in MATLAB (2015). One model realization with the parameters given in Table 2 takes about 5 min on a desktop computer. For the simulation of effective conductivity and permeability as described in Sect. 4.1, a Fortran code parallelized on a 24-cores machine is used. Calculations take about 2 h and 40 min for effective conductivity and permeability, respectively.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Neumann, M., Abdallah, B., Holzer, L. et al. Stochastic 3D Modeling of Three-Phase Microstructures for Predicting Transport Properties: A Case Study. Transp Porous Med 128, 179–200 (2019). https://doi.org/10.1007/s11242-019-01240-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11242-019-01240-y

Keywords

Navigation