Skip to main content
Top
Published in:
Cover of the book

Open Access 2023 | OriginalPaper | Chapter

3. Tortuosity-Porosity Relationships: Review of Empirical Data from Literature

Authors : Lorenz Holzer, Philip Marmet, Mathias Fingerle, Andreas Wiegmann, Matthias Neumann, Volker Schmidt

Published in: Tortuosity and Microstructure Effects in Porous Media

Publisher: Springer International Publishing

Activate our intelligent search to find suitable subject content or patents.

search-config
loading …

Abstract

It is generally assumed that transport resistance in porous media, which can also be expressed as tortuosity, correlates somehow with the pore volume fraction. Hence, mathematical expressions such as the Bruggeman relation (i.e., τ2 = ε−1/2) are often used to describe tortuosity (τ)—porosity (ε) relationships in porous materials. In this chapter, the validity of such mathematical expressions is critically evaluated based on empirical data from literature. More than 2200 datapoints (i.e., τε couples) are collected from 69 studies on porous media transport. When the empirical data is analysed separately for different material types (e.g., for battery electrodes, SOFC electrodes, sandstones, packed spheres etc.), the resulting τ versus ε—plots do not show clear trend lines, that could be expressed with a mathematical expression. Instead, the datapoints for different materials show strongly scattered distributions in rather ill-defined ‘characteristic’ fields. Overall, those characteristic fields are strongly overlapping, which means that the τε characteristics of different materials cannot be separated clearly. When the empirical data is analysed for different tortuosity types, a much more consistent pattern becomes apparent. Hence, the observed τε pattern indicates that the measured tortuosity values strongly depend on the involved type of tortuosity. A relative order of measured tortuosity values then becomes apparent. For example, the values observed for direct geometric and mixed tortuosities are concentrated in a relatively narrow band close to the Bruggeman trend line, with values that are typically < 2. In contrast, indirect tortuosities show higher values, and they scatter over a much larger range. Based on the analysis of empirical data, a detailed pattern with a very consistent relative order among the different tortuosity types can be established. The main conclusion from this chapter is thus that the tortuosity value that is measured for a specific material, is much more dependent on the type of tortuosity than it is dependent on the material and its microstructure. The empirical data also illustrates that tortuosity is not strictly bound to porosity. As the pore volume decreases, the more scattering of tortuosity values can be observed. Consequently, any mathematical expression that aims to provide a generalized description of τε relationships in porous media must be questioned. A short section is thus provided with a discussion of the limitations of such mathematical expressions for τε relationships. This discussion also includes a description of the rare and special cases, for which the use of such mathematical expressions can be justified.
Notes

Supplementary Information

The online version contains supplementary material available at https://​doi.​org/​10.​1007/​978-3-031-30477-4_​3.

3.1 Introduction

The review of concepts and theories (see Chap. 2) reveals a multitude of different approaches, how tortuosity can be defined and measured nowadays. This diversity raises the question, whether the different tortuosity types can be used interchangeably, which is, however, only the case when they reveal identical or very similar results when applied to the same sample and microstructure. If it turns out, that different tortuosity types reveal different results, then the question arises, whether these differences are systematic, predictable, and understandable. A deeper understanding of inherent differences will improve our interpretation of measured tortuosity data and sharpen the scientific discussion of the topic.
To address these issues, we compile and analyse empirical data from the literature in this chapter. Table 3.1 represents a list of data sets with tortuosity-porosity (τ − ε) couples collected in almost 70 different studies. Each line in this table represents a set of data points for a specific class of porous material, which is characterized with a specific tortuosity type. More than 2200 data points were collected semi-automatically from literature (using the webplotdigitizer [70]), in order to investigate and identify specific τε patterns.
τε relationships are collected for the following 7 important classes of porous materials:
(a)
Solid oxide fuel cell (SOFC) electrodes and sintered ceramics,
 
(b)
Gas diffusion layers (GDL) of polymer electrolyte membrane fuel cells (PEM-FC) and other fibrous materials (e.g., paper, particle filters),
 
(c)
Battery electrodes, mostly from Li-ion batteries (LIB),
 
(d)
Geological materials (sandstones, clays, soils),
 
(e)
2D-models of granular materials (packed circles, ellipses, squares, or rectangles)
 
(f)
3D-models of granular materials (packed spheres and ellipsoids, mono-sized and poly-dispersed), and experimental model materials (e.g., glass beads)
 
(g)
3D-models of networked pore structures (from stochastic simulation) and foams.
 
In Sect. 3.2 we analyse this data to see, whether different materials classes reveal specific τε patterns. In our comparison of empirical data, we also consider the different tortuosity types as a criterion for discrimination. In Sect. 3.3, the data from the seven material classes is thus also used to investigate the impact of different tortuosity types on the observed τ–ε patterns. Thereby, we use the classification scheme from Chap. 2 with the main categories: (I) direct geometric τ, (II) mixed τ and (III) indirect physics-based tortuosities (see also Fig. 2.​8). In Sect. 3.4, we consider some specific examples with datasets, where different tortuosity types are applied to the very same microstructure. From these investigations, a consistent pattern can be deduced, which shows how various tortuosity types differ from each other (Sect. 3.5). Finally, in Sect. 3.6, the mathematical descriptions of τε relationships in literature (e.g., Bruggeman, Archie, Maxwell etc.) are reviewed in the light of the findings in previous sections.

3.2 Empirical Data for Different Materials and Microstructure Types

The empirical data in tables 3.1 covers a large variety of materials with a wide range of porosities and with different microstructure characteristics. The data includes, for example, simple structures consisting of mono-sized spheres, but also more complex structures of packed fibers, foams, and composite materials (e.g., SOFC electrodes).
The empirical tortuosity-porosity (τε) data is plotted in Fig. 3.1a–g for each specific material type separately. Overall, these plots show a relatively large scatter, so that no specific τε relationship that could be described with a simple trend line and/or a mathematical expression (such as τ = εx) can be attributed to the specific material types. This finding indicates that even within a single material type the structural variety is large. However, when taking a closer look at Fig. 3.1a–g, for each material type a characteristic field (i.e., a dense cloud of data points) can be observed in the τε plots.
Figure 3.1h illustrates that the characteristic fields of the different material types tend to have a strong overlap. Hence, it seems that τε relationships are not suitable to capture the strong differences between the involved types of materials and structures. Nevertheless, for simple microstructures, such as packed spheres and ellipsoids, the scatter of characteristic fields is relatively small (see Fig. 3.2 e, f). For these rather simple microstructures, the corresponding data points are in a narrow band close to and often slightly above the Bruggeman trend line (i.e., T = τ2 = 1/ε0.5). For all other types of materials and microstructures, the tortuosity-porosity (τε) data shows as strong scattering.
It must be emphasized that the plots in Fig. 3.1 do not distinguish between different types of tortuosities, which may be one reason for the observed scatter of τε datapoints.
The underlying source file for Fig. 3.1, with detailed information to 69 references, can be downloaded from the electronic appendix (Supplementary File 3.1).
Table 3.1
a–g contains list with empirical tortuosity-porosity (τ–ε)-data from literature 2204 tortuosity-porosity (τε) data points are collected from 69 references in literature. The corresponding source file for Table 3.1, which contains additional detailed information, can be downloaded from the electronic appendix (Supplementary File 3.1). It must be emphasized that ‘porosity’ in this table always means the effective porosity, where this information is available. The effective porosity represents the fraction of pore space, which forms a contiguous network, and which excludes trapped pores. In subsequent figures derived from this table the effective porosity is plotted.
Table 3.1a SOFC electrodes and porous ceramics (e.g., separation membrane for SOEC, ceramic catalyst support, or sintered alumina) Legend: (Nr of dataset) Reference [Ref Nr], τ type, (τ methodical details) investigated material
(1)
Grew et al. [1], τindir_diff, (from sim with Laplace solver, Fick’s diff) Ni-YSZ anode
(2)
Grew et al. [2], τindir_diff, (from sim with Laplace solver, Fick’s diff) Ni-YSZ anode
(3)
Iwai et al. [3], τindir_diff, (from sim with random walk, bulk diff) Ni-YSZ anode
(4)
Kishimoto et al. [4], τindir_diff, (from sim with LBM, bulk diff) Ni-YSZ anode
(5)
Cooper et al. [5], τindir_diff, (CFD-FVM, Avizo, random walk, ‘StarCCM+’) LSCF
(6)
Cooper et al. [6], comparison of various τ-types specified below, LSCF cathode
(6a)
τdir_pore_centroid, (image analysis/Avizo Fire, voxel based)
(6b)
τindir _diff_sim. (sim conductivity, ‘StarCCM+’/Laplace s., Fourier’s law, mesh based)
(6c)
τindir _diff_sim. (sim bulk diffusion, ‘AvizoXlab’/Laplace solver, Fick’s law, voxel based)
(6d)
τindir _diff_sim, (simbulk diffusion, ‘TauFactor’/Laplace solver, Fick’s law, voxel based)
(6e)
τindir _diff_sim. (simulation of bulk diffusion, (in-house)/random walk, voxel based)
(7)
Wilson et al. [7], τindir_diff, (sim with Laplace solver, Fick’s diff) Ni-YSZ anode
(8)
Tjaden et al. [8], comparison of tortuosity types, porous YSZ support layer:
(8a)
τdir_FMM, (image analysis/in-house (Matlab)/FIB-SEM)
(8b)
τdir_FMM, (image analysis/in-house (Matlab)/X-ray nano CT)
(8c)
τindir_ therm, (sim. of thermal cond., ‘StarCCM+’/Laplace, Fourier, mesh based/FIB-SEM)
(8d)
τindir_therm, (sim. of thermal cond., ‘StarCCM+’/Laplace, Fourier, mesh based/XnCT)
(8e)
τindir_diff, (diffusion cell experiment/Gas diffusion at 30 and 100 °C/Fick’s law)
(9)
Joos et al. [9], τindir_diff, (from simulation with Laplace solver) LSCF cathode
(10)
Laurencin et al. [10], τindir_diff, (from sim with Laplace solver) Ni-YSZ anode
(11)
Holzer et al. [11], τdir_medial_axis, (Avizo skeletonization/Matlab) Ni-YSZ anode
(12a)
Pecho et al. [12], τdir_geodesic, (Image analysis, in-house) Ni-YSZ anode
(12b)
Pecho et al. [12], supplementary material
(13)
Holzer et al. [13], τdir_geodesic, (Image analysis, in-house) Ni-YSZ anode
(14a)
Wiedenmann et al. [14], τdir_medial_axis, (Avizo/Matlab) sintered Olivine/Wollastonite
(14b)
Wiedenmann et al. [14], τindir_ele_exp, (from EIS experiment) Olivine/Wollastonite
(15)
Zheng et al. [15], τindir_diff, (from sim with Laplace solver) stochastic model for anode
(16)
Lichtner et al. [16], τindir_diff, (from sim with Laplace solver, GeoDict) LSM-YSZ cathode
(17)
Almar et al. [17], τindir_diff, (from sim with Laplace solver) LSCF and BSCF cathodes
(18)
Endler et al. [18], τindir_diff, (from sim with Laplace solver) LSCF cathode
(19)
Holzer et al. [19], 2 tortuosity types, porous Zr-oxide used as Diaphragm in pH sensor:
(19a)
τdir_geodesic, (image analysis/in-house)
(19b)
τindir_ele, (simulation of electrical conduction, GeoDict/Laplace solver/Ohm’s law)
(20)
Haj et al. [20], τdir_geodesic, (Image analysis, in-house) sintered Ni
(21a)
Kishimoto et al. [21], τindir_diff, (Laplace solver, Fick’s diff) pores in CGO anode
(21b)
Kishimoto et al. [21], τindir_ele, (Laplace solver, Ohm’s law) solid phase in CGO anode
(22a)
Shanti et al. [22], τindir_diff, (from sim with Laplace solver, gas diff), sintered alumina
(22b)
Shanti et al. [22], τdir_skeleton, (Amira skeletonization/shortest path analysis), alumina
Table 3.1b
PEM FC (GDL) and fibrous materials. Legend: (Nr of dataset) Reference [Ref Nr], τ type, (τ methodical details) investigated material
(23)
Yu et al. [23], τindir_diff_exp, (from diffusion experiment) Pt electrode
(24)
Sarkar and Bhattacharyya [24], τmixed_hydr_Vav, (from LBM-sim, Navier–Stokes) GDL through-plane dir
(25a)
Garcia-Salaberry et al. [25], τindir_diff, (from LBM-sim) GDL through-plane, var. thicknesses
(25b)
Garcia-Salaberry et al. [25], τindir_diff, (from LBM-sim) GDL in-plane, various thicknesses
(26)
Froning et al. [26], τmixed_hydr_Vav, (from LBM-sim, NS gas flow) GDL in-plane direction
(27a)
Flückiger et al. [27], τindir_diff_exp, (from experiment) GDL dry, no Filler
(27b)
Flückiger et al. [27], τindir_diff_exp, (from experiment) GDL dry, with Filler
(28)
Froning, et al. [28], τmixed_hydr_Vav, (from LBM, NS) PEM GDL real/virtual, w/o Filler
(29)
Holzer et al. [29], 2 tortuosity types, PEM GDL dry IP and TP, compression series:
(29a)
τdir_geodesic, (image analysis, in-house) var. thickness from in-situ μ-CT compression experiment
(29b)
τindir_ele (simulation of electrical conduction, GeoDict/Laplace solver/Ohm’s law)
(30)
Holzer et al. [30], 2 tortuosity types, PEM GDL wet IP and TP, μ-CT imbibition experiment:
(30a)
τdir_geodesic, (image analysis/in-house, from dynamic XCT-imbibition experiment)
(30b)
τindir_ele, (simulation of electrical conduction, GeoDict/Laplace solver/Ohm’s law)
(31)
Huang et al. [31], τdir FMM, (from image analysis) fibrous cloths
Table 3.1c
Battery electrodes. Legend: (Nr of dataset) Reference [Ref Nr], τ type, (τ methodical details) investigated material
(32)
Cooper et al. [32], LI-Battery (LiFePO4):
(32a)
τdir_pore_centroid, (image analysis/Avizo Fire, voxel based)
(32b)
τindir_therm, (sim. thermal cond., ‘StarCCM+’/Laplace solver Fourier’s law, mesh based)
(33)
Tariq et al. [33], τindir_diff, (Star-CCM+, CD adapco, FVM) LIB (graphite)
(34)
Shearing et al. [34], τindir_diff, (Star-CCM + , CD adapco, FVM) LIB (LMO)
(35)
Taiwo et al. [35], τdir_FMM, (image analysis, in-house SW) LIB (LCO, LMO, graphite)
(36)
Almar et al. [36], τindir_diff, (GeoDict, Laplace solver) LIB
(37)
Shearing et al. [37], τindir_diff, (TauFactor, Laplace solver) LIB (graphite)
(38)
Hutzenlaub et al. [38], τindir_diff, (GeoDict, Laplace solver) LIB
(39)
Ebner et al. [39], τindir_diff, (Bruggeman Est., Laplace) LIB (LCO, LMO, graphite)
(40)
Hamann et al. [40], τdir_FMM, (image analysis, Matlab, Fiji) LLCZN garnet electrolyte
(41)
Landesfeind et al. [41], τindir_ele_exp, (EIS experiment) LIB (3–6% binder)
(42a)
Landesfeind et al. [42], τindir_ele_exp, (EIS experiment) LIB (NMC, NCA, graphite)
(42b)
Landesfeind et al. [42], τindir_ele, (Laplace solver) LIB (NMC, NCA, graphite)
(43)
Various authors, cited in Landesfeind et al. [43], τindir_ele, LIB
Table 3.1d
Earth Science and Geo-engineering (sandstones, clays/soils/mortar). Legend: (Nr of dataset) Reference [Ref Nr], τ type, (τ methodical details) investigated material
(44)
Berg [44], τdir_skeleton (flux-weighted, e-Core network model) Fontainebleau sandstone
(45)
Provis et al. [45], τindir_diff, (Random walk sim., in-house) Alkali activated fly ash mortar
(46a)
Klinkenberg [46], cited in Ziarani [47] τindir_ele, (sim. Laplace solver) soil
(46b)
Klinkenberg [46, 47] τindir_ele, (Laplace) unconsolidated sand and glass beads
(46c)
Klinkenberg [46, 47] τindir_ele, (Laplace) sandstones and limestones
(48)
Keller et al. [48], τdir_medial_axis, (Avizo skeletonization/Matlab) Opalinus Clay
(49)
Berg [49], τdir_skeleton (flux-weighted, e-Core network model) Bentheimer sandstone
(50)
Keller et al. [50]τdir_medial_axis, (Avizo skeletonization/Matlab) Opalinus Clay
(51)
Various authors cited in Kristensen et al. [51], τindir_diff_exp, (experiments) soils
(52)
Boving et al. [52] (cited in [51]), τindir_diff_exp, (tracer diffusion experiments) soils
(53)
Nemati [53], τdir_skeleton (flux-weighted, Pore network model) Berea sandstone
Table 3.1e
Granular porous media I: 2D models (circles, ellipsoids, rectangles). Legend: (Nr of dataset) Reference [Ref Nr], τ type, (τ methodical details) investigated material
(54a)
Saomoto and Katagiri [54], τmixed_hydr_Streamline, (Comsol + image analysis) 2D monosized
(54b)
Saomoto and Katagiri [54], τmixed_ele_Streamline, (Comsol + image analysis) 2D monosized
(55)
Saomoto et al. [55], mixed tortuosity types, monosized 2D ellipsoids, aspect ratios 1- 5
(55a)
τmixed_hydr_Streamline, (sim. of flow with Comsol/Image Analysis (IA) of 2D vel. field)
(55b)
τmixed_ele_Streamline, (simulation of el. conduction with Comsol/IA of 2D velocity field)
(55c)
τmixed_hydr_Vav, (simulation of flow with Comsol/IA of 2D velocity field)
(55d)
τmixed_ele_Vav, (simulation of el. conduction with Comsol/IA of 2D velocity field)
(56)
Nabovati et al. [56], τmixed_hydr_Vav, (LBM, Navier Stokes) 2D squares and rectangles
(57)
Ghassemi and Pak [57], τmixed_diff_Vav, (LBM) 2D spheres polydispersed
Table 3.1f
Granular porous media II: 3D models (packed spheres, ellipsoids, sands) and model materials (glass beads, sand). Legend: (Nr of dataset) Reference [Ref Nr], τ type, (τ methodical details) investigated material
(39)
Ebner et al. [39], τindir_diff, (Bruggeman Estimator/Laplace) sphere packing model
(58)
Gommes et al. [58], τdir_geodesic, (Matlab, in-house) Battery Poisson sphere model
(59)
Chung et al. [59], τindir_diff, (Batt3D/Laplace) Battery sphere model with ESyS
(60)
Sobieski [60], τdir_PTM (in-house, virtual 3D with DEM) polydispersed spheres
(61a)
Sheikh and Pak [61], τmixed_diff_Vav (LBM C++, IA) 3D polydispersed spheres
(61b)
Sheikh and Pak [61], τindir_diff (LBM C++) polydispersed spheres with DEM
(62)
Pawlowski et al. [62], τmixed_hydr_SL (Open Foam/VGSMax/ImageVis3D) chromatogr.
(63)
Al-Roush and Madhoun [63], τdir_medial_axis (Tort3D/Matlab) Silica Sand polydisp.
(64a)
Hormann et al. [64], τdir_geodesic (Fiji, in-house), Silica monoliths/chromatogr.
(64b)
Hormann et al. [64], τdir_medial_axis (Fiji, skeletonize3D/analyzeSkeleton) Silica
(46)
Klinkenberg [46], in Ziarani [47, 61], τindir_ele_exp, non-consolidated sand
(65)
Johnson et al. [65], τindir_diff (TauFactor) packed agar/cellulose for chromatogr.
Table 3.1g
Foams and stochastic models of networked structures with continuous phases. Legend: (Nr of dataset) Reference [Ref Nr], τ type, (τ methodical details) investigated material
(66a)
Stenzel et al. [66], τdir_geodesic (in-house) SSGM stochastic spatial graph model
(66b)
Stenzel et al. [66], τindir_ele (GeoDict, Laplace solver) SSGM graph model, contin. phase
(67)
Gaiselmann et al. [67], τdir_medial_axis, (Avizo skeleton/Matlab) SSGM graph model
(68)
Knackstedt et al. [68], τindir_diff (Laplace, bulk diffusion) PU open cell foams
(69)
Vicente et al. [69], τdir_FMM (in-house image analysis) Metal open cell foams
(22a)
Shanti et al. [22], τdir_medial_axis (Image analysis) Alumina packed spheres polydispersed
(22b)
Shanti et al. [22], τindir_diff (Laplace, bulk diffusion) Alumina packed spheres polydisp.

3.3 Empirical Data for Different Tortuosity Types

The same empirical data from Tables 3.1, which in Fig. 3.1 is grouped into different material types, is now replotted in Fig. 3.2, and thereby regrouped into different tortuosity types. The results in Fig. 3.2 illustrate a pronounced pattern, whereby the indirect tortuosities (Fig. 3.2a) scatter over a much larger range than the mixed (Fig. 3.2b) and geometric (Fig. 3.2c, d) tortuosity types.
In Fig. 3.2e geometric and mixed tortuosities are plotted together. But compared to Fig. 3.2a–d, they are plotted with a different scale on the y-axis. The characteristic field in red, representing geodesic and FMM tortuosities, shows the lowest tortuosity values. The measured values for geodesic and FMM tortuosities are often slightly below the Bruggeman trend line and only rarely they take values larger than 2. In contrast, the Bruggeman trend line typically defines the lower bound for mixed tortuosities (green characteristic field) and for those geometric tortuosities, which are derived from a skeleton (medial axis and PTM, blue characteristic field). For porosities below 0.3, the medial axis/PTM tortuosities often also reach values greater than 2. It must be mentioned here that for mixed tortuosities, the empirical data at low porosities is missing, but it is expected that the mixed tortuosity values also show a strong increase with decreasing porosity, in a similar way as it is observed, e.g., for medial axis and PTM tortuosities.
The comparison of empirical results in Figs. 3.1 and 3.2, indicate that the large scatter observed for most materials and microstructure types can be mainly attributed to data points containing indirect tortuosity types. A surprising insight is the fact that the different tortuosity types (Fig. 3.2) give a more characteristic pattern than the different material types (Fig. 3.1). The latter were discussed in the previous section. This indicates that in general the value of tortuosity is more strongly influenced by the type of tortuosity than by the type of material or microstructure. This finding strongly emphasizes the necessity of clearly defining (and choosing carefully) the type of tortuosity, which is used for the characterization of porous media.

3.4 Direct Comparison of Tortuosity Types Based on Selected Data Sets

The data reviewed in the previous sections documents a kind of universal pattern with systematic differences between the various tortuosity types. In this section, we focus on examples that illustrate the difference between specific tortuosity types for a given material and microstructure. This investigation is based on selected data sets from Tables 3.1, where more than one tortuosity type is characterized for the same material or microstructure. For each example, we refer to the corresponding (Nr of data set) in Tables 3.1, which correspond to the [Ref Nr].

3.4.1 Example 1: Indirect Versus Direct Pore Centroid Tortuosity

τindir_diff versus τdir_pore_centroid from SOFC and battery electrodes.
Cooper et al. [5, 6] (dataset Nr 6a-e) used 5 different methods to extract tortuosity from an SOFC cathode material (LSCF). The data set includes four indirect tortuosities that were determined with different simulation approaches (6b: τindir_therm StarCCM+, mesh-based, 6c: τindir_diff, AvizoXlab, voxel-based, 6d: τindir_diff, TauFactor, voxel-based, 6e: τindir_diff, random walk algorithm, voxel-based). As shown in Fig. 3.3, all four approaches give very similar results for the indirect tortuosities (red circles). In comparison, the direct geometric tortuosity (6a: τdir_pore_centroid) for the same sample is significantly lower (red squares).
Cooper et al. [32] (dataset Nr 32a-b, see Fig. 3.3, blue symbols) also presented a comparison of direct geometric tortuosities (32a: τdir_pore_centroid) with indirect diffusional tortuosities (32b: τindir_therm) for a Li-ion battery electrode (LiFePO4). The overall pattern and even the specific tortuosity-porosity values are very similar when comparing the battery electrode (blue) with the SOFC electrode (red). For both materials the values for indirect tortuosity indir_diff/therm) are consistently higher than those for geometric tortuosity (τdir_pore_centroid). From this data, Cooper et al. [5] deduced the following relationship between indirect and direct tortuosities:
$$\tau_{dir\_pore\_centroid} = 0.5\ln \tau_{diff\_indir} + 1$$
(3.1)

3.4.2 Example 2: Indirect Versus Direct Medial Axis Tortuosity

τindir_ele_exp versus τdir_medial_axis from porous ceramic membranes.
Wiedenmann et al. [14] (Dataset Nr. 14 in Table 3.1) presented a comparison of direct medial axis tortuosities (14a: τdir_medial_axis) versus indirect electrical tortuosities (14b: τindir_ele, from EIS experiments) for two different separation membranes in alkaline electrolysis cells, consisting of sintered olivine and wollastonite. As shown in Fig. 3.4 the microstructures vary from fine grained and dense (ε = 0.27) to coarse-grained and open porous (ε = 0.80). Despite the large variation of porosity, the values for medial axis tortuosity (filled squares) are all in a very narrow range (1.62–1.84). The 3D visualizations in Fig. 3.4 [14] document nicely, that the basic geometry of all pore networks remains very similar for all samples, except for the coarseness, which increases with porosity (i.e., scaling). When changing porosity, the obtained values for geometric tortuosity thus remain almost constant (1.73 ± 0.11).
In contrast, the indirect tortuosities (open circles) increase significantly with decreasing porosity from 1.4 to 2.2 for olivine (blue) and from 1.6 to 2.4 for wollastonite (red). Wiedenmann et al. [14] and Holzer et al. [13] documented that the effective properties in these samples scale with constrictivity (i.e., bottlenecks), but not with geometric tortuosity. Therefore, they concluded that the variation of indirect tortuosities in this example rather represents the resistive effects arising from variations in the size of characteristic bottlenecks, rather than effects from path length variations.

3.4.3 Example 3: Indirect Versus Direct Geodesic Tortuosity

τindir_ele/therm_sim versus τindir_diff_exp versus τdir_geodesic/FMM from porous Zr-oxide and PEM GDL.
Figure 3.5 represents a comparison of direct geometric tortuosities (geodesic/FMM) with indirect tortuosities (diffusive/electric/thermal—from simulation and experiment) based on data from Holzer et al. [19], Tjaden et al. [8] and Holzer et al. [29, 30]. As shown in Fig. 3.5, all four studies show that the values for geometric tortuosities (τdir_geodesic and τdir_FMM, closed squares) are systematically lower than the results for indirect tortuosity (open circles and crosses). The samples of the three studies cover a wide range of effective porosities from 4 to 76 vol-%. In addition, the involved microstructures in sintered ceramics (Zr-oxide, YSZ) are very different from those in fibrous PEM GDL. Despite this obvious microstructural variation, all samples show the same consistent pattern. The direct geometric tortuosities (geodesic/FMM) vary hardly and are always below the Bruggeman trend line. Only for very low porosities (ε < 0.2) the geodesic tortuosities start to increase moderately. Apparently, the geodesic path lengths are not very sensitive to variation of the pore volume fraction. In contrast the indirect tortuosities (open symbols and crosses) are always significantly higher than the Bruggeman trend line and they also show much stronger variation. For each series of material/microstructure, a trend of increasing indirect tortuosities is observed when lowering the porosity.

3.4.4 Example 4: Indirect Versus Medial Axis Versus Geodesic Tortuosity

τindir_ele_sim versus τdir_medial_axis versus τdir_geodesic from stochastic 3D structures.
Gaiselmann et al. [67] and Stenzel et al. [66] performed in-depth investigations on the relationship between microstructure characteristics and effective transport properties based on virtual 3D structures generated by stochastic modelling. The so-called stochastic spatial graph model (SSGM) provides 3D structures with a connected transporting phase even at low volume fractions. A large number of microstructures covering a wide range of microstructure characteristics (i.e., volume fractions, phase size distributions, constrictivity and path lengths) were created and analysed. Three different tortuosities (medial axis, geodesic, and indirect tortuosities) can be compared based on datasets Nr. 66 and Nr. 67. It should be noted, that in these studies, electric conduction in the solid phase and associated solid phase microstructure was investigated. However, the effect of microstructure (e.g., tortuosity) on transport in solid phases is basically the same as in the pore phase.
The results in Fig. 3.6 document that the indirect tortuosities (τindir_ele_sim) are consistently higher than the direct geometric tortuosities. Figure 3.6 also reveals slight differences between direct medial axis and direct geodesic tortuosities. The medial axis tortuosity (τdir_medial_axis, red symbols, Nr 67) is slightly higher than the Bruggeman trend line. In contrast, the geodesic tortuosity (τdir_geodesic, black, Nr 66b) is usually below the Bruggeman trend line. However, for small volume fractions of the transporting phase (i.e., ε < 20 vol-%), both tortuosities (i.e., geodesic, and medial axis) show similar values.

3.4.5 Example 5: Direct Medial Axis Versus Direct Geodesic Tortuosity

τdir_medial_axis versus τdir_geodesic from SOFC anodes and silica monoliths.
The relationship between two direct geometric tortuosities (i.e., geodesic, and medial axis tortuosities) is also investigated by Holzer et al. [11] and Pecho et al. [12] for SOFC anodes and by Hormann et al. [64] for silica monoliths. Figure 3.7 clearly shows that geodesic tortuosities (squares) are systematically lower than the medial axis tortuosities (triangles), even though they are measured for the same samples. Furthermore, the Bruggeman trend line can be roughly considered as the boundary between the characteristic fields for these two geometric tortuosity types. These findings are compatible with the results from Gaiselmann et al. [67] and Stenzel et al. [66] in Example 4 (Fig. 3.6).

3.4.6 Example 6: Mixed Streamline Versus Mixed Volume Averaged Tortuosity

τmixed_ele/hydr_streamline versus τmixed_ele/hydr_Vav (mixed tortuosity types) for simulated particle packing (2D and 3D).
This example uses datasets with mixed tortuosity types, including the streamline tortuosity (τmixed_phys_streamline) and the volume-averaged tortuosity (τmixed_phys_Vav) for flow, conduction, and diffusion.
Saomoto et al. [55] simulated hydraulic flow and electrical conduction in simple 2D structures consisting of mono-sized circles and/or ellipsoids. Both mixed tortuosities (i.e., streamline and area-averaged) are extracted from the electric and hydraulic flow fields and their values are plotted in Fig. 3.8. Surprisingly the results for both mixed tortuosity types are almost identical (if considering results for one specific type of transport). For example, the electric streamline tortuosity (red square with crosses) is nearly identical to the electric volume-averaged tortuosity (blue square with crosses) (i.e., τmixed_ele_streamline ≅ τmixed_ele_Vav). The same holds for hydraulic tortuosities (i.e., τmixed_hydr_streamline ≅ τmixed_hydr_Vav). However, a significant difference is observed between electric and hydraulic tortuosities. The characteristic field for mixed electric tortuosities (i.e., τmixed_ele_streamline, τmixed_ele_Vav), which is highlighted in red color, is lower than the characteristic field for mixed hydraulic tortuosities (i.e., τmixed_hydr_streamline, τmixed_hydr_Vav), which is highlighted in blue. The boundary between these characteristic fields is roughly identical with the Bruggeman trend line.
Sheikh and Pak [61] (green crosses) reported diffusive volume-averaged tortuosities (τmixed_diff_ Vav) from 3D poly-dispersed spheres, which are close to the values of mixed electric tortuosities in Saomoto et al. [55], but lower than the mixed hydraulic tortuosities in [55]. This finding is compatible with the general expectation that diffusive and electrical transport properties and associated tortuosities are almost identical to each other and that the hydraulic tortuosities are generally somewhat higher, see e.g., Clennell [71].
Finally, results of volume averaged hydraulic tortuosities (τmixed_hydr_Vav) for gas diffusion layers (GDL) in PEM fuel cells are presented by Froning et al. [28] (crosses highlighted in yellow). Note that the microstructures of fibrous GDL considered by Froning et al. [28] are significantly different from those of poly-dispersed sphere packing in Sheikh and Pack [61]. Nevertheless, the characteristic values of τmixed_hydr_Vav for these two materials (GDL in [28] and packed spheres in [61]) are rather similar. This shows again that the tortuosity values are usually more strongly depending on the type of tortuosity and less strongly on the type of material and/or microstructure.
Overall, the results presented in this example show that the values of mixed tortuosities are relatively low (i.e., close to the Bruggeman trend line). Values larger than 2 are rarely observed and can only be expected for structures with either low porosities (ε < 0.2) or with strong anisotropy effects. Similar as observed previously for the various geometric tortuosities, also the mixed tortuosities show a relatively small scatter, and the values are usually in the range between 1 and 2 (i.e., roughly compatible with Carman’s estimation of √2). This is in remarkable contrast to the indirect tortuosities (not analysed in this example), which scatter over a much larger range and often take values much larger than 2.

3.5 Relative Order of Tortuosity Types

3.5.1 Summary of Empirical Data: Global Pattern of Tortuosity Types

The empirical data from literature reveals a consistent pattern among the tortuosity types, in the sense that certain tortuosity types give consistently higher values than others. This relative order of tortuosity types, which is schematically illustrated in Fig. 3.9, is valid for a wide range of materials with very different microstructures.
Basically, the indirect tortuosities scatter over a much wider range than the direct geometric and the mixed tortuosity types. The direct and mixed types rarely take values greater than 2, whereas for the indirect tortuosities much larger values are measured,—in some cases even higher than 20.
For the geometric tortuosities, two subgroups can be distinguished. For medial axis and path tracking method (PTM) tortuosities, the Bruggeman trend line represents the lower bound. In contrast, geodesic and fast marching method (FMM) tortuosities typically show values that are equal or even lower than the Bruggeman trend line.
The values of mixed tortuosity types (streamline and volume averaged tortuosities) roughly overlap with the values for the direct geometric tortuosities (i.e., usually the mixed tortuosities are also close to the Bruggeman trend line—see e.g., Fu et al. [72]). Example 6 indicates that the mixed streamline tortuosities are identical with the mixed volume-averaged tortuosities, provided that the same transport mechanism is considered (see e.g., Saomoto et al. [55]). However, the mixed electrical, diffusional, or thermal tortuosities are consistently lower than the mixed hydraulic tortuosities. The Bruggeman trend line separates the two characteristic fields for mixed electrical/diffusional and for mixed hydraulic tortuosities (see Fig. 3.8).

3.5.2 Interpretation of Different Tortuosity Categories

3.5.2.1 Direct and Mixed Tortuosities

The direct and mixed tortuosities are based on geometric analyses and therefore they can be considered as true or realistic measures for the path lengths through the porous medium under investigation. Consequently, in order to predict the impact of microstructure on effective transport properties, it is not sufficient to merely consider the geometric or mixed tortuosities, since other morphological effects (e.g., bottlenecks/constrictivity) in addition to path length variation also have an influence on the effective transport properties. Predictions of effective transport properties based on distinct estimations of the path length effect (tortuosity), bottleneck effect (constrictivity) and viscous drag at pore walls (hydraulic radius) are extensively discussed in Chap. 5. Using tortuosity types that give a realistic estimation of the true path length is the key to understanding and differentiating different microstructure effects that limit the transport in porous media.

3.5.2.2 Indirect Tortuosities

The indirect tortuosities are derived from effective (or relative) properties that are known from experiment or simulation. The indirect tortuosities can thus be considered as a measure for the bulk microstructure resistance against transport. The large values and the large variability observed for indirect tortuosities are due to the fact, that they capture all different kinds of microstructure limitations, including resistive effects from narrow bottlenecks. Indirect tortuosities are thus not a realistic measure for the length of transport paths since they tend to overestimate the effect of pure path lengths significantly. Indirect tortuosities can also be interpreted as fudge factor that describes the ratio of relative transport property over porosity (i.e., √(σrel/ε) or √(Drel/ε)). For viscous flow, the indirect hydraulic tortuosity is rarely calculated, since viscous drag expressed by hydraulic radius makes calculation more complicated (see discussion in Chap. 2).

3.6 Tortuosity–Porosity Relationships in Literature

3.6.1 Mathematical Expressions for τε Relationships and Their Limitations

Numerous mathematical expressions describing tortuosity–porosity (τε) relationships can be found in literature. The different τε relationships are reviewed by Shen and Chen [73], Ghanbarian [74], Tjaden et al. [75] and Idris et al. [76]. Table 3.2 represents a selection of mathematical τε relationships from literature.
Table 3.2
Mathematical tortuosity—porosity (τε) relationships from literature
Nr.
τ type
\({\tau }^{2} { }\left( {{or \tau }!} \right) = { f}\left( {\varepsilon } \right)\)
Material type or microstructure type
x = ..
Reference(s)
1
Geom. model
\(\tau_{geometric} = \frac{1}{2}\left[ {1\frac{1}{2}\sqrt {1 - \varepsilon } + \frac{{\sqrt {\left( {1 - \sqrt {1 - \varepsilon } } \right)^{2} + \left( {1 - \varepsilon } \right)/4} }}{{1 - \sqrt {1 - \varepsilon } }}} \right]\)
Geometric 2D model of square particles
 
Yu and Li [81]
3,
2
Geom. model
\(\tau_{geometric} = 1.23\frac{{\left( {1 - \varepsilon } \right)^{4/3} }}{{ x^{2} \varepsilon }}\)
Packed particles → 
Packed spheres → 
0.75
1
Lanfrey et al. [82]
4
Geom. model
\(\tau_{geometric} = \left( {\frac{19}{{18}}} \right)^{{\ln \left( \varepsilon \right)/\ln \left( {8/9} \right)}}\)
Pore fractal model
 
Li and Yu [83]
5,
6,
7
Hydr
\(\tau_{hydr}^{2} = 1 - x\ln \left( \varepsilon \right)\)
Spheres + fibers →
Plates + flakes →
High aspect ratio particles →
0.5
1
3
Pech and Renaud, cited in Comiti and Renaud [84]
8
Hydr
\(\tau_{hydr}^{2} = \varepsilon^{ - x}\)
Binary mixture of spheres
0.4
Mota et al. [85]
9
Hydr
\(\tau_{hydr}^{2} = \frac{\varepsilon }{{1 - \left( {1 - \varepsilon } \right)^{2/3} }}\)
Isotropic granular material
 
Du Plessis et al. [86]
10
11
Hydr
\(\tau_{hydr}^{2} = \sqrt {\frac{2\varepsilon }{{3\left[ {1 - x\left( {1 - \varepsilon } \right)^{2/3} } \right]}} + \frac{1}{3}}\)
Monosized spheres:
Cubic packing →
Tetrahedral packing →
1.209
1.108
Ahmadi [87]
12
Hydr
\(\tau_{mix\_hyd\_Vav} = 1 + 0.8 \left( {1 - \varepsilon } \right)\)
Mono-sized solid rectangles in 2D
 
Koponen et al. [88]
13
Electr
\(\tau_{ele}^{2} = 1 + 0.5\left( {1 - \varepsilon } \right)\)
\(\tau_{ele}^{2} = \frac{{\left( {3 - \varepsilon } \right)}}{2}\)
Dilute suspension of spheres
 
Maxwell, 1873 [89]
14,
15
Electr
\(\tau_{ele}^{2} = \varepsilon^{1 - x}\)
Cementation Exponents:
For rocks →
For sediments →
1.1
2
Archie, 1942 [79]
16
Electr
\({\varvec{\tau}}_{{{\varvec{ele}}}}^{2} = {\varvec{\varepsilon}}^{{1 - {\varvec{x}}}}\), identical with Bruggeman (see Nr 27)
Packing of poly-dispersed spheres
1.5
Archie, 1942 [79]
17,
18,
19,
20,
21,
22
Electr
\(\tau_{ele}^{2} = 1 - x\ln \left( \varepsilon \right)\)
Spherical particles →
Monosized spheres →
Cubic particles →
Cylinders →
Overlapping spheres →
Monosized spheres →
0.41
0.49
0.63
1.00
0.5
0.5
Pech and Renaud, cited in Comiti and Renaud [84]
23
Diff
\(\tau_{diff}^{2} = \varepsilon^{ - x}\)
Mono-sized spheres
0.4
Mota et al. [85], Delgado [90]
24
Diff
\(\tau_{diff}^{2} = \frac{\varepsilon }{{1 - \left( {1 - \varepsilon } \right)^{1/3} }}\)
Catalyst
 
Beeckman [91]
25
Diff
\(\tau_{diff}^{2} = \varepsilon + x\left( {1 - \varepsilon } \right)\)
Sand, silt, sediments
2
Iversen and Jørgensen [92]
26
Non spec
\(\tau^{2} = 2 - \varepsilon\)
   
Petersen [93]
27
Diff
\(\tau_{diff}^{2} = \varepsilon^{{ - {\raise0.7ex\hbox{$1$} \!\mathord{\left/ {\vphantom {1 2}}\right.\kern-0pt} \!\lower0.7ex\hbox{$2$}}}} = \varepsilon^{1 - x}\)
for spheres identical with Archie
(See Nr 16)
Polydisperse granular media:
Spheres →
Cylinders →
1.5
2
Bruggeman 1935 [94]
see also Tjaden et al. [95]
28
Electr
\(\tau_{diff}^{2} = x_{1} \varepsilon^{{1 - x_{2} }}\)
Battery electrodes and separators
x1: 0.1–2.6
x2:1.27–5.2
Thorat [96], cit. in Tjaden [75]
29
Electr
\(\tau_{ele}^{2} = \varepsilon^{ - x}\)
Battery electrode
0.5–2
Ebner [39]
30
Diff
\(\tau_{diff}^{2} = \varepsilon^{ - 1/3}\)
Monodisperse granular media
 
van Brakel and Heertjes [97]
31
Hydr
\(\tau_{diff}^{2} = 1 - \ln \left( \varepsilon \right)\)
Overlapping cylinders
 
Tomadakis and Sotirchos [98]
32
Diff
\(\tau_{diff}^{2} = 1 - \frac{\ln \left( \varepsilon \right)}{2}\)
Overlapping spheres
 
Weissberg [99]
33
Diff
\(\tau_{diff} = 1 - x\ln \left( \varepsilon \right)\)
Clays dry
and hygroscopic
0.357
0.503
Sun et al. [100]
34
Hydr
\(\tau_{hydr}^{2} = \left( {\frac{{\left( {2 - \varepsilon } \right)}}{\varepsilon }} \right)^{2}\)
Cation exchange resin
 
Mackie [101]
35
Diff
\(\tau_{diff} = \left[ {1 - x\left( {1 - \varepsilon } \right)^{ - 1} } \right]\)
Granular media with: spheres →
Cubes →
Large and →
Small parallelepipeds →
0.6
0.73
1.07
1.21
Pisani [102]
36
Hydr
\(\tau_{mixed\_hydr\_Vav}^{2} = x_{1} - x_{2} \ln \left( {\varepsilon^{2} } \right)\)
Monosized spheres
x1: 1.1842
x2: 0.6579
Jin et al. [103]
37
Hydr
\(\tau_{mixed\_hydr\_Vav}^{2} = x_{1} - x_{2} \ln \left( {\varepsilon^{2} } \right)\)
Monosized spheres
x1: 0.9463
x2: 0.7173
Duda [104], cit. in Jin [103]
38
Hydr
\(\tau_{mixed\_hydr\_Vav} = x_{1} \left( {1 - \varepsilon } \right) + x_{2}\)
Monosized spheres
x1: 0.8002
x2: 1.0454
Jin et al. [103]
39
Hydr
\(\tau_{mixed\_hydr\_Vav} = x_{1} \left( {1 - \varepsilon } \right) + x_{2}\)
Monosized spheres
x1: 0.9119
x2: 0.9340
Duda [104], cit. in Jin [103]
40
Hydr
\(\tau_{mixed\_hydr\_Vav} = x_{1} - x_{2} \ln \left( \varepsilon \right)\)
Monosized spheres
x1: 1.1133
x2: 0.4845
Jin et al. [103]
41
Hydr
\(\tau_{mixed\_hydr\_Vav} = x_{1} - x_{2} \ln \left( \varepsilon \right)\)
Monosized spheres
x1: 1.0104
x2: 0.5541
Duda [104], cit. in Jin [103]
42
Hydr
\(\tau_{mixed\_hydr\_Vav} = 1 - x_{1} \ln \left( \varepsilon \right)\)
2D monosized spheres/cubes
0.5/0.541
Saomoto and Katagiri [54]
43
Ele
\(\tau_{mixed\_ele\_Vav} = 1 - x_{1} \ln \left( \varepsilon \right)\)
2D monosized spheres/cubes
0.2/0.19
Saomoto and Katagiri [54]
It must be emphasized that in most references the tortuosity type is not defined clearly. Information on ‘τ type’ is thus often interpreted from the context of a given paper, where possible. ‘Material type’ describes the microstructure for which the τε relationship was derived. Note that depending on the authors, the lefthand side of the equation is sometimes written as tortuosity (τ) and sometimes as tortuosity factor (τ2)
Note that very different mathematical expressions are proposed—usually logarithmic and power-law functions. As shown in Fig. 3.10a, the resulting τε curves diverge greatly from each other. It may be argued that the observed variety results from the fact, that these τε relationships are derived for different material types and different microstructures (e.g., battery electrodes, clays). However, most of these relationships are derived for packed spheres.
The observed chaotic picture of mathematical expressions (Fig. 3.10a) contrasts with the empirical data, which shows a clear pattern when plotted separately for different tortuosity types, as summarized in Fig. 3.9. Moreover, the empirical data typically results in characteristic τε fields (as shown in Fig. 3.2 for the different tortuosity types), but usually it does not result in clearly defined τε curves. This is particularly not the case when plotted for specific material types (see Fig. 3.1) and thereby not distinguishing the involved tortuosity types.
The large scatter for the mathematical expressions in Fig. 3.10a thus illustrates that there is no consensus how tortuosity varies with porosity. In fact, this finding is questioning the underlying assumption that variations of tortuosity and porosity are strictly related, so that the τε relationship could be described with one universal mathematical formula.
The behaviour of microstructure characteristics with varying porosity can be investigated in a more general way, when 3D image analysis is applied to a large number of 3D microstructure models. Results from such studies are shown in Fig. 3.10b–e (taken from Neumann et al. [77] and Stenzel et al. [78]). It must be emphasized that these investigations are using only one specific tortuosity type, which is geodesic tortuosity. These data sets indicate that the microstructure characteristics (ε, β, τdir_geodesic) can vary in a large range within the theoretically possible constellations, except for high porosities, where tortuosity asymptotically tends to 1 (see Fig. 3.10d). For small porosities the variation of geodesic tortuosity is particularly large, but also at very high porosities (ε > 0.8) there is still significant variation in the tortuosity values (Fig. 3.10d). A similar behaviour is expected for all the other direct geometric and mixed tortuosities. These examples thus indicate that there exists no clear correlation between porosity and tortuosity (and associated effective path length, respectively), which could be described with a universal mathematical τε expression. As shown in Fig. Fig. 3.10e, there exists also no strict correlation between relative conductivity and porosity. Hence, it is not justified to expect a strict correlation between indirect tortuosity and porosity, since the first one is derived from relative conductivity.

3.6.2 Mathematical Expressions for τε Relationships and Their Justification

Nevertheless, mathematical expressions for tortuosity–porosity (τε) relationships may have their justification in context with specific cases of controlled microstructure variation. Such a case was described by Archie [79], who presented experimental data for porous sediments saturated with an electrolyte. For the special case, where all samples originate from the same sedimentary unit, the experimental results show a correlation between electric resistance and porosity. Archie’s law (see Eqs. 2.​22 and 2.​23: FR = ε−m = 1/σrel) describes this correlation using a so-called cementation factor (m) as exponent. Wyllie and Rose [80] redefined the relationship between electric resistance (formation factor, FR, respectively) and porosity by introducing the so-called structural factor (Eq. 2.​24: FR = Xele = 1/σrel). Thereby, Xele can be interpreted as being equivalent with Carman’s tortuosity factor (Xele = T = τele2). This leads to Eq. 2.​25 (i.e., τindir_ele2 = ε/σrel), which is widely used to compute indirect tortuosity. Thereby, indirect tortuosity can be considered as some kind of proportionality between porosity and effective properties (e.g., transport resistance or conductivity). For the special case where effective or relative properties (FR, σrel) correlate with porosity, it follows that there must be also a strict correlation with indirect tortuosity, such that Archie’s law can be reformulated as follows:
$$\tau_{indir\_ele}^{2} = \varepsilon^{1 - m}$$
(3.2)
It must be emphasized that Archie’s cementation factor (m > 1) is a fit parameter, which is valid only for a special case of microstructure variation. The cementation factor takes a specific value for a series of rocks, which all have undergone the same diagenetic process (i.e., solidification process transforming loose sediment into a solid rock). The common history of these sediments in the same geo-environment led to a characteristic variation of the microstructure, so that sizes of pores and bottlenecks, pore connectivity and transport path lengths vary in a characteristic way with porosity. Hence, in this case the correlation of microstructure with porosity is controlled by the diagenetic conditions and by the associated cementation process. Due to this controlled correlation, it is possible to find a suitable mathematical expression for the τ − ε relationship of rocks in a single sedimentary unit. Thereby the indirect tortuosity represents the lumped sum of all microstructure effects. With respect to Archie’s law, it must be realized that the fitting of the cementation factor (m) has to be repeated if the rocks originate from a different sedimentary unit, because then these rocks were exposed to different diagenetic conditions, and therefore they are characterized by different τε relationships.
Such special cases are also known from materials engineering when microstructure variations are performed in a controlled way. An example is described by Holzer et al. [19], where sintering temperatures are changed systematically, but all the other parameters (e.g., composition and sinter time) for the fabrication of porous ceramic membranes are kept constant. This results in a systematic variation of microstructure characteristics (τdir_geodesic, β) and effective properties (σeff, σrel) with porosity (ε). Hence, for such controlled microstructure variation a suitable mathematical expression can be found for the observed τε relationship. However, this mathematical expression may no longer be valid, when one of the other fabrication parameters (e.g., composition or sinter time, grinding and particle size distribution of ceramic powders) is changed.
The Bruggeman equation is a special case of Archie’s τε relationship (see Eq. 3.2), where the ‘cementation factor’ (m) is fitted for a special microstructure type. For example, it was found that m = 1.5 for packed spheres (and m = 2 for cylindrical particles), which leads to
$$\tau_{indir\_ele}^{2} = \varepsilon^{ - 0.5} .$$
(3.3)
In battery research, a modified version of the Bruggeman equation, given by
$$\tau_{indir\_ele}^{2} = \gamma \varepsilon^{1 - \alpha }$$
(3.4)
is nowadays often used (see Thorat et al. [96]), where γ is an additional fit parameter. Various authors [96, 105110] presented experimental and numerical fits of Eq. 3.4 (e.g., Nrs 27–30 in Table 3.2) for different battery electrodes, which was critically reviewed by Tjaden et al. [75]. Thereby it is well illustrated that the variations of γ (0.1–2.6) and α (1.27–5.2) are very large and the resulting τε curves differ significantly from each other, depending on the type of battery material. In addition, Tjaden et al. [75] also report examples from literature, which show that for many battery materials the Bruggeman equation and its modifications are simply not applicable. Tjaden et al. [75] thus concluded that tortuosity-porosity relationships such as the Bruggeman equation are only applicable to microstructures ‘which are similar to the microstructure used to derive the respective relationship’ (i.e., for special cases).
These selected examples related to Archie’s and Bruggeman’s equations illustrate that τε relationships should not be mistaken as universal laws. In general, when varying materials microstructures are considered, the different microstructure characteristics (ε, β, τ, rh, rmin, rmax) can vary independently from each other,—at least to some degree. Therefore, empirical data shows significant scatter of τε datapoints for different material types (Fig. 3.1) and also for different tortuosity types (Fig. 3.2). A similar scatter of datapoints is also observed for other morphological characteristics (e.g., constrictivity) and effective properties (see Fig. 3.10b–e). The empirical data illustrate that, in general, there exists no simple and clear correlation between porosity and the other relevant microstructure characteristics (tortuosity, constrictivity, hydraulic radius).
Hence, we conclude that mathematical formulas for tortuosity-porosity relationships are valid only when the following special conditions are fulfilled:
(a)
τε relationship is defined for a specific type of tortuosity.
 
A suitable classification scheme and associated nomenclature are given in Sect. 2.​6 (see Fig. 2.​8). The empirical data show that the scatter of data points is generally much smaller for direct geometric and mixed tortuosity types (Fig. 3.2b–e), compared to the indirect tortuosities (Fig. 3.2a).
(b)
τε relationship is defined for a specific type of material and microstructure.
 
Thereby, the materials under consideration fulfil a systematic microstructure variation (e.g., rocks from the same sedimentary unit, which all had similar conditions during sedimentation and diagenesis). For simple microstructures such as packed spheres it is more probable that microstructure variation (i.e., densification) results in a systematic correlation between tortuosity and porosity (Fig. 3.1e, f), compared to more complex microstructures (e.g., SOFC electrodes, fibrous materials, foams), which tend to show a larger scatter (Figs. 3.1g, h) of τ-ε couples.

3.7 Summary

Nowadays, there are numerous methods available for the characterization of tortuosity in porous media. Despite the progress in characterization, there still exist many controversies, misconceptions, and confusions about tortuosity, which mainly come from the fact that the awareness for systematic differences between tortuosity types is missing. Hence, in many studies and discussions, the different tortuosity types are not clearly distinguished and addressed. As a first step toward solving this problem, we proposed to use a clear terminology. For this purpose, a new classification scheme and a new tortuosity-nomenclature were introduced in Chap. 2 (see e.g., Fig. 2.​8). As a second step, the nature, and the extent of the systematic differences between the various tortuosity types need to be documented and illustrated. In order to investigate these differences, a large collection of empirical data from 69 different references was presented and analysed in this chapter (see Table 3.1).
Based on the collection of empirical data from literature, plots of tortuosity (τ) versus porosity (ε) are presented for different classes of materials and microstructures. For simple microstructures such as monosized sphere packings, the τε plot shows a characteristic field that defines a relatively narrow band close to the Bruggeman trend line. However, for most material classes, which have more complicated, disordered, and stochastic microstructures (such as fuel cell electrodes, foams, rocks, and soils), the corresponding characteristic fields in the τε plots are expanded over much wider domains. In addition, the characteristic fields for these materials classes are strongly overlapping. This overlap is observed even for material classes with significantly different microstructural architectures (e.g., cellular foams, fibrous textiles or sintered ceramics fabricated from powders). Obviously, the different materials and microstructures cannot be distinguished easily based on their τε characteristics.
For the same collection of empirical data, τε characteristics are replotted, but this time separately for the different types of tortuosities. Thereby, systematic differences can be observed among the main tortuosity categories: The indirect tortuosities show relatively high tortuosity values (typically higher than 2 and sometimes up to 20) and their τε characteristics scatter over a wide range. In contrast, τε characteristics of the direct geometric tortuosities and of the mixed tortuosities are typically concentrated in a narrow band close to the Bruggeman trend line (i.e., τ < 1.5 for ε > 0.3).
Systematic differences are also observed between tortuosity types within the same tortuosity category. For example, within the mixed category, the hydraulic tortuosity is consistently higher than the electrical and diffusional tortuosities. Similarly, within the category of direct geometric tortuosities, the medial axis tortuosity is consistently higher than the geodesic type. The τε plots of empirical data thus document a relative order among the different tortuosity types. This consistent pattern is graphically illustrated in Fig. 3.9.
Hence, a consistent τε pattern is observed when the empirical data is plotted for different tortuosity types. However, when the same data is plotted for different materials and microstructures, the corresponding pattern of the characteristic fields is less clear. It follows that the tortuosity value that is measured for a specific material, is much more dependent on the type of tortuosity than it is dependent on the material and its microstructure. This illustrates the need for a clear distinction between the different tortuosity types, including the need for a careful selection of a suitable method and the use of a clear terminology (i.e., nomenclature).
Based on the detailed description of the underlying definitions (see Chap. 2) and based on the documentation of the characteristic τε patterns (present chapter), the main characteristics of the three main tortuosity categories can be summarized as follows:
The direct tortuosities (geodesic, FMM, medial axis, PTM, etc.) as well as the mixed tortuosities (streamline and volume-averaged) are based on geometric analyses and therefore, they provide realistic estimations of the true path lengths. The differences among these tortuosity types are relatively small, and they reflect the existing variations of the underlying geometric and methodological concepts.
The indirect tortuosities are derived from effective (or relative) properties that are known from experiment or simulation. The indirect tortuosities can thus be considered as a measure for the bulk microstructure resistance against transport, which includes also other resistive effects such as the bottleneck effect. Indirect tortuosities should therefore not be misinterpreted as a realistic measure for the length of transport pathways, but they should be rather considered as a bulk factor or fudge factor, which describes the ratio of relative transport property over porosity (e.g., τindir_ele = √(σrel/ε)).
Finally, the empirical data also illustrates that tortuosity is not strictly bound to porosity. As the pore volume decreases, the more scattering of tortuosity values can be observed. Consequently, any mathematical expression that aims to provide a generalized description of τε relationships in porous media must be treated with caution (especially in cases without specification of the corresponding type of tortuosity).
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0 International License (http://​creativecommons.​org/​licenses/​by/​4.​0/​), which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter's Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the chapter's Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder.
Appendix

Electronic supplementary material

Below is the link to the electronic supplementary material.
Literature
1.
go back to reference K. Grew, A.A. Peracchio, J.R. Izzo, W. Chiu, Nondestructive imaging and analysis of transport processes in the solid oxide fuel cell anode, in ECS Transactions, vol. 25 (ECS, 2009), pp. 1861–1870 K. Grew, A.A. Peracchio, J.R. Izzo, W. Chiu, Nondestructive imaging and analysis of transport processes in the solid oxide fuel cell anode, in ECS Transactions, vol. 25 (ECS, 2009), pp. 1861–1870
2.
go back to reference K.N. Grew, A.A. Peracchio, A.S. Joshi, J.R. Izzo Jr., W.K.S. Chiu, Characterization and analysis methods for the examination of the heterogeneous solid oxide fuel cell electrode microstructure. Part 1: volumetric measurements of the heterogeneous structure. J. Power Sources 195, 7930 (2010) K.N. Grew, A.A. Peracchio, A.S. Joshi, J.R. Izzo Jr., W.K.S. Chiu, Characterization and analysis methods for the examination of the heterogeneous solid oxide fuel cell electrode microstructure. Part 1: volumetric measurements of the heterogeneous structure. J. Power Sources 195, 7930 (2010)
3.
go back to reference H. Iwai et al., Quantification of SOFC anode microstructure based on dual beam FIB-SEM technique. J. Power Sources 195, 955 (2010)CrossRef H. Iwai et al., Quantification of SOFC anode microstructure based on dual beam FIB-SEM technique. J. Power Sources 195, 955 (2010)CrossRef
4.
go back to reference M. Kishimoto, H. Iwai, M. Saito, H. Yoshida, Quantitative evaluation of solid oxide fuel cell porous anode microstructure based on focused ion beam and scanning electron microscope technique and prediction of anode overpotentials. J. Power Sources 196, 4555 (2011)CrossRef M. Kishimoto, H. Iwai, M. Saito, H. Yoshida, Quantitative evaluation of solid oxide fuel cell porous anode microstructure based on focused ion beam and scanning electron microscope technique and prediction of anode overpotentials. J. Power Sources 196, 4555 (2011)CrossRef
5.
go back to reference S.J. Cooper, M. Kishimoto, F. Tariq, R.S. Bradley, A.J. Marquis, N.P. Brandon, J.A. Kilner, P.R. Shearing, Microstructural analysis of an LSCF cathode using in situ tomography and simulation. ECS Trans. 57, 2671 (2013)CrossRef S.J. Cooper, M. Kishimoto, F. Tariq, R.S. Bradley, A.J. Marquis, N.P. Brandon, J.A. Kilner, P.R. Shearing, Microstructural analysis of an LSCF cathode using in situ tomography and simulation. ECS Trans. 57, 2671 (2013)CrossRef
6.
go back to reference S.J. Cooper, A. Bertei, P.R. Shearing, J.A. Kilner, N.P. Brandon, TauFactor: an open-source application for calculating tortuosity factors from tomographic data. SoftwareX 5, 203 (2016)CrossRef S.J. Cooper, A. Bertei, P.R. Shearing, J.A. Kilner, N.P. Brandon, TauFactor: an open-source application for calculating tortuosity factors from tomographic data. SoftwareX 5, 203 (2016)CrossRef
7.
go back to reference J.R. Wilson, W. Kobsiriphat, R. Mendoza, H.-Y. Chen, J.M. Hiller, D.J. Miller, K. Thornton, P.W. Voorhees, S.B. Adler, S.A Barnett, Three-dimensional reconstruction of a solid-oxide fuel-cell anode. Nat. Mater. 5, 541 (2006) J.R. Wilson, W. Kobsiriphat, R. Mendoza, H.-Y. Chen, J.M. Hiller, D.J. Miller, K. Thornton, P.W. Voorhees, S.B. Adler, S.A Barnett, Three-dimensional reconstruction of a solid-oxide fuel-cell anode. Nat. Mater. 5, 541 (2006)
8.
go back to reference B. Tjaden, J. Lane, P.J. Withers, R.S. Bradley, D.J.L. Brett, P.R. Shearing, The application of 3D imaging techniques, simulation and diffusion experiments to explore transport properties in porous oxygen transport membrane support materials. Solid State Ion 288, 315 (2016) B. Tjaden, J. Lane, P.J. Withers, R.S. Bradley, D.J.L. Brett, P.R. Shearing, The application of 3D imaging techniques, simulation and diffusion experiments to explore transport properties in porous oxygen transport membrane support materials. Solid State Ion 288, 315 (2016)
9.
go back to reference J. Joos, M. Ender, T. Carraro, A. Weber, E. Ivers-Tiffée, Representative volume element size for accurate solid oxide fuel cell cathode reconstructions from focused ion beam tomography data. Electrochim. Acta 82, 268 (2012)CrossRef J. Joos, M. Ender, T. Carraro, A. Weber, E. Ivers-Tiffée, Representative volume element size for accurate solid oxide fuel cell cathode reconstructions from focused ion beam tomography data. Electrochim. Acta 82, 268 (2012)CrossRef
10.
go back to reference J. Laurencin, R. Quey, G. Delette, H. Suhonen, P. Cloetens, P. Bleuet, Characterisation of solid oxide fuel cell Ni–8YSZ substrate by synchrotron X-ray nano-tomography: from 3D reconstruction to microstructure quantification. J. Power Sources 198, 182 (2012)CrossRef J. Laurencin, R. Quey, G. Delette, H. Suhonen, P. Cloetens, P. Bleuet, Characterisation of solid oxide fuel cell Ni–8YSZ substrate by synchrotron X-ray nano-tomography: from 3D reconstruction to microstructure quantification. J. Power Sources 198, 182 (2012)CrossRef
11.
go back to reference L. Holzer, B. Iwanschitz, Th. Hocker, L. Keller, O. Pecho, G. Sartoris, Ph. Gasser, B. Muench, Redox cycling of Ni–YSZ anodes for solid oxide fuel cells: influence of tortuosity, constriction and percolation factors on the effective transport properties. J. Power Sources 242, 179 (2013) L. Holzer, B. Iwanschitz, Th. Hocker, L. Keller, O. Pecho, G. Sartoris, Ph. Gasser, B. Muench, Redox cycling of Ni–YSZ anodes for solid oxide fuel cells: influence of tortuosity, constriction and percolation factors on the effective transport properties. J. Power Sources 242, 179 (2013)
12.
go back to reference O. Pecho, O. Stenzel, B. Iwanschitz, P. Gasser, M. Neumann, V. Schmidt, M. Prestat, T. Hocker, R. Flatt, L. Holzer, 3D microstructure effects in Ni-YSZ anodes: prediction of effective transport properties and optimization of Redox stability. Materials 8, 5554 (2015)PubMedPubMedCentralCrossRef O. Pecho, O. Stenzel, B. Iwanschitz, P. Gasser, M. Neumann, V. Schmidt, M. Prestat, T. Hocker, R. Flatt, L. Holzer, 3D microstructure effects in Ni-YSZ anodes: prediction of effective transport properties and optimization of Redox stability. Materials 8, 5554 (2015)PubMedPubMedCentralCrossRef
13.
go back to reference L. Holzer, D. Wiedenmann, B. Münch, L. Keller, M. Prestat, P. Gasser, I. Robertson, B. Grobéty, The influence of constrictivity on the effective transport properties of porous layers in electrolysis and fuel cells. J. Mater. Sci. 48, 2934 (2013)CrossRef L. Holzer, D. Wiedenmann, B. Münch, L. Keller, M. Prestat, P. Gasser, I. Robertson, B. Grobéty, The influence of constrictivity on the effective transport properties of porous layers in electrolysis and fuel cells. J. Mater. Sci. 48, 2934 (2013)CrossRef
14.
go back to reference D. Wiedenmann et al., Three-dimensional pore structure and ion conductivity of porous ceramic diaphragms. AIChE J. 59, 1446 (2013)CrossRef D. Wiedenmann et al., Three-dimensional pore structure and ion conductivity of porous ceramic diaphragms. AIChE J. 59, 1446 (2013)CrossRef
15.
go back to reference K. Zheng, Y. Zhang, L. Li, M. Ni, On the tortuosity factor of solid phase in solid oxide fuel cell electrodes. Int. J. Hydrogen Energy 40, 665 (2015)CrossRef K. Zheng, Y. Zhang, L. Li, M. Ni, On the tortuosity factor of solid phase in solid oxide fuel cell electrodes. Int. J. Hydrogen Energy 40, 665 (2015)CrossRef
16.
go back to reference A.Z. Lichtner, D. Jauffrès, D. Roussel, F. Charlot, C.L. Martin, R.K. Bordia, Dispersion, connectivity and tortuosity of hierarchical porosity composite SOFC cathodes prepared by freeze-casting. J. Eur. Ceram. Soc. 35, 585 (2015)CrossRef A.Z. Lichtner, D. Jauffrès, D. Roussel, F. Charlot, C.L. Martin, R.K. Bordia, Dispersion, connectivity and tortuosity of hierarchical porosity composite SOFC cathodes prepared by freeze-casting. J. Eur. Ceram. Soc. 35, 585 (2015)CrossRef
17.
go back to reference L. Almar, J. Szasz, A. Weber, E. Ivers-Tiffée, Oxygen transport kinetics of mixed ionic-electronic conductors by coupling focused ion beam tomography and electrochemical impedance spectroscopy. J. Electrochem. Soc. 164, F289 (2017) L. Almar, J. Szasz, A. Weber, E. Ivers-Tiffée, Oxygen transport kinetics of mixed ionic-electronic conductors by coupling focused ion beam tomography and electrochemical impedance spectroscopy. J. Electrochem. Soc. 164, F289 (2017)
18.
go back to reference C. Endler-Schuck, J. Joos, C. Niedrig, A. Weber, E. Ivers-Tiffée, The chemical oxygen surface exchange and bulk diffusion coefficient determined by impedance spectroscopy of porous La0.58Sr0.4Co0.2Fe0.8O3 - δ (LSCF) cathodes. Solid State Ion 269, 67 (2015) C. Endler-Schuck, J. Joos, C. Niedrig, A. Weber, E. Ivers-Tiffée, The chemical oxygen surface exchange and bulk diffusion coefficient determined by impedance spectroscopy of porous La0.58Sr0.4Co0.2Fe0.8O3 - δ (LSCF) cathodes. Solid State Ion 269, 67 (2015)
19.
go back to reference L. Holzer et al., Fundamental relationships between 3D pore topology, electrolyte conduction and flow properties: towards knowledge-based design of ceramic diaphragms for sensor applications. Mater. Des. 99, 314 (2016)CrossRef L. Holzer et al., Fundamental relationships between 3D pore topology, electrolyte conduction and flow properties: towards knowledge-based design of ceramic diaphragms for sensor applications. Mater. Des. 99, 314 (2016)CrossRef
20.
go back to reference S. Haj Ibrahim, M. Neumann, F. Klingner, V. Schmidt, T. Wejrzanowski, Analysis of the 3D microstructure of tape-cast open-porous materials via a combination of experiments and modeling. Mater. Des. 133, 216 (2017) S. Haj Ibrahim, M. Neumann, F. Klingner, V. Schmidt, T. Wejrzanowski, Analysis of the 3D microstructure of tape-cast open-porous materials via a combination of experiments and modeling. Mater. Des. 133, 216 (2017)
21.
go back to reference M. Kishimoto, M. Lomberg, E. Ruiz-Trejo, N.P. Brandon, Numerical modeling of nickel-infiltrated gadolinium-doped ceria electrodes reconstructed with focused ion beam tomography. Electrochim. Acta 190, 178 (2016)CrossRef M. Kishimoto, M. Lomberg, E. Ruiz-Trejo, N.P. Brandon, Numerical modeling of nickel-infiltrated gadolinium-doped ceria electrodes reconstructed with focused ion beam tomography. Electrochim. Acta 190, 178 (2016)CrossRef
22.
go back to reference N.O. Shanti, V.W.L. Chan, S.R. Stock, F. De Carlo, K. Thornton, K.T. Faber, X-ray micro-computed tomography and tortuosity calculations of percolating pore networks. Acta Mater. 71, 126 (2014)CrossRef N.O. Shanti, V.W.L. Chan, S.R. Stock, F. De Carlo, K. Thornton, K.T. Faber, X-ray micro-computed tomography and tortuosity calculations of percolating pore networks. Acta Mater. 71, 126 (2014)CrossRef
23.
go back to reference Z. Yu, R.N. Carter, J. Zhang, Measurements of pore size distribution, porosity, effective oxygen diffusivity, and tortuosity of PEM fuel cell electrodes. Fuel Cells 12, 557 (2012)CrossRef Z. Yu, R.N. Carter, J. Zhang, Measurements of pore size distribution, porosity, effective oxygen diffusivity, and tortuosity of PEM fuel cell electrodes. Fuel Cells 12, 557 (2012)CrossRef
24.
go back to reference J. Sarkar, S. Bhattacharyya, Application of graphene and graphene-based materials in clean energy-related devices Minghui. Arch. Thermodyn. 33, 23 (2012)CrossRef J. Sarkar, S. Bhattacharyya, Application of graphene and graphene-based materials in clean energy-related devices Minghui. Arch. Thermodyn. 33, 23 (2012)CrossRef
25.
go back to reference P.A. García-Salaberri, J.T. Gostick, G. Hwang, A.Z. Weber, M. Vera, Effective diffusivity in partially-saturated carbon-fiber gas diffusion layers: effect of local saturation and application to macroscopic continuum models. J. Power Sources 296, 440 (2015)CrossRef P.A. García-Salaberri, J.T. Gostick, G. Hwang, A.Z. Weber, M. Vera, Effective diffusivity in partially-saturated carbon-fiber gas diffusion layers: effect of local saturation and application to macroscopic continuum models. J. Power Sources 296, 440 (2015)CrossRef
26.
go back to reference D. Froning, J. Yu, G. Gaiselmann, U. Reimer, I. Manke, V. Schmidt, W. Lehnert, Impact of compression on gas transport in non-woven gas diffusion layers of high temperature polymer electrolyte fuel cells. J. Power Sources 318, 26 (2016)CrossRef D. Froning, J. Yu, G. Gaiselmann, U. Reimer, I. Manke, V. Schmidt, W. Lehnert, Impact of compression on gas transport in non-woven gas diffusion layers of high temperature polymer electrolyte fuel cells. J. Power Sources 318, 26 (2016)CrossRef
27.
go back to reference R. Flückiger, S.A. Freunberger, D. Kramer, A. Wokaun, G.G. Scherer, F.N. Büchi, Anisotropic, effective diffusivity of porous gas diffusion layer materials for PEFC. Electrochim. Acta 54, 551 (2008)CrossRef R. Flückiger, S.A. Freunberger, D. Kramer, A. Wokaun, G.G. Scherer, F.N. Büchi, Anisotropic, effective diffusivity of porous gas diffusion layer materials for PEFC. Electrochim. Acta 54, 551 (2008)CrossRef
28.
go back to reference D. Froning, J. Brinkmann, U. Reimer, V. Schmidt, W. Lehnert, D. Stolten, 3D analysis, modeling and simulation of transport processes in compressed fibrous microstructures, using the lattice Boltzmann method. Electrochim. Acta 110, 325 (2013)CrossRef D. Froning, J. Brinkmann, U. Reimer, V. Schmidt, W. Lehnert, D. Stolten, 3D analysis, modeling and simulation of transport processes in compressed fibrous microstructures, using the lattice Boltzmann method. Electrochim. Acta 110, 325 (2013)CrossRef
29.
go back to reference L. Holzer, O. Pecho, J. Schumacher, Ph. Marmet, O. Stenzel, F.N. Büchi, A. Lamibrac, B. Münch, 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 (2017) L. Holzer, O. Pecho, J. Schumacher, Ph. Marmet, O. Stenzel, F.N. Büchi, A. Lamibrac, B. Münch, 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 (2017)
30.
go back to reference L. Holzer, O. Pecho, J. Schumacher, Ph. Marmet, F. N. Büchi, A. Lamibrac, B. Münch, Microstructure-property relationships in a gas diffusion layer (GDL) for polymer electrolyte fuel cells, part II: pressure-induced water injection and liquid permeability. Electrochim. Acta 241, 414 (2017) L. Holzer, O. Pecho, J. Schumacher, Ph. Marmet, F. N. Büchi, A. Lamibrac, B. Münch, Microstructure-property relationships in a gas diffusion layer (GDL) for polymer electrolyte fuel cells, part II: pressure-induced water injection and liquid permeability. Electrochim. Acta 241, 414 (2017)
31.
go back to reference X. Huang, Q. Wang, W. Zhou, D. Deng, Y. Zhao, D. Wen, J. Li, Morphology and transport properties of fibrous porous media. Powder Technol. 283, 618 (2015)CrossRef X. Huang, Q. Wang, W. Zhou, D. Deng, Y. Zhao, D. Wen, J. Li, Morphology and transport properties of fibrous porous media. Powder Technol. 283, 618 (2015)CrossRef
32.
go back to reference S.J. Cooper et al., Image based modelling of microstructural heterogeneity in LiFePO4 electrodes for Li-ion batteries. J. Power Sources 247, 1033 (2014)CrossRef S.J. Cooper et al., Image based modelling of microstructural heterogeneity in LiFePO4 electrodes for Li-ion batteries. J. Power Sources 247, 1033 (2014)CrossRef
33.
go back to reference F. Tariq, V. Yufit, M. Kishimoto, P.R. Shearing, S. Menkin, D. Golodnitsky, J. Gelb, E. Peled, N.P. Brandon, Three-dimensional high resolution X-ray imaging and quantification of lithium ion battery mesocarbon microbead anodes. J. Power Sources 248, 1014 (2014)CrossRef F. Tariq, V. Yufit, M. Kishimoto, P.R. Shearing, S. Menkin, D. Golodnitsky, J. Gelb, E. Peled, N.P. Brandon, Three-dimensional high resolution X-ray imaging and quantification of lithium ion battery mesocarbon microbead anodes. J. Power Sources 248, 1014 (2014)CrossRef
34.
go back to reference P.R. Shearing, N.P. Brandon, J. Gelb, R. Bradley, P.J. Withers, A.J. Marquis, S. Cooper, S.J. Harris, Multi length scale microstructural investigations of a commercially available Li-ion battery electrode. J. Electrochem. Soc. 159, A1023 (2012)CrossRef P.R. Shearing, N.P. Brandon, J. Gelb, R. Bradley, P.J. Withers, A.J. Marquis, S. Cooper, S.J. Harris, Multi length scale microstructural investigations of a commercially available Li-ion battery electrode. J. Electrochem. Soc. 159, A1023 (2012)CrossRef
35.
go back to reference O.O. Taiwo, D.P. Finegan, D.S. Eastwood, J.L. Fife, L.D. Brown, J.A. Darr, P.D. Lee, D.J.L. Brett, P.R. Shearing, Comparison of three-dimensional analysis and stereological techniques for quantifying lithium-ion battery electrode microstructures. J. Microsc. 263, 280 (2016)PubMedPubMedCentralCrossRef O.O. Taiwo, D.P. Finegan, D.S. Eastwood, J.L. Fife, L.D. Brown, J.A. Darr, P.D. Lee, D.J.L. Brett, P.R. Shearing, Comparison of three-dimensional analysis and stereological techniques for quantifying lithium-ion battery electrode microstructures. J. Microsc. 263, 280 (2016)PubMedPubMedCentralCrossRef
36.
go back to reference L. Almar, J. Joos, A. Weber, E. Ivers-Tiffée, Microstructural feature analysis of commercial Li-ion battery cathodes by focused ion beam tomography. J. Power Sources 427, 1 (2019)CrossRef L. Almar, J. Joos, A. Weber, E. Ivers-Tiffée, Microstructural feature analysis of commercial Li-ion battery cathodes by focused ion beam tomography. J. Power Sources 427, 1 (2019)CrossRef
37.
go back to reference P.R. Shearing, L.E. Howard, P.S. Jørgensen, N.P. Brandon, S.J. Harris, Characterization of the 3-dimensional microstructure of a graphite negative electrode from a Li-ion battery. Electrochem. Commun. 12, 374 (2010)CrossRef P.R. Shearing, L.E. Howard, P.S. Jørgensen, N.P. Brandon, S.J. Harris, Characterization of the 3-dimensional microstructure of a graphite negative electrode from a Li-ion battery. Electrochem. Commun. 12, 374 (2010)CrossRef
38.
go back to reference T. Hutzenlaub, A. Asthana, J. Becker, D.R. Wheeler, R. Zengerle, S. Thiele, FIB/SEM-based calculation of tortuosity in a porous LiCoO2 cathode for a Li-ion battery. Electrochem. Commun. 27, 77 (2013)CrossRef T. Hutzenlaub, A. Asthana, J. Becker, D.R. Wheeler, R. Zengerle, S. Thiele, FIB/SEM-based calculation of tortuosity in a porous LiCoO2 cathode for a Li-ion battery. Electrochem. Commun. 27, 77 (2013)CrossRef
39.
go back to reference M. Ebner, D.W. Chung, R.E. García, V. Wood, Tortuosity anisotropy in lithium-ion battery electrodes. Adv. Energy Mater. 4, 1 (2014) M. Ebner, D.W. Chung, R.E. García, V. Wood, Tortuosity anisotropy in lithium-ion battery electrodes. Adv. Energy Mater. 4, 1 (2014)
40.
go back to reference T. Hamann, L. Zhang, Y. Gong, G. Godbey, J. Gritton, D. McOwen, G. Hitz, E. Wachsman, The effects of constriction factor and geometric tortuosity on Li-ion transport in porous solid-state Li-ion electrolytes. Adv. Funct. Mater. 30 (2020) T. Hamann, L. Zhang, Y. Gong, G. Godbey, J. Gritton, D. McOwen, G. Hitz, E. Wachsman, The effects of constriction factor and geometric tortuosity on Li-ion transport in porous solid-state Li-ion electrolytes. Adv. Funct. Mater. 30 (2020)
41.
go back to reference J. Landesfeind, A. Eldiven, H.A. Gasteiger, Influence of the binder on lithium ion battery electrode tortuosity and performance. J. Electrochem. Soc. 165, A1122 (2018)CrossRef J. Landesfeind, A. Eldiven, H.A. Gasteiger, Influence of the binder on lithium ion battery electrode tortuosity and performance. J. Electrochem. Soc. 165, A1122 (2018)CrossRef
42.
go back to reference J. Landesfeind, M. Ebner, A. Eldiven, V. Wood, H.A. Gasteiger, Tortuosity of battery electrodes: validation of impedance-derived values and critical comparison with 3D tomography. J. Electrochem. Soc. 165, A469 (2018)CrossRef J. Landesfeind, M. Ebner, A. Eldiven, V. Wood, H.A. Gasteiger, Tortuosity of battery electrodes: validation of impedance-derived values and critical comparison with 3D tomography. J. Electrochem. Soc. 165, A469 (2018)CrossRef
43.
go back to reference J. Landesfeind, J. Hattendorff, A. Ehrl, W.A. Wall, H.A. Gasteiger, Tortuosity determination of battery electrodes and separators by impedance spectroscopy. J. Electrochem. Soc. 163, A1373 (2016)CrossRef J. Landesfeind, J. Hattendorff, A. Ehrl, W.A. Wall, H.A. Gasteiger, Tortuosity determination of battery electrodes and separators by impedance spectroscopy. J. Electrochem. Soc. 163, A1373 (2016)CrossRef
44.
go back to reference C.F. Berg, Permeability description by characteristic length, tortuosity, constriction and porosity. Transp. Porous Media 103, 381 (2014)CrossRef C.F. Berg, Permeability description by characteristic length, tortuosity, constriction and porosity. Transp. Porous Media 103, 381 (2014)CrossRef
45.
go back to reference J.L. Provis, R.J. Myers, C.E. White, V. Rose, J.S.J. Van Deventer, X-ray microtomography shows pore structure and tortuosity in alkali-activated binders. Cem. Concr. Res. 42, 855 (2012)CrossRef J.L. Provis, R.J. Myers, C.E. White, V. Rose, J.S.J. Van Deventer, X-ray microtomography shows pore structure and tortuosity in alkali-activated binders. Cem. Concr. Res. 42, 855 (2012)CrossRef
46.
go back to reference L.J. Klinkenberg, Analogy between diffusion and elecrical conductivity in porous rocks. GSA Bull. 62, 559 (1951)CrossRef L.J. Klinkenberg, Analogy between diffusion and elecrical conductivity in porous rocks. GSA Bull. 62, 559 (1951)CrossRef
47.
go back to reference A.S. Ziarani, R. Aguilera, Pore-throat radius and tortuosity estimation from formation resistivity data for tight-gas sandstone reservoirs. J. Appl Geophy 83, 65 (2012)CrossRef A.S. Ziarani, R. Aguilera, Pore-throat radius and tortuosity estimation from formation resistivity data for tight-gas sandstone reservoirs. J. Appl Geophy 83, 65 (2012)CrossRef
48.
go back to reference L.M. Keller, L. Holzer, R. Wepf, P. Gasser, 3D geometry and topology of pore pathways in Opalinus clay: implications for mass transport. Appl Clay Sci 52, 85 (2011)CrossRef L.M. Keller, L. Holzer, R. Wepf, P. Gasser, 3D geometry and topology of pore pathways in Opalinus clay: implications for mass transport. Appl Clay Sci 52, 85 (2011)CrossRef
49.
go back to reference C.F. Berg, Re-examining Archie’s law: conductance description by tortuosity and constriction. Phys. Rev. E 86, 046314 (2012)CrossRef C.F. Berg, Re-examining Archie’s law: conductance description by tortuosity and constriction. Phys. Rev. E 86, 046314 (2012)CrossRef
50.
go back to reference L.M. Keller, L. Holzer, R. Wepf, P. Gasser, B. Münch, P. Marschall, On the application of focused ion beam nanotomography in characterizing the 3D pore space geometry of Opalinus clay. Phys. Chem. Earth 36, 1539 (2011)CrossRef L.M. Keller, L. Holzer, R. Wepf, P. Gasser, B. Münch, P. Marschall, On the application of focused ion beam nanotomography in characterizing the 3D pore space geometry of Opalinus clay. Phys. Chem. Earth 36, 1539 (2011)CrossRef
51.
go back to reference A.H. Kristensen, A. Thorbjørn, M.P. Jensen, M. Pedersen, P. Moldrup, Gas-phase diffusivity and tortuosity of structured soils. J. Contam. Hydrol. 115, 26 (2010)PubMedCrossRef A.H. Kristensen, A. Thorbjørn, M.P. Jensen, M. Pedersen, P. Moldrup, Gas-phase diffusivity and tortuosity of structured soils. J. Contam. Hydrol. 115, 26 (2010)PubMedCrossRef
52.
go back to reference T.B. Boving, P. Grathwohl, Tracer diffusion coefficients in sedimentary rocks: correlation to porosity and hydraulic conductivity. J. Contam. Hydrol. 53, 85 (2001)PubMedCrossRef T.B. Boving, P. Grathwohl, Tracer diffusion coefficients in sedimentary rocks: correlation to porosity and hydraulic conductivity. J. Contam. Hydrol. 53, 85 (2001)PubMedCrossRef
53.
go back to reference R. Nemati, J. Rahbar Shahrouzi, R. Alizadeh, A stochastic approach for predicting tortuosity in porous media via pore network modeling. Comput. Geotech. 120, 103406 (2020) R. Nemati, J. Rahbar Shahrouzi, R. Alizadeh, A stochastic approach for predicting tortuosity in porous media via pore network modeling. Comput. Geotech. 120, 103406 (2020)
54.
go back to reference H. Saomoto, J. Katagiri, Direct comparison of hydraulic tortuosity and electric tortuosity based on finite element analysis. Theor. Appl. Mech. Lett. 5, 177 (2015)CrossRef H. Saomoto, J. Katagiri, Direct comparison of hydraulic tortuosity and electric tortuosity based on finite element analysis. Theor. Appl. Mech. Lett. 5, 177 (2015)CrossRef
55.
go back to reference H. Saomoto, J. Katagiri, Particle shape effects on hydraulic and electric tortuosities: a novel empirical tortuosity model based on van Genuchten-type function. Transp. Porous Media 107, 781 (2015)CrossRef H. Saomoto, J. Katagiri, Particle shape effects on hydraulic and electric tortuosities: a novel empirical tortuosity model based on van Genuchten-type function. Transp. Porous Media 107, 781 (2015)CrossRef
56.
go back to reference A. Nabovati, A.C.M. Sousa, Fluid flow simulation in random porous media at pore level using the lattice Boltzmann method. Sci. Technol. 2, 226 (2007) A. Nabovati, A.C.M. Sousa, Fluid flow simulation in random porous media at pore level using the lattice Boltzmann method. Sci. Technol. 2, 226 (2007)
57.
go back to reference A. Ghassemi, A. Pak, Pore scale study of permeability and tortuosity for flow through particulate media using lattice Boltzmann method. Int. J. Numer. Anal. Methods Geomech. 35, 886 (2011)CrossRef A. Ghassemi, A. Pak, Pore scale study of permeability and tortuosity for flow through particulate media using lattice Boltzmann method. Int. J. Numer. Anal. Methods Geomech. 35, 886 (2011)CrossRef
58.
go back to reference C.J. Gommes, A.-J. Bons, S. Blacher, J.H. Dunsmuir, A.H. Tsou, Practical methods for measuring the tortuosity of porous materials from binary or gray-tone tomographic reconstructions. AIChE J. 55, 2000 (2009)CrossRef C.J. Gommes, A.-J. Bons, S. Blacher, J.H. Dunsmuir, A.H. Tsou, Practical methods for measuring the tortuosity of porous materials from binary or gray-tone tomographic reconstructions. AIChE J. 55, 2000 (2009)CrossRef
59.
go back to reference D.W. Chung, M. Ebner, D.R. Ely, V. Wood, R. Edwin García, Validity of the Bruggeman relation for porous electrodes. Model. Simul. Mat. Sci. Eng. 21, (2013) D.W. Chung, M. Ebner, D.R. Ely, V. Wood, R. Edwin García, Validity of the Bruggeman relation for porous electrodes. Model. Simul. Mat. Sci. Eng. 21, (2013)
60.
go back to reference W. Sobieski, The use of path tracking method for determining the tortuosity field in a porous bed. Granul. Matter 18, 1 (2016)CrossRef W. Sobieski, The use of path tracking method for determining the tortuosity field in a porous bed. Granul. Matter 18, 1 (2016)CrossRef
61.
go back to reference B. Sheikh, A. Pak, Numerical investigation of the effects of porosity and tortuosity on soil permeability using coupled three-dimensional discrete-element method and lattice Boltzmann method. Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 91, 1 (2015)CrossRef B. Sheikh, A. Pak, Numerical investigation of the effects of porosity and tortuosity on soil permeability using coupled three-dimensional discrete-element method and lattice Boltzmann method. Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 91, 1 (2015)CrossRef
62.
go back to reference S. Pawlowski, N. Nayak, M. Meireles, C.A.M. Portugal, S. Velizarov, J.G. Crespo, CFD modelling of flow patterns, tortuosity and residence time distribution in monolithic porous columns reconstructed from X-ray tomography data. Chem. Eng. J. 350, 757 (2018)CrossRef S. Pawlowski, N. Nayak, M. Meireles, C.A.M. Portugal, S. Velizarov, J.G. Crespo, CFD modelling of flow patterns, tortuosity and residence time distribution in monolithic porous columns reconstructed from X-ray tomography data. Chem. Eng. J. 350, 757 (2018)CrossRef
63.
go back to reference R.I. Al-Raoush, I.T. Madhoun, TORT3D: a MATLAB code to compute geometric tortuosity from 3D images of unconsolidated porous media. Powder Technol. 320, 99 (2017)CrossRef R.I. Al-Raoush, I.T. Madhoun, TORT3D: a MATLAB code to compute geometric tortuosity from 3D images of unconsolidated porous media. Powder Technol. 320, 99 (2017)CrossRef
64.
go back to reference K. Hormann, V. Baranau, D. Hlushkou, A. Höltzel, U. Tallarek, Topological analysis of non-granular, disordered porous media: determination of pore connectivity, pore coordination, and geometric tortuosity in physically reconstructed silica monoliths. New J. Chem. 40, 4187 (2016)CrossRef K. Hormann, V. Baranau, D. Hlushkou, A. Höltzel, U. Tallarek, Topological analysis of non-granular, disordered porous media: determination of pore connectivity, pore coordination, and geometric tortuosity in physically reconstructed silica monoliths. New J. Chem. 40, 4187 (2016)CrossRef
65.
go back to reference T.F. Johnson, F. Iacoviello, D.J. Hayden, J.H. Welsh, P.R. Levison, P.R. Shearing, D.G. Bracewell, Packed bed compression visualisation and flow simulation using an erosion-dilation approach. J. Chromatogr. A 1611, 1 (2020)CrossRef T.F. Johnson, F. Iacoviello, D.J. Hayden, J.H. Welsh, P.R. Levison, P.R. Shearing, D.G. Bracewell, Packed bed compression visualisation and flow simulation using an erosion-dilation approach. J. Chromatogr. A 1611, 1 (2020)CrossRef
66.
go back to reference O. Stenzel, O. Pecho, L. Holzer, M. Neumann, V. Schmidt, predicting effective conductivities based on geometric microstructure characteristics. AIChE J. 62, 1834 (2016)CrossRef O. Stenzel, O. Pecho, L. Holzer, M. Neumann, V. Schmidt, predicting effective conductivities based on geometric microstructure characteristics. AIChE J. 62, 1834 (2016)CrossRef
67.
go back to reference G. Gaiselmann, M. Neumann, V. Schmidt, O. Pecho, T. Hocker, L. Holzer, Quantitative relationships between microstructure and effective transport properties based on virtual materials testing. AIChE J. 60, 1983 (2014)CrossRef G. Gaiselmann, M. Neumann, V. Schmidt, O. Pecho, T. Hocker, L. Holzer, Quantitative relationships between microstructure and effective transport properties based on virtual materials testing. AIChE J. 60, 1983 (2014)CrossRef
68.
go back to reference M.A. Knackstedt, C.H. Arns, M. Saadatfar, T.J. Senden, A. Sakellariou, A.P. Sheppard, R.M. Sok, W. Schrof, H. Steininger, Virtual materials design: properties of cellular solids derived from 3D tomographic images. Adv. Eng. Mater. 7, 238 (2005)CrossRef M.A. Knackstedt, C.H. Arns, M. Saadatfar, T.J. Senden, A. Sakellariou, A.P. Sheppard, R.M. Sok, W. Schrof, H. Steininger, Virtual materials design: properties of cellular solids derived from 3D tomographic images. Adv. Eng. Mater. 7, 238 (2005)CrossRef
69.
go back to reference J. Vicente, F. Topin, J.V. Daurelle, Open celled material structural properties measurement: from morphology to transport properties. Mater. Trans. 47, 2195 (2006)CrossRef J. Vicente, F. Topin, J.V. Daurelle, Open celled material structural properties measurement: from morphology to transport properties. Mater. Trans. 47, 2195 (2006)CrossRef
71.
go back to reference M.B. Clennell, Tortuosity: a guide through the maze, in Developments In Petrophysics, ed by M.A. Lovell, P.K. Harvey (Geol. Soc. Spec. Publ. No. 122, 1997), pp. 299–344 M.B. Clennell, Tortuosity: a guide through the maze, in Developments In Petrophysics, ed by M.A. Lovell, P.K. Harvey (Geol. Soc. Spec. Publ. No. 122, 1997), pp. 299–344
72.
go back to reference J. Fu, H.R. Thomas, C. Li, Tortuosity of porous media: image analysis and physical simulation. Earth Sci. Rev. 1 (2020) J. Fu, H.R. Thomas, C. Li, Tortuosity of porous media: image analysis and physical simulation. Earth Sci. Rev. 1 (2020)
73.
go back to reference L. Shen, Z. Chen, Critical review of the impact of tortuosity on diffusion. Chem. Eng. Sci. 62, 3748 (2007)CrossRef L. Shen, Z. Chen, Critical review of the impact of tortuosity on diffusion. Chem. Eng. Sci. 62, 3748 (2007)CrossRef
74.
go back to reference B. Ghanbarian, A.G. Hunt, R.P. Ewing, M. Sahimi, Tortuosity in porous media: a critical review. Soil Sci. Soc. Am. J. 77, 1461 (2013)CrossRef B. Ghanbarian, A.G. Hunt, R.P. Ewing, M. Sahimi, Tortuosity in porous media: a critical review. Soil Sci. Soc. Am. J. 77, 1461 (2013)CrossRef
75.
go back to reference B. Tjaden, D.J.L. Brett, P.R. Shearing, Tortuosity in electrochemical devices: a review of calculation approaches. Int. Mater. Rev. 63, 47 (2018)CrossRef B. Tjaden, D.J.L. Brett, P.R. Shearing, Tortuosity in electrochemical devices: a review of calculation approaches. Int. Mater. Rev. 63, 47 (2018)CrossRef
76.
go back to reference A. Idris, A. Muntean, B. Mesic, A review on predictive tortuosity models for composite films in gas barrier applications. J. Coat Technol. Res. (2022) A. Idris, A. Muntean, B. Mesic, A review on predictive tortuosity models for composite films in gas barrier applications. J. Coat Technol. Res. (2022)
77.
go back to reference M. Neumann, O. Stenzel, F. Willot, L. Holzer, V. Schmidt, Quantifying the influence of microstructure on effective conductivity and permeability: virtual materials testing. Int. J. Solids Struct. 184, 211 (2020)CrossRef M. Neumann, O. Stenzel, F. Willot, L. Holzer, V. Schmidt, Quantifying the influence of microstructure on effective conductivity and permeability: virtual materials testing. Int. J. Solids Struct. 184, 211 (2020)CrossRef
78.
go back to reference O. Stenzel, O. Pecho, L. Holzer, M. Neumann, V. Schmidt, Big Data for microstructure-property relationships: a case study of predicting effective conductivities. AIChE J. 63, 4224 (2017)CrossRef O. Stenzel, O. Pecho, L. Holzer, M. Neumann, V. Schmidt, Big Data for microstructure-property relationships: a case study of predicting effective conductivities. AIChE J. 63, 4224 (2017)CrossRef
79.
go back to reference G.E. Archie, The electrical resistivity log as an aid in determining some reservoir characteristics. Trans. AIME 146 (1942) G.E. Archie, The electrical resistivity log as an aid in determining some reservoir characteristics. Trans. AIME 146 (1942)
80.
go back to reference M.R.J. Wyllie, W.D. Rose, Some theoretical considerations related to the quantitative evaluation of the physical characteristics of reservoir rock from electrical log data. J. Petrol. Technol. 2, 105 (1950)CrossRef M.R.J. Wyllie, W.D. Rose, Some theoretical considerations related to the quantitative evaluation of the physical characteristics of reservoir rock from electrical log data. J. Petrol. Technol. 2, 105 (1950)CrossRef
81.
go back to reference B. Yu, J. Li, Some fractal characters of porous media. Fractals 9, 365 (2001)CrossRef B. Yu, J. Li, Some fractal characters of porous media. Fractals 9, 365 (2001)CrossRef
82.
go back to reference P.-Y. Lanfrey, Z.V. Kuzeljevic, M.P. Dudukovic, Tortuosity model for fixed beds randomly packed with identical particles. Chem. Eng. Sci. 65, 1891 (2010)CrossRef P.-Y. Lanfrey, Z.V. Kuzeljevic, M.P. Dudukovic, Tortuosity model for fixed beds randomly packed with identical particles. Chem. Eng. Sci. 65, 1891 (2010)CrossRef
83.
go back to reference J.-H. Li, B.-M. Yu, Tortuosity of flow paths through a Sierpinski Carpet. Chin. Phys. Lett. 28, 034701 (2011)CrossRef J.-H. Li, B.-M. Yu, Tortuosity of flow paths through a Sierpinski Carpet. Chin. Phys. Lett. 28, 034701 (2011)CrossRef
84.
go back to reference J. Comiti, M. Renaud, A new model for determining mean structure parameters of fixed beds from pressure drop measurements: application to beds packed with parallelepipedal particles. Chem. Eng. Sci. 44, 1539 (1989)CrossRef J. Comiti, M. Renaud, A new model for determining mean structure parameters of fixed beds from pressure drop measurements: application to beds packed with parallelepipedal particles. Chem. Eng. Sci. 44, 1539 (1989)CrossRef
85.
go back to reference M. Mota, J.A. Teixeira, A. Yelshin, Binary spherical particle mixed beds porosity and permeability relationship measurement. Trans. Filtr. Soc. 1, 101 (2001) M. Mota, J.A. Teixeira, A. Yelshin, Binary spherical particle mixed beds porosity and permeability relationship measurement. Trans. Filtr. Soc. 1, 101 (2001)
86.
go back to reference J.P. Du Plessis, J.H. Masliyah, Flow through isotropic granular porous media. Transp. Porous Media 6, 207 (1991)CrossRef J.P. Du Plessis, J.H. Masliyah, Flow through isotropic granular porous media. Transp. Porous Media 6, 207 (1991)CrossRef
87.
go back to reference M.M. Ahmadi, S. Mohammadi, A.N. Hayati, Analytical derivation of tortuosity and permeability of monosized spheres: a volume averaging approach. Phys. Rev. E 83, 026312 (2011)CrossRef M.M. Ahmadi, S. Mohammadi, A.N. Hayati, Analytical derivation of tortuosity and permeability of monosized spheres: a volume averaging approach. Phys. Rev. E 83, 026312 (2011)CrossRef
88.
go back to reference A. Koponen, M. Kataja, J. Timonen, Tortuous flow in porous media. Phys. Rev. E 54, 406 (1996)CrossRef A. Koponen, M. Kataja, J. Timonen, Tortuous flow in porous media. Phys. Rev. E 54, 406 (1996)CrossRef
89.
go back to reference J.C. Maxwell, A Treatise on Electricity and Magnetism, vol. I (Clarendon Press, London, 1873) J.C. Maxwell, A Treatise on Electricity and Magnetism, vol. I (Clarendon Press, London, 1873)
90.
go back to reference J.M.P.Q. Delgado, A simple experimental technique to measure tortuosity in packed beds. Can. J. Chem. Eng. 84, 651 (2008)CrossRef J.M.P.Q. Delgado, A simple experimental technique to measure tortuosity in packed beds. Can. J. Chem. Eng. 84, 651 (2008)CrossRef
91.
go back to reference J.W. Beeckman, Mathematical description of heterogeneous materials. Chem. Eng. Sci. 45, 2603 (1990)CrossRef J.W. Beeckman, Mathematical description of heterogeneous materials. Chem. Eng. Sci. 45, 2603 (1990)CrossRef
92.
go back to reference N. Iversen, B.B. Jørgensen, Diffusion coefficients of sulfate and methane in marine sediments: influence of porosity. Geochim. Cosmochim. Acta 57, 571 (1993)CrossRef N. Iversen, B.B. Jørgensen, Diffusion coefficients of sulfate and methane in marine sediments: influence of porosity. Geochim. Cosmochim. Acta 57, 571 (1993)CrossRef
93.
go back to reference E.E. Petersen, Diffusion in a pore of varying cross section. AIChE J. 4, 343 (1958)CrossRef E.E. Petersen, Diffusion in a pore of varying cross section. AIChE J. 4, 343 (1958)CrossRef
94.
go back to reference D.A.G. Bruggeman, Berechnung Verschiedener Physikalischer Konstanten von Heterogenen Substanzen. I. Dielektrizitätskonstanten Und Leitfähigkeiten Der Mischkörper Aus Isotropen Substanzen. Ann. Phys. 416, 636 (1935) D.A.G. Bruggeman, Berechnung Verschiedener Physikalischer Konstanten von Heterogenen Substanzen. I. Dielektrizitätskonstanten Und Leitfähigkeiten Der Mischkörper Aus Isotropen Substanzen. Ann. Phys. 416, 636 (1935)
95.
go back to reference B. Tjaden, S.J. Cooper, D.J. Brett, D. Kramer, P.R. Shearing, On the origin and application of the Bruggeman correlation for analysing transport phenomena in electrochemical systems. Curr. Opin. Chem. Eng. 12, 44 (2016)CrossRef B. Tjaden, S.J. Cooper, D.J. Brett, D. Kramer, P.R. Shearing, On the origin and application of the Bruggeman correlation for analysing transport phenomena in electrochemical systems. Curr. Opin. Chem. Eng. 12, 44 (2016)CrossRef
96.
go back to reference I.V. Thorat, D.E. Stephenson, N.A. Zacharias, K. Zaghib, J.N. Harb, D.R. Wheeler, Quantifying tortuosity in porous Li-ion battery materials. J. Power Sources 188, 592 (2009)CrossRef I.V. Thorat, D.E. Stephenson, N.A. Zacharias, K. Zaghib, J.N. Harb, D.R. Wheeler, Quantifying tortuosity in porous Li-ion battery materials. J. Power Sources 188, 592 (2009)CrossRef
97.
go back to reference J. Van Brakel, P.M. Heertjes, Analysis of diffusion in macroporous media in terms of a porosity, a tortuosity and a constrictivity factor. Int. J. Heat Mass Transf. 1093 (1974) J. Van Brakel, P.M. Heertjes, Analysis of diffusion in macroporous media in terms of a porosity, a tortuosity and a constrictivity factor. Int. J. Heat Mass Transf. 1093 (1974)
98.
go back to reference M.M. Tomadakis, S.V. Sotirchos, Transport through random arrays of conductive cylinders dispersed in a conductive matrix. J. Chem. Phys. 104, 6893 (1996)CrossRef M.M. Tomadakis, S.V. Sotirchos, Transport through random arrays of conductive cylinders dispersed in a conductive matrix. J. Chem. Phys. 104, 6893 (1996)CrossRef
99.
go back to reference H.L. Weissberg, Effective diffusion coefficient in porous media. J. Appl. Phys. 34, 2636 (1963)CrossRef H.L. Weissberg, Effective diffusion coefficient in porous media. J. Appl. Phys. 34, 2636 (1963)CrossRef
100.
go back to reference Z. Sun, X. Tang, G. Cheng, Numerical simulation for tortuosity of porous media. Microporous Mesoporous Mater. 173, 37 (2013)CrossRef Z. Sun, X. Tang, G. Cheng, Numerical simulation for tortuosity of porous media. Microporous Mesoporous Mater. 173, 37 (2013)CrossRef
101.
go back to reference J.S. Mackie, P. Meares, The diffusion of electrolytes in a cation-exchange resin membrane. Proc. R. Soc. A 232, 498 (1955) J.S. Mackie, P. Meares, The diffusion of electrolytes in a cation-exchange resin membrane. Proc. R. Soc. A 232, 498 (1955)
102.
go back to reference L. Pisani, Simple expression for the tortuosity of porous media. Transp. Porous Media 88, 193 (2011)CrossRef L. Pisani, Simple expression for the tortuosity of porous media. Transp. Porous Media 88, 193 (2011)CrossRef
103.
go back to reference Y. Jin, J.B. Dong, X. Li, Y. Wu, Kinematical measurement of hydraulic tortuosity of fluid flow in porous media. Int. J. Mod. Phys. C 26 (2015) Y. Jin, J.B. Dong, X. Li, Y. Wu, Kinematical measurement of hydraulic tortuosity of fluid flow in porous media. Int. J. Mod. Phys. C 26 (2015)
104.
go back to reference A. Duda, Z. Koza, M. Matyka, Hydraulic tortuosity in arbitrary porous media flow. Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 84, (2011) A. Duda, Z. Koza, M. Matyka, Hydraulic tortuosity in arbitrary porous media flow. Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 84, (2011)
105.
go back to reference M. Ebner, V. Wood, Tool for tortuosity estimation in lithium ion battery porous electrodes. J. Electrochem. Soc. 162, A3064 (2015)CrossRef M. Ebner, V. Wood, Tool for tortuosity estimation in lithium ion battery porous electrodes. J. Electrochem. Soc. 162, A3064 (2015)CrossRef
106.
go back to reference M. Doyle, Comparison of modeling predictions with experimental data from plastic lithium ion cells. J. Electrochem. Soc. 143, 1890 (1996)CrossRef M. Doyle, Comparison of modeling predictions with experimental data from plastic lithium ion cells. J. Electrochem. Soc. 143, 1890 (1996)CrossRef
107.
go back to reference P. Arora, M. Doyle, A.S. Gozdz, R.E. White, J. Newman, Comparison between computer simulations and experimental data for high-rate discharges of plastic lithium-ion batteries. J. Power Sources 88, 219 (2000)CrossRef P. Arora, M. Doyle, A.S. Gozdz, R.E. White, J. Newman, Comparison between computer simulations and experimental data for high-rate discharges of plastic lithium-ion batteries. J. Power Sources 88, 219 (2000)CrossRef
108.
go back to reference D. Kehrwald, P.R. Shearing, N.P. Brandon, P.K. Sinha, S.J. Harris, Local tortuosity inhomogeneities in a lithium battery composite electrode. J. Electrochem. Soc. 158, A1393 (2011)CrossRef D. Kehrwald, P.R. Shearing, N.P. Brandon, P.K. Sinha, S.J. Harris, Local tortuosity inhomogeneities in a lithium battery composite electrode. J. Electrochem. Soc. 158, A1393 (2011)CrossRef
109.
go back to reference J. Cannarella, C.B. Arnold, Ion transport restriction in mechanically strained separator membranes. J. Power Sources 226, 149 (2013)CrossRef J. Cannarella, C.B. Arnold, Ion transport restriction in mechanically strained separator membranes. J. Power Sources 226, 149 (2013)CrossRef
110.
go back to reference N.A. Zacharias, D.R. Nevers, C. Skelton, K. Knackstedt, D.E. Stephenson, D.R. Wheeler, Direct measurements of effective ionic transport in porous Li-ion electrodes. J. Electrochem. Soc. 160, A306 (2013)CrossRef N.A. Zacharias, D.R. Nevers, C. Skelton, K. Knackstedt, D.E. Stephenson, D.R. Wheeler, Direct measurements of effective ionic transport in porous Li-ion electrodes. J. Electrochem. Soc. 160, A306 (2013)CrossRef
Metadata
Title
Tortuosity-Porosity Relationships: Review of Empirical Data from Literature
Authors
Lorenz Holzer
Philip Marmet
Mathias Fingerle
Andreas Wiegmann
Matthias Neumann
Volker Schmidt
Copyright Year
2023
DOI
https://doi.org/10.1007/978-3-031-30477-4_3

Premium Partners