2,619
Views
52
CrossRef citations to date
0
Altmetric
Critical Assessment

Critical Assessment 18: elastic and thermal properties of porous materials – rigorous bounds and cross-property relations

&

Abstract

A critical assessment of model relations describing the porosity dependence of elastic properties (Young's modulus) and thermal properties (thermal conductivity) is given. It is shown that there are essentially five types of admissible predictive model relations for the relative Young's modulus and thermal conductivity of isotropic porous materials. The cross-property relations resulting from the complete analogy between the model relations for the elastic moduli and thermal conductivity of isotropic porous materials are reviewed and compared. Finally, it is shown that the fact that relative Young's moduli are not equal to relative thermal conductivities except for materials with translational symmetry, i.e. the mere existence and necessity of non-trivial cross-property relations, proves so-called minimum solid area models to be wrong.

Introduction

Many materials of engineering interest are porous, and in many cases porosity is intentionally introduced in order to meet certain structural or functional requirements, e.g. light weight, low heat capacity, increased flexibility, good insulating performance (thermal, electrical or acoustic), high fluid permeability (in filters), high surface area (in catalysts and catalyst supports), etc.Citation1CitationCitationCitation4 Porous materials are heterogeneous materials for which the salient feature of the microstructure is the existence of a pore space. Thus, porous materials may be considered as a special case of heterogeneous materials that may be modelled as two-phase systems, consisting of a solid ‘phase’ (which may, of course, be itself a heterogeneous material, e.g. a phase mixture or composite consisting of several solid phases or a single-phase polycrystalline material) and a void phase (either vacuum voids or voids filled with fluids, usually a gas). Properties sensu stricto are coefficients in linear constitutive equations, e.g. Hooke's law or Fourier's law. Among these are the elastic constants (tensile, shear and bulk modulus as well as Poisson ratio), thermal conductivity, specific heat and thermal expansion coefficients.Citation5 The properties of heterogeneous materials, including porous materials, are usually called ‘effective properties’.Citation6 The theory to calculate effective properties from the constituent phase properties and the microstructure is often called ‘micromechanics’Citation7 or ‘theory of composites’,Citation8 although ‘theory of heterogeneous materials’, including ‘theory of porous media’ as a special field, is probably the most appropriate term.

Since the aforementioned elastic and thermal properties of a porous material of a given composition are primarily determined by the porosity (i.e. the volume fraction of pores), and only to a minor degree by ot1her microstructural features (e.g. pore size, shape and connectivity), and since porosity is usually easy to measure, it is reasonable to consider property–porosity dependences in the first place and then possibly take into account other microstructural details if desired, according to their relative importance. Now it can be shown that some of these effective properties of porous materials are indeed independent of the porosity, e.g. the specific heat per unit mass and the thermal expansion coefficients.Citation9 Among the more challenging effective properties are the elastic constants and thermal conductivity, which are strongly dependent on the porosity and other microstructural features of the porous materials. Therefore it is not surprising that the literature is full of traditional model relations that have been proposed to describe the porosity dependence of these properties and are summarised both in monographs and journal papers.Citation10CitationCitationCitationCitationCitation15 However, many of these relations are redundant (since they have been ‘rediscovered’ several times and denoted by different names), and not all of these relations are admissible from the physical point of view.

In particular, in the field of thermal conductivity, it seems to be still widely believed that the parallel and series models, the Maxwell-type models and the self-consistent model (sometimes called Bruggeman model or, misleadingly, ‘effective medium theory’ model/EMT model) are the only parameter-free predictive model relations for conductivity prediction,Citation14 although it is well known that Coble–Kingery-type relations, power–law relations and exponential relations have a quite similar theoretical status,Citation16 the parallel, series and Maxwell-type models just correspond to the micromechanical bounds and in the case of vacuum pores the self-consistent model degenerates to the linear approximation and predicts a spurious percolation threshold,Citation6 i.e. describes non-linear behaviour only when the thermal conductivity of the pore phase is sufficiently large, but not as an intrinsic consequence of the microstructure (as it should be).

On the other hand, the field of elastic moduli description is dominated by a completely unacceptable relation, the simple exponential, apart from a large amount of other traditional relations that are cited under different names. This has led to the unpleasant situation that traditional modulus–porosity relations are cited over and over again, even in quite recent papers,Citation15 although it is well known for quite some time that many of these relations are inadmissible for predictive purposes, and some of them are misleading or inconvenient even when used for fitting. A look into the current literature shows that this practice, in which trivial mixture rules are listed side by side with modified relations and special results that hold only for very special geometries, and rigorous results are listed side by side with empirical fit relations, all this using inconsequent terminology, is bound to confuse not only the beginner. Moreover, it seems that the complete formal analogy of relations for describing elastic moduli and thermal conductivity, described in one of our earlier papers,Citation16 is not yet generally known outside the ceramics community.

The present critical assessment is an attempt to clarify this situation by recalling the rational criteria based on which the admissibility of any relation can be tested and by presenting the most important non-empirical model relations that are at least admissible for predictive purposes and have turned out to be successful for porous real-world materials. Among other things, it is emphasised that specific differences between the relations usable for predicting the porosity dependence of thermal conductivity and Young's modulus imply the existence of non-trivial cross-property relations. Since all of these non-trivial cross-property relations, which are reviewed in this paper for the first time, directly contradict so-called minimum-solid area models,Citation10,Citation11,Citation17Citation25 it is necessary to abandon the latter.

Rigorous bounds for elastic moduli and thermal conductivity

The elastic moduli and thermal conductivity of multiphase materials are strongly dependent on the microstructure, in particular on the volume fractions of the constituent phases. These properties are generally restricted from above and below by upper and lower bounds. The simplest of these bounds are the so-called one-point bounds, corresponding to the (volume-weighted) arithmetic and harmonic means, respectively, called Wiener boundsCitation26 in conductivity context (and several other context such as electrical conductivity, dielectric permittivity and magnetic permeability) and Voigt–Reuss boundsCitation27,Citation28 or Paul boundsFootnote* Citation29 in elasticity context. In the sequel these bounds are denoted as Wiener–Paul bounds (WP bounds).

In the special case of isotropic microstructures narrower bounds, so-called two-point bounds, are available, or more precisely, reduce to bounds requiring only volume fraction (one-point) information.Citation6 These are called Hashin–Shtrikman bounds (HS bounds),Citation6CitationCitation8 both for the thermal conductivity (and several other second-order tensor properties such as electrical conductivity, dielectric permittivity and magnetic permeabilityCitation30) and the elastic moduliCitation31,Citation32 (properties based on fourth-order tensors defined for the special case of isotropic materials). When the phase contrast between the properties of the constituent phases is not too high, these bounds are fully sufficient for estimating the effective properties (e.g. elastic moduli or thermal conductivity) of multiphase materials.Footnote Citation6CitationCitationCitation9,Citation32

When the phase contrast is higher, however, the bounds become wider and thus the estimates of the effective properties become very imprecise. In particular, it can easily be shownCitation32Citation,Citation34 that when one of the phases has a zero (or negligible) property value, as in the case of elastic moduli and thermal conductivity of many common porous materials, the lower bounds degenerate to zero (or negligibly small values) in the whole range of pore volume fractions (porosities) from zero to one (and thus become useless), while the upper one-point bounds (WP bounds) reduce toCitation16 (1)

where denotes the porosity (volume fraction of pores) and we have defined the relative property as the (dimensionless) ratio of the effective property of the porous material as a whole and the property of the dense, pore-free, solid (which can be a single solid phase, usually polycrystalline, or a solid phase mixture). Note that the solid portion of the porous material can form a matrix (in the case of a matrix-inclusion or closed-cell microstructure containing closed isolated pores) or a skeleton (in the case of a bicontinuous or open-cell microstructure). Of course, the solid property itself may have to be calculated by an appropriate averaging of tensor components of the individual crystallitesCitation6CitationCitationCitation9 and (in the case of a solid phase mixture) by applying the aforementioned rigorous bounds. Note that here and in the sequel denotes any of the elastic moduli (tensile or Young's modulus , shear modulus and bulk modulus ), or thermal conductivity .

While the upper one-point bound (upper WP bound), equation (1), holds generally for arbitrary microstructures, including anisotropic ones, in the special case of isotropic porous materials the upper two-point bounds (upper HS bounds) are (2) for the tensile modulus (Young's modulus), at least to an excellent approximation,Citation32 and (3) for the thermal conductivity (exactly).Citation9,Citation16,Citation34

It has to be emphasised that, when the phase contrast attains several orders of magnitude or even approaches infinity (as in the case of porous materials), the simple arithmetic mean of the upper and lower bounds (either WP or HS bounds) cannot be expected to provide a realistic estimate of the effective property anymore, because it leads to a spurious value of 0.5 for the relative property in the limit . Although weighted means can be defined that avoid this difficulty,Citation35CitationCitation37 it is clear that these empirical means do not contribute anything essential to a physical understanding of property–porosity dependences. Therefore, for porous materials (in contrast to composites with a sufficiently low phase property contrast) model-based predictive relations are needed in addition to the aforementioned rigorous bounds (which are principally not model-based, i.e. have not been derived for specific microstructures, but have been derived on the basis of very universal variational principlesCitation30,Citation31).

Model relations for elastic moduli and thermal conductivity

Once the solid property is reliably known (e.g. for a single polycrystalline solid phase from orientational averaging of the monocrystal tensor components,Citation6CitationCitationCitation9,Citation38CitationCitationCitation41 for multiphase materials by calculating arithmetic or other means of upper and lower one- or two-point boundsCitation32,Citation34,Citation36,Citation37), several model relations are available that can be invoked for predicting the porosity dependence of the effective property of the corresponding porous material. In fact, the parallel model, (4)

which corresponds formally to the upper one-point bound (upper WP bound), equation (1), and which exactly predicts the axial property component of porous materials with translational symmetry (translational invariance), is such a model. The remarkable feature of this relation is the fact that it has the same form for all elastic moduli and thermal conductivity and is independent of the cross-section of the pore channels. However, for the same reasons it is clear that equation (4) can never be a good prediction for other directions than the axial one or for materials without translational symmetry, e.g. isotropic porous materials.

All model relations for isotropic materials must include a specific coefficient that depends on the property in question and takes into account pore shape and, in the case of elastic moduli, the influence of the Poisson ratio (at least that of the solid phase, in the sequel called ‘solid Poisson ratio’ ). In contrast to widespread belief, however, the value of this coefficient is not freely assignable, but must be coupled to the exact solution of the corresponding single-inclusion problem.Citation6CitationCitation8,Citation42CitationCitationCitationCitationCitation47 The ensuing linear relation (5)

with the coefficient dependent on the property, pore shape and solid Poisson ratio (in the case of elastic moduli) may be considered as the exact solution of the single-inclusion problem, i.e. a single pore in an infinitely extended solid matrix. With respect to the properties of real-world materials this exact solution is called the ‘non-interaction approximation’,Citation7 since the pore is assumed to be isolated and infinitely distant from other pores. Although this linear model relation can be expected to hold only for materials with very low porosity values, and predicts a spurious percolation threshold for higher porosity,Citation6 it serves as a benchmark relation to which all other (non-linear) admissible relations should reduce in the case of low porosity. We note in passing that so-called self-consistent approximations,Citation6CitationCitation8 in conductivity context attributed to Bruggeman,Citation48 in elasticity context due to HillCitation49 and Budiansky,Citation50 reduce to the linear relation, equation (5), as soon as the pore phase has a zero (or negligibly small) property value, the case that is the main focus of this paper. That means the non-linearity of self-consistent models is a consequence of the pore phase property and not, as should be, an intrinsic feature of the porous microstructure itself. In particular, according to the self-consistent approximation, materials with vacuum pores would exhibit a linear dependence of the elastic moduli and thermal conductivity. This is in contrast to what is observed in reality. It is thus clear that, when the pore phase has a zero or negligibly small property value, self-consistent models, including the popular Bruggeman model for thermal conductivity (sometimes, misleadingly, called ‘EMT’ model), are not only unrealistic but simply redundant, because they are in this case identical to the linear relation (non-interaction approximation), equation (5). Therefore they may be ignored in this context completely.

Non-linear model relations that obey the rigorous bound, equations (1)–(3), and at the same time reduce to the linear relation, equation (4), in the low-porosity case are the Maxwell-type model, (6) in conductivity context often called ‘Maxwell–Eucken model’Citation14 which is formally identical to the upper two-point bound (upper HS bound) and the Hasselman relation for elastic moduli,Citation51 the Coble–Kingery relation,Citation52CitationCitation54 (7) which is usually incorrectly cited in textbooks, the power–law relation,Citation55 (8) which corresponds to the differential scheme approximationCitation56CitationCitationCitation59 and includes the Gibson–Ashby model for open-cell foamsCitation60 as a special case) and our exponential relation,Citation16,Citation32,Citation61,Citation62 (9)

In all these relations the value of depends on the property in question and the pore shape and may therefore be called a ‘property-pore-shape coefficient’. Moreover, in the case of elastic moduli, this coefficient depends also on the solid Poisson ratio, but for Young's modulus this dependence is completely negligible (at least for solid Poisson ratios in the range 0.1–0.4).Citation32 For spherical pores (and approximately for other isometric convex pore shapes with aspect ratios around unity) the value of this coefficient is 2 for Young's modulus and 3/2 for thermal conductivity.Citation16 For strongly anisometric pores these values are different (tending to infinity for infinitely thin, oblate pores, i.e. microcracks, and to 2.34 and 1.67 for prolate pores in the case of and , respectivelyCitation46,Citation47). It has to be emphasised that the values can never be identical for and (at least as long as the solid phase is non-auxeticCitation32). The set of non-linear model relations given above for isotropic porous materials (equations (6)–(9)) is exhaustive in the sense that these relations are the only empirical-parameter-free model relations that reduce to the exact solution, equation (5), for vanishingly small porosity and at the same time do not violate the upper bounds (i.e. inter alia, they ensure that when , when ). All other relations occurring in the literature do not fulfil a least one of these criteria and are therefore ignored here. In particular, there are many more or less useful fit relations, both for the elastic moduli and for thermal conductivity. Among the most popular of these are two-parameter fit relations (generalised power-law relations), which are attributed to Phani and NiyogiCitation63 in elasticity context and to McLachlanCitation64 in conductivity context. Less well known, but equally useful, are our one-parameter fit relations, which are based on the Coble–Kingery-type relations, equation (7), but include a parameter that can be interpreted as a percolation threshold.Citation65,Citation66 On the other hand, simple (Spriggs) exponential relations,Citation67 which are preferentially used in the context of so-called minimum solid area models,Citation10,Citation11,Citation17Citation25 necessarily violate the upper bounds for sufficiently high porosity and should therefore be avoided altogether, both for predictive and fitting purposes.

Porosity dependence of the Poisson ratio

Due to the tensorial character of elastic properties (note that the proportionality coefficient in Hooke's law is a fourth-order tensor, so that even after considering all property symmetries the stiffness matrix has 21 elastic constants in the case of triclinic monocrystals, which are, of course, further reduced for materials of higher symmetryCitation33) two elastic constants are needed for a complete description of the elastic behaviour even in the simplest case of isotropic materials. Therefore it would be extremely convenient to dispose of a simple estimate for the Poisson ratio of porous materials. Unfortunately, however, the porosity dependence of the Poisson ratio is an unsolved problem so far. Nevertheless, when auxetic behaviour is ignored, most authors agree on the fact that with increasing porosity the effective Poisson ratio approaches an asymptotic value. This value has been differently identified to be 0.333, 0.25 or 0.2.Citation1, Citation7, Citation68 Although experimental evidence is not unambiguous,Citation69 there are strong arguments in favour of the latter value.Citation70 In particular, both the differential scheme approximationCitation59 and the self-consistent approximationCitation50 predict the Poisson ratio of porous materials to approach 0.2 with increasing porosity.Citation70 Moreover, according to the Mori–Tanaka approach,Citation71 also materials with microstructures realising the upper HS bounds tend to this value.Citation70 In particular, all these models and also the model relations mentioned above (linear, power-law and exponential relation) lead to the conclusion that the effective Poisson ratio of porous materials with solid Poisson ratio does not dependent on porosity at all.Citation70

Cross-property relations

Cross-property relations connect the effective or relative values of one property to those of another, different one. They can take the form of equations or inequalities and may be universally valid or not. The most well-known cross-property relation is probably the Levin relationCitation72 for the calculation of the effective thermal expansion coefficient of two-phase materials (composites).Citation6CitationCitation8 For porous materials, this relation leads to the conclusion that the effective thermal expansion coefficient is independent of porosity.Citation9

Another relatively well-known cross-property relation (in the form of an inequality) is the Milton–Torquato cross-property bound (MT bound), a so-called ‘elementary’ bound between the relative bulk modulus and the relative thermal conductivity.Citation6Citation,Citation8 For isotropic porous materials with non-negative solid Poisson ratio (, i.e. when the solid phase or phase mixture itself is assumed to be non-auxetic) the MT bound can be written asCitation73 (10)

In this relation is the relative bulk modulus ( being the effective bulk modulus of the porous material and the bulk modulus of the dense, pore-free solid) and is the relative thermal conductivity. Invoking the elasticity standard relationsCitation33 (11) and(12) and their counterparts for the dense solid (completely analogous, just replacing , , and by , , and , respectively) the corresponding inequality (here called MT bound) for the shear modulus can be calculated as(13) and for the tensile modulus (Young's modulus) as(14)

Note that, for these cross-property bounds to be valid, non-auxeticity of the porous materials as a whole is not required (only of the solid, see above). However, simpler, but also weaker (i.e. less restrictive, less useful), bounds are available when the porous material is non-auxetic.Citation73

Less well known, but more restrictive (and thus more useful) are the so-called ‘translation’ bounds, which have been derived by Berryman and MiltonCitation74 for porous materials and by Gibiansky and TorquatoCitation75 for general isotropic two-phase composites (here called BMGT bounds). For isotropic porous materials with a non-auxetic solid phase (or phase mixture), i.e. , the BMGT bound for the bulk modulus can be written asCitation73 (15)

The corresponding results for the shear modulus and the tensile modulus are obtained in an obvious way by applying the elasticity standard relations, equations (11) and (12), i.e.(16) (17)

In the extreme case the BMGT bound for the bulk modulus, equation (15), reduces to the MT bound for the bulk modulus, equation (10), and the bounds for and are simplified accordingly, whereas for (i.e. incompressible materials) the right hand sides of the bounds for and approach infinity (∞) and thus become useless in practice. Note that for incompressible materials the bulk modulus is infinity per definitionem and not bounded by the thermal conductivity at all.

It should be emphasised that the cross-property bounds for the bulk modulus depend only on the solid Poisson ratio , i.e. the Poisson ratio of the solid phase (or phase mixture), whereas the cross-property bounds for the shear and tensile moduli ( and ) depend also on the effective Poisson ratio of the porous material as a whole. However, assuming that the effective Poisson ratio usually approaches the value of 0.2 (see the discussion above), it is clear that the effective Poisson ratio cannot decrease to values below 0.2 as long as the solid Poisson ratio is higher than 0.2 (and, vice versa, cannot increase to values above 0.2, as long as is lower than 0.2).

Both the (very wide) MT cross-property bounds and the (more restrictive) BMGT cross-property bounds allow the calculation of upper bounds on the elastic moduli of isotropic porous materials as soon as the thermal conductivity of these materials is known (e.g. from measurements or finite element calculations). Cross-property bounds of this type are generally better (i.e. more precise, more restrictive or, in other words, lower) than the upper HS bounds, i.e. the best bounds that can be calculated based on the knowledge of porosity alone (for isotropic porous materials). The reason is that prior knowledge of one (effective or relative) property, here the thermal conductivity, which is a result not only of the porosity, but implicitly reflects all microstructural details, allows one to elegantly circumvent the explicit quantification of these microstructural details for determining another property (here elastic moduli). At the same time both MT and BMGT bounds may serve as a cross-check for the admissibility of the following cross-property relations.

In contrast to the cross-property bounds above, which are inequalities, cross-property relations in the form of equations (equalities) principally do not claim to be of universal validity, but their advantage is that they are extremely useful in practice, because they may lead to very concrete predictions of effective property values, as soon as some very rudimentary qualitative information on the type of microstructure beyond volume fractions and isotropy is available.

The most trivial cross-property relation of this type is valid for anisotropic porous materials with translational symmetry: in the direction of translation the relative axial Young's modulus is indeed equal to the relative axial thermal conductivity, i.e.(18) (cross-property relation based on the parallel model, equation (4), which is formally identical to the upper one-point bound = upper WP bound, see equation (1)). Of course, as discussed above, for isotropic porous materials this relation cannot be valid. Probably the earliest attempt to set up a cross-property relation of this type for isotropic porous materials has been undertaken by Sevostianov et al.Citation76, Citation77 using the Mori–Tanaka approach.Citation71 Their cross-property relation is(19)

(Sevostianov, Kovácˇik and Simancˇík/SKS cross-property relation). More recently, we have shown that an excellent approximation to this SKS cross-property relation can be derived in a simple way from the Maxwell-type model, equation (6) (which is formally identical to the upper two-point bound = upper HS bound, equations (2) and (3)).Citation73 Thus, for practical purposes, the SKS cross-property relation can be replaced by the much simpler cross-property relation,(20)

It can easily be shown that this cross-property relation deviates from the exact SKS cross-property relation by less than 4.3% for solid Poisson ratios in the range (difference in relative properties < 0.0011) and by less than 1.4% for (difference in relative properties <0.0003).Citation73

Nevertheless, although the upper HS bounds correspond theoretically to realisable microstructures,Citation6CitationCitation8 such microstructures are at best extremely rare among real-world materials. In fact, in the field of inorganic materials we are not aware of a single example reported in the literature for which the property–porosity dependence is well described by the upper HS bound (Maxwell model). Therefore, for the vast majority of real-world material microstructures the SKS cross-property relation and its excellent approximation, equation (20), cannot be expected to provide realistic property predictions. Similar conclusions hold for the cross-property relations (21) and(22) which can be derived in an elementary way from the linear relations and the Coble–Kingery relations, equations (5) and (7), for Young's modulus and thermal conductivity, respectively.

On the other hand, in a previous paperCitation78 we have shown that the cross-property relation for isotropic porous materials obeying either the power–law relation, equation (8), or our exponential relation, equation (9), is the same, viz. (23)

This extremely simple cross-property relation is valid for microstructures exhibiting either a power-law or an exponential dependence of the properties on porosity and can therefore be expected to be realistic at least for all microstructures in between these two limits (actually, recent research results indicate its validity far beyond these limits). This fact makes this cross-property relation more general than any of the other cross-property relations, although the detailed determination of its limits of ‘validity beyond’ is a desideratum of future research.

A detailed comparison of these cross-property relations with experimental values is beyond the scope of this paper, but it should have become clear to the reader that all cross-property relations for isotropic porous materials, irrespective of one or the other is more realistic for the material in question, lead to the conclusion that the porosity dependence of the (relative) thermal conductivity is never identical with that of the (relative) Young's modulus. The same is true for other elastic moduli. In other words, all cross-property relations lead to the conclusion that for isotropic porous materials (and most anisotropic materials as well, except for those with translational symmetry in the axial direction) we have . Of course, this conclusion is in direct contradiction with minimum solid area models which claim that .Citation10,Citation11,Citation17,Citation18 This claim is wrong. Therefore it is clear that minimum solid area models (also called minimum contact area models)Citation16CitationCitationCitationCitationCitationCitationCitationCitationCitation25 represent a blind alley of engineering science that should be abandoned.

Summary, conclusion and outlook

Based on a rational micromechanical framework and using the rigorous one-point (Voigt-Wiener) and two-point (HS) bounds, a critical assessment of model relations describing the porosity dependence of elastic properties (in particular Young's modulus) and thermal properties (in particular thermal conductivity) has been given. It has been shown that – in contrast to common belief and the impression evoked by many publications – the set of admissible and useful model relations for calculating the Young's modulus and thermal conductivity of isotropic porous materials is not very large. More specifically, when empirical fit relations with freely assignable fit parameters are ignored, the set of admissible predictive model relations reduces essentially to five types: linear relations (for small porosities), Maxwell-type relations, Coble–Kingery-type relations, power–law relations (corresponding to the well-known Gibson–Ashby relations for open-cell foams) and our (Pabst–Gregorová-type) exponential relations. Note that the popular simple (Spriggs-type) exponential relations have to be a priori discarded from this set of potentially predictive relations, because they necessarily violate the upper bounds for sufficiently high porosity. Finally, it is shown that – due to the same underlying geometrical models – there is a complete formal analogy between the model relations for the elastic moduli and thermal conductivity of porous materials and that, as a consequence, non-trivial cross-property relations exist between the relative Young's modulus and the relative thermal conductivity of isotropic porous materials. These cross-property relations arise due to rigorous mathematics and are not to be considered as empirical findings. Irrespective of the question as to which of the cross-property relations is optimal for a certain microstructure or type of porous material, all cross-property relations for isotropic porous materials lead to the conclusion that the relative Young's modulus (or any other elastic modulus) is not equal to the relative thermal conductivity. Thus, the mere existence of non-trivial cross-property relations of this type provides a refutation of the simplistic assumption of so-called minimum solid area models (also called minimum contact area models) that the relative Young's modulus equals the relative thermal conductivity. As a consequence, minimum solid or contact area models represent a blind alley of engineering science that should be abandoned.

Acknowledgement

This work is part of the project ‘Preparation and characterisation of oxide and silicate ceramics with controlled microstructure and modelling of microstructure–property relations’ (GA15-18513S), supported by the Czech Science Foundation (GACˇR).

Notes

*PaulCitation29 was apparently the first to recognise the significance of Voigt- and Reuss-type relations, i.e. arithmetic and harmonic volume-weighted means, respectively, as upper and lower bounds for multiphase materials.

Of course, these ‘microstructure-insensitive’ bounds and all model relations mentioned below remain valid only as long as elastic moduli and thermal conductivity are well defined via their corresponding linear constitutive equations, i.e. as long as the materials behave in a linear elastic manner according to Hooke's law,Citation33 i.e. do not exhibit non-linear elastic or elastoplastic behaviour, and heat transfer obeys Fourier's law,Citation5 i.e. does not involve significant radiative or convective mechanisms.Citation34

References

  • L. J. Gibson and M. F. Ashby: ‘The mechanics of foams: basic results’, 2nd edn, Chap. 5, ‘Cellular solids – structure and properties’, 175–234; 1997, Cambridge, Cambridge University Press.
  • L. J. Gibson and M. F. Ashby: ‘Thermal, electrical and acoustic properties of foams’, 2nd edn, Chap. 7, ‘Cellular solids – structure and properties’, 283–308; 1997, Cambridge, Cambridge University Press.
  • Cellular ceramics – structure, manufacturing, properties and applications’, (ed. M. Scheffler and P. Colombo), 1–645; 2005, Weinheim, Wiley-VCH.
  • Cellular and porous materials – thermal properties simulation and prediction’, (ed. A. Öchsner et al.), 1–422; 2008, Weinheim, Wiley-VCH.
  • W. Pabst: ‘The linear theory of thermoelasticity from the viewpoint of rational thermomechanics’, Ceram. Silik., 2005, 49, 254–263.
  • S. Torquato: ‘Random heterogeneous materials – microstructure and macroscopic properties’, 403–646; 2002, New York, Springer.
  • K. Markov: ‘Micromechanics of heterogeneous media’, in ‘Heterogeneous media – micromechanics modeling and simulations’, (ed. K. Markov and L. Preziosi), 1–162; 2000, Boston, Birkhäuser.
  • G. W. Milton: ‘The theory of composites’, 75–498; 2002, Cambridge, Cambridge University Press.
  • W. Pabst and E. Gregorová: ‘Effective thermal and thermoelastic moduli of alumina, zirconia and alumina-zirconia composite ceramics’ in ‘New developments in materials science research’ (ed. B. M. Caruta), 77–137; 2007, New York, Nova Science Publishers.
  • R. W. Rice: ‘Porosity and microcrack dependence of elastic properties at low to moderate temperatures’, Chap. 3, ‘Porosity of ceramics’, 100–167; 1998, New York, Marcel Dekker.
  • R. W. Rice: ‘The porosity and microcrack dependence of thermal and electrical conductivities and other nonmechanical properties’, Chap. 7, ‘Porosity of ceramics’, 315–374; 1998, New York, Marcel Dekker.
  • M. Kaviany: ‘Conduction heat transfer’ in ‘Principles of heat transfer in porous media’, 119–156; 1995, New York, Springer.
  • J. B. Wachtman: ‘Mechanical properties of porous ceramics’ in ‘Mechanical properties of ceramics’, 409–412; 1996, New York, Wiley-Interscience.
  • L. Gong, Y. Wang, X. Cheng, R. Zhang and H. Zhang: ‘Thermal conductivity of highly porous mullite materials’, Int. J. Heat Mass Transfer, 2013, 67, 253–259. doi: 10.1016/j.ijheatmasstransfer.2013.08.008
  • J. A. Choren, S. M. Heinrich and M. B. Silver-Thorn: ‘Young's modulus and volume porosity relationships for additive manufacturing applications’, J. Mater. Sci., 2013, 48, 5103–5112. doi: 10.1007/s10853-013-7237-5
  • W. Pabst, E. Gregorová and G. Tichá: ‘Effective properties of suspensions, composites and porous materials’, J. Eur. Ceram. Soc., 2007, 27, 479–482. doi: 10.1016/j.jeurceramsoc.2006.04.169
  • R. W. Rice: ‘Evaluation and extension of physical property–porosity models based on minimum solid area’, J. Mater. Sci., 1996, 31, 102–118. doi: 10.1007/BF00355133
  • R. W. Rice: ‘Comparison of physical property–porosity behaviour with minimum solid area models’, J. Mater. Sci., 1996, 31, 1509–1528. doi: 10.1007/BF00357860
  • A. K. Mukhopadhyay and K. K. Phani: ‘Young's modulus–porosity relations: an analysis based on a minimum contact area model’, J. Mater. Sci., 1998, 33, 69–72. doi: 10.1023/A:1004385327370
  • C. Reynaud and F. Thevenot: ‘Porosity dependence of mechanical properties of porous sintered SiC. Verification of the minimum solid area model’, J. Mater. Sci. Lett., 2000, 19, 871–874. doi: 10.1023/A:1006741616088
  • M. L. Hentschel and N. W. Page: ‘Elastic properties of powders during compaction. Part 3: evaluation of models’, J. Mater. Sci., 2006, 41, 7902–7925. doi: 10.1007/s10853-006-0875-0
  • K. U. O'Kelly, A. J. Carr and B. A. O. McCormack: ‘Minimum solid area models applied to the prediction of Young's modulus for cancellous bone’, J. Mater. Sci.: Mater. Med., 2003, 14, 379–384.
  • H. N. Yoshimura, A. L. Molisani, N. E. Narita, P. F. Cesar and H. Goldenstein: ‘Porosity dependence of elastic constants in aluminum nitride ceramics’, Mater. Res., 2007, 10, 127–133. doi: 10.1590/S1516-14392007000200006
  • M. Lubrick, R. G. Maev, F. Severin and V. Leshchynsky: ‘Young's modulus of low-pressure cold sprayed composites: an analysis based on a minimum contact area model’, J. Mater. Sci., 2008, 43, 4953–4961. doi: 10.1007/s10853-008-2729-4
  • L.-H. He, O. C. Standard, T. T. Y. Huang, B. A. Latella and M. V. Swain: ‘Mechanical behaviour of porous hydroxyapatite’, Acta Biomater., 2008, 4, 577–586. doi: 10.1016/j.actbio.2007.11.002
  • O. Wiener: ‘Die Theorie des Mischkörpers für das Feld der stationären Strömung’, Abh. Math. Phys. Klasse Königl. Sächs. Ges. Wiss., 1912, 32, 509–604.
  • W. Voigt: ‘Über die Beziehung zwischen den beiden Elastizitätskonstanten isotroper Körper’, Wiedemann Ann., 1889, 274, 573–587. doi: 10.1002/andp.18892741206
  • A. Reuss: ‘Berechnung der Fliessgrenze von Mischkristallen auf Grund der Plastizitätsbedingung für Einkristalle’, Z. Angew. Math. Mech., 1929, 9, 49–58. doi: 10.1002/zamm.19290090104
  • B. Paul: ‘Prediction of the elastic constants of multiphase materials’, Trans. AIME, 1960, 218, 36–41.
  • Z. Hashin and S. Shtrikman: ‘A variational approach to the theory of the effective magnetic permeability of multiphase materials’, J. Appl. Phys., 1962, 33, 3125–3131. doi: 10.1063/1.1728579
  • Z. Hashin and S. Shtrikman: ‘A variational approach to the theory of the elastic behaviour of multiphase materials’, J. Mech. Phys. Solids, 1963, 11, 127–140. doi: 10.1016/0022-5096(63)90060-7
  • W. Pabst and E. Gregorová: ‘Effective elastic properties of alumina–zirconia composite ceramics – Part 2: micromechanical modeling’, Ceram. Silik., 2004, 48, 14–23.
  • W. Pabst and E. Gregorová: ‘Effective elastic properties of alumina–zirconia composite ceramics – Part 1: rational continuum theory of linear elasticity’, Ceram. Silik., 2003, 47, 1–7.
  • W. Pabst and J. Hostaša: ‘Thermal conductivity of ceramics – from monolithic to multiphase, from dense to porous, from micro to nano’ in ‘Advances in materials science research’, (ed. M. C. Wythers), Vol. 7, 1–112; 2011, New York, Nova Science Publishers.
  • J. K. Carson, S. J. Lovatt, D. J. Tanner and A. C. Cleland: ‘Predicting the effective conductivity of unfrozen, porous food’, J. Food. Eng., 2006, 75, 297–307. doi: 10.1016/j.jfoodeng.2005.04.021
  • W. Pabst and E. Gregorová: ‘The sigmoidal average – a powerful tool for predicting the thermal conductivity of composite ceramics’, J. Phys.: Conf. Ser., 2012, 395, 012021.
  • W. Pabst, E. Gregorová, I. Sedlárˇová and M. Cˇerný: ‘Preparation and characterization of porous alumina-zirconia composite ceramics’, J. Eur. Ceram. Soc., 2011, 31, 2721–2731. doi: 10.1016/j.jeurceramsoc.2011.01.011
  • R. Hill: ‘The elastic behaviour of a crystalline aggregate’, Proc. Phys. Soc. Lond. A, 1952, 65, 349–354. doi: 10.1088/0370-1298/65/5/307
  • E. Kröner: ‘Berechnung der elastischen Konstanten des Vielkristalls aus den Konstanten des Einkristalls’, Z. Phys., 1958, 151, 504–518. doi: 10.1007/BF01337948
  • W. Pabst, G. Tichá and E. Gregorová: ‘Effective elastic properties of alumina–zirconia composite ceramics – Part 3: calculation of elastic moduli of polycrystalline alumina and zirconia from monocrystal data’, Ceram. Silik., 2004, 48, 41–48.
  • W. Pabst, E. Gregorová, T. Uhlírˇová and A. Musilová: ‘Elastic properties of mullite and mullite-containing ceramics – Part 1: theoretical aspects and review of monocrystal data’, Ceram. Silik., 2013, 57, 265–274.
  • J. C. Maxwell: ‘A treatise on electricity and magnetism, Vol. 1’, 435–449; 1873, Oxford, Clarendon Press ( reprint 1954, New York, Dover).
  • J. M. Dewey: ‘The elastic constants of materials loaded with non-rigid fillers’, J. Appl. Phys., 1947, 18, 578–581. doi: 10.1063/1.1697691
  • J. K. Mackenzie: ‘Elastic constants of a solid containing spherical holes’, Proc. Phys. Soc. Lond. B, 1950, 63, 2–11. doi: 10.1088/0370-1301/63/1/302
  • J. D. Eshelby: ‘The determination of the elastic field of an elliptical inclusion, and related problems’, Proc. R. Soc. Lond. A, 1957, 241, 376–396. doi: 10.1098/rspa.1957.0133
  • W. Pabst and E. Gregorová: ‘Conductivity of porous materials with spheroidal pores’, J. Eur. Ceram. Soc., 2014, 34, 2757–2766. doi: 10.1016/j.jeurceramsoc.2013.12.040
  • W. Pabst and E. Gregorová: ‘Young's modulus of isotropic porous materials with spheroidal pores’, J. Eur. Ceram. Soc., 2014, 34, 3195–3207. doi: 10.1016/j.jeurceramsoc.2014.04.009
  • D. Bruggeman: ‘Berechnung verschiedener physikalischer Konstanten von heterogenen Substanzen’, Ann. Phys. (Leipzig), 1935, 416, 636–664. doi: 10.1002/andp.19354160705
  • R. Hill: ‘A self-consistent mechanics of composite materials’, J. Mech. Phys. Solids, 1965, 13, 213–222. doi: 10.1016/0022-5096(65)90010-4
  • B. Budiansky: ‘On the elastic moduli of some heterogeneous materials’, J. Mech. Phys. Solids, 1965, 13, 223–227. doi: 10.1016/0022-5096(65)90011-6
  • D. P. H. Hasselman: ‘On the porosity dependence of the elastic moduli of polycrystalline refractory materials’, J. Am. Ceram. Soc., 1962, 45, 452–453. doi: 10.1111/j.1151-2916.1962.tb11191.x
  • R. L. Coble and W. D. Kingery: ‘Effect of porosity on physical properties of sintered alumina’, J. Am. Ceram. Soc., 1956, 39, 377–385. doi: 10.1111/j.1151-2916.1956.tb15608.x
  • W. Pabst and E. Gregorová: ‘Note on the so-called Coble–Kingery formula for the effective tensile modulus of porous ceramics’, J. Mater. Sci. Lett., 2003, 22, 959–962. doi: 10.1023/A:1024600627210
  • W. Pabst: ‘Simple second-order expression for the porosity dependence of thermal conductivity’, J. Mater. Sci., 2005, 40, 2667–2669. doi: 10.1007/s10853-005-2101-x
  • W. Pabst and E. Gregorová: ‘Derivation of the simplest exponential and power-law relations for the effective tensile modulus of porous ceramics via functional equations’, J. Mater. Sci. Lett., 2003, 22, 1673–1675. doi: 10.1023/B:JMSL.0000004645.77295.b5
  • S. Boucher: ‘On the effective moduli of isotropic two-phase elastic composites’, J. Compos. Mater., 1974, 8, 82–89. doi: 10.1177/002199837400800108
  • R. McLaughlin: ‘A study of the differential scheme for composite materials’, Int. J. Eng. Sci., 1977, 15, 237–244. doi: 10.1016/0020-7225(77)90058-1
  • A. N. Norris: ‘A differential scheme for the effective moduli of composites’, Mech. Mater., 1985, 4, 1–16. doi: 10.1016/0167-6636(85)90002-X
  • R. W. Zimmerman: ‘Elastic moduli of a solid containing spherical inclusions’, Mech. Mater., 1991, 12, 17–24. doi: 10.1016/0167-6636(91)90049-6
  • L. J. Gibson and M. F. Ashby: ‘The mechanics of three-dimensional cellular materials’, Proc. R. Soc. Lond. A, 1982, 382, 43–59. doi: 10.1098/rspa.1982.0088
  • W. Pabst and E. Gregorová: ‘Mooney-type relation for the porosity dependence of the effective tensile modulus of ceramics’, J. Mater. Sci., 2004, 39, 3213–3215. doi: 10.1023/B:JMSC.0000025863.55408.c9
  • G. Tichá, W. Pabst and D. S. Smith: ‘Predictive model for the thermal conductivity of porous materials with matrix-inclusion type microstructure’, J. Mater. Sci., 2005, 40, 5045–5047. doi: 10.1007/s10853-005-1818-x
  • K. K. Phani and S. K. Niyogi: ‘Young's modulus of porous brittle solids’, J. Mater. Sci., 1987, 22, 257–263. doi: 10.1007/BF01160581
  • D. S. McLachlan: ‘Equation for the conductivity of metal-insulator mixtures’, J. Phys. C: Solid State, 1985, 18, 1891–1897. doi: 10.1088/0022-3719/18/9/022
  • W. Pabst and E. Gregorová: ‘New relation for the porosity dependence of the effective tensile modulus of brittle materials’, J. Mater. Sci., 2004, 39, 3501–3503. doi: 10.1023/B:JMSC.0000026961.12735.2a
  • W. Pabst and E. Gregorová: ‘A new percolation-threshold relation for the porosity dependence of thermal conductivity’, Ceram. Int., 2006, 32, 89–91. doi: 10.1016/j.ceramint.2004.12.007
  • R. M. Spriggs: ‘Expression for effect of porosity on elastic modulus of polycrystalline refractory materials, particularly aluminum oxide’, J. Am. Ceram. Soc., 1961, 44, 628–629. doi: 10.1111/j.1151-2916.1961.tb11671.x
  • N. Ramakrishnan and V. S. Arunachalam: ‘Effective elastic moduli of porous ceramic materials’, J. Am. Ceram. Soc., 1993, 76, 2745–2752. doi: 10.1111/j.1151-2916.1993.tb04011.x
  • A. R. Boccaccini: ‘Comment on “Effective elastic moduli of porous ceramic materials”’, J. Am. Ceram. Soc., 1994, 77, 2779–2781. doi: 10.1111/j.1151-2916.1994.tb04679.x
  • W. Pabst and E. Gregorová: ‘The Poisson ratio of porous materials’ in ‘Characterisation of porous solids VIII’, (ed. S. Kaskel et al.), 424–431; 2009, Cambridge, RSC Publishing.
  • T. Mori and K. Tanaka: ‘Average stress in matrix and average elastic energy of materials with misfitting inclusions’, Acta Metal., 1973, 21, 571–574. doi: 10.1016/0001-6160(73)90064-3
  • V. M. Levin: ‘Thermal expansion coefficients of heterogeneous materials’, Mech. Solids, 1967, 21, 9–17.
  • W. Pabst and E. Gregorová: ‘Cross-property relations between elastic and thermal properties of porous ceramics’, Adv. Sci. Technol., 2006, 45, 107–112. doi: 10.4028/www.scientific.net/AST.45.107
  • J. G. Berryman and G. W. Milton: ‘Microgeometry of random composites and porous media’, J. Phys. D: Appl. Phys., 1988, 21, 87–94. doi: 10.1088/0022-3727/21/1/013
  • L. V. Gibiansky and S. Torquato: ‘Connection between the conductivity and bulk modulus of isotropic composite materials’, Proc. R. Soc. Lond. A, 1996, 452, 253–283. doi: 10.1098/rspa.1996.0015
  • I. Sevostianov, J. Kovácˇik and F. Simancˇík: ‘Correlation between elastic and electric properties for metal foams: theory and experiment’, Int. J. Fract., 2002, 114, 23–28. doi: 10.1023/A:1022674130262
  • I. Sevostianov, J. Kovácˇik and F. Simancˇík: ‘Elastic and electric properties of closed-cell aluminum foams – cross-property connection’, Mater. Sci. Eng. A, 2006, 420, 87–99. doi: 10.1016/j.msea.2006.01.064
  • W. Pabst and E. Gregorová: ‘A cross-property relation between the tensile modulus and the thermal conductivity of porous ceramics’, Ceram. Int., 2007, 33, 9–12. doi: 10.1016/j.ceramint.2005.07.009

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.