Proc. R. Soc. A (2006) 462, 1167–1180 doi:10.1098/rspa.2005.1614 Published online 17 January 2006
Mesoscale bounds in finite elasticity and thermoelasticity of random composites B Y Z. F. K HISAEVA 1
AND
M. O STOJA -S TARZEWSKI 1,2, *
1
Department of Mechanical Engineering, McGill University, Montre´al, Quebec H3A 2K6, Canada 2 McGill Institute for Advanced Materials, Ernest Rutherford Building, 3600 University Street, Montreal, Quebec H3A 2T8, Canada
Under consideration is the problem of size and response of the Representative Volume Element (RVE) in the setting of finite elasticity of statistically homogeneous and ergodic random microstructures. Through the application of variational principles, a scale dependent hierarchy of strain energy functions (i.e. mesoscale bounds) is derived for the effective strain energy function of the RVE. In order to account for thermoelastic effects, the variational principles are first generalized, and then analogous bounds on the effective thermoelastic response are derived. It is also shown that, in contradiction to results obtained for random linear composites, the hierarchy on the effective strain energy function in nonlinear elasticity cannot be split into volumetric and isochoric terms, while the hierarchy on the effective free energy function cannot be divided into purely mechanical and thermal contributions. Keywords: random composites; representative volume element; mesoscale bounds; finite elasticity; finite thermoelasticity
1. Introduction The prediction of macroscopic properties of heterogeneous materials is of great importance in many engineering applications. Evidently, any material displays heterogeneity on a microscale and has different properties depending on its inherent physical and chemical composition. The issue which commonly arises when dealing with such materials—especially now, in the age of bio- and nanotechnology—is the validity of the separation of scales of continuum mechanics: d ! L!!Lmacro ; ð1:1Þ where L is the size of the representative volume element (RVE). As is well known, (1.1) allows one to smooth the heterogeneous microstructure, characterized by the microscale d, and work with an effective continuum over the range of scales from L through Lmacro. The RVE is clearly set-up in two basic cases (Ostoja-Starzewski 1999, 2001): (i) a unit cell in a periodic microstructure and (ii) a domain containing * Author and address for correspondence: Department of Mechanical and Industrial Engineering, University of Illinois at Urbana-Champaign, Urbana, IL 61801, USA. (
[email protected]). Received 6 July 2005 Accepted 14 November 2005
1167
q 2006 The Royal Society
1168
Z. F. Khisaeva and M. Ostoja-Starzewski
infinitely many microscale elements (e.g. crystals) in a random medium. Overwhelmingly, in the case of random media, the prediction of effective properties of heterogeneous materials is usually based on an a priori assumption of existence of the RVE, without clearly specifying its size L relative to d. In this paper, we develop scale-dependent bounds on the effective response of random composites for physically nonlinear materials subjected to large isothermal and non-isothermal deformations. We follow the approach that originated some 15 years ago (Huet 1990; Sab 1992), also driven by the need to derive continuum random field models from microstructures (Ostoja-Starzewski & Wang 1989; see also the review Ostoja-Starzewski 2001). We shall deal with ‘apparent’ material properties at finite strain, that is, when the body is smaller than the RVE and when deformations are large. The idea is to consider properties of a composite material of the size smaller than the representative volume and apply variational principles along with different types of boundary conditions to investigate scale-dependent relationships for various types of material microstructure. These results lay the ground for a quantitative estimation of the real size of the RVE, currently under study.
2. General theory Consider a thermoelastic body undergoing equilibrium deformation in a threedimensional Euclidian space. Let x and X denote the position vectors of a particle inside the body in reference and current configurations, respectively; then the deformation of the body can be described as x Z cðXÞ. The deformation gradient tensor is given by FZ
vcðXÞ Z Grad x Z Grad u C 1 vX
or Fij Z
vxi vu Z i C dij ; vXj vXj
ð2:1Þ
where Grad u is the displacement gradient in material description, dij is the Kronecker delta, and the indices, i, j, k, take on the values 1, 2, 3. Equation of equilibrium in the absence of body forces has the form Div P Z 0
or
vPij Z 0: vXj
ð2:2Þ
Here, P denotes a generally non-symmetric first Piola–Kirchhoff stress tensor and Div is the divergence calculated with respect to the reference configuration. The key assumption of the finite hyperelasticity theory is the existence of a strain energy function jZj(Fij) per unit volume of an undeformed body, which depends on the deformation of the object and its material properties. Restricting ourselves to the reference configuration, the equation of state of the material can be written as Pij Z
vj : vFij
ð2:3Þ
If temperature changes are to be considered, the strain energy density should be replaced by the free energy as a function of both deformation gradient and temperature. Proc. R. Soc. A (2006)
Bounds in finite elasticity and thermoelasticity
1169
It is worth mentioning that the first Piola–Kirchhoff stress tensor and deformation gradient tensor are the most convenient tensors to be used since the averaging theorems and Hill condition can be formulated for these quantities in the way analogous to the infinitesimal deformation theory (for a discussion of these theorems, the reader is referred to §4). Combining (2.3) and (2.1), equation (2.2) can be rewritten as v vj Z 0: ð2:4Þ vXj vui;j Here and elsewhere, ui,j denotes the deformation gradient tensor, where the unit tensor is omitted for simplicity of notation. 3. Minimum theorems in finite elasticity and thermoelasticity Variational principles for a material undergoing finite strains have been studied extensively by many writers (e.g. Zubov 1971; de Veubeke 1972; Lee & Shield 1980; Ogden 1984; Gao 1999). If the principle of stationary potentional energy in finite elasticity can be formulated similarly to that in linear elasticity, the complementary energy principle cannot be established without certain restrictions placed on the strain-energy function. In nonlinear elasticity, the strain-energy function may be non-convex and, therefore, non-invertible, i.e. the deformation gradient cannot be expressed in terms of stress. Even the inversion of convex strain-energy functions is extremely difficult in general, which places a limitation on the use of the complementary energy theorem. To overcome the difficulty of inversion, different ‘complementary-type’ variational principles have been introduced in the literature. From the point of view of the bounding problem, the variational principles proposed by Lee & Shield (1980) seem more advantageous to use. It allows one to avoid a difficult and often impossible task of inverting the strain-energy density function. In §3a we outline this theory and in §3b extend it to the case of nonlinear thermoelastostatics. (a ) Variational principles in finite elastostatics Consider the functional PfUi g Z
ð V0
ð jðUi;j ÞdV K ti0 Ui dS;
ð3:1Þ
ST
where V0 is material volume in reference state, Ui is an admissible displacement field such that Ui Z ui0 on the portion of the boundary, where displacement is prescribed, SU, and t0i is prescribed nominal traction on the remaining part of the boundary ST. Equation (3.1) represents the finite elasticity counterpart of the principle of stationary potential energy in linear elasticity theory. The functional P{Ui} is stationary for the actual solution ui, which satisfies the equilibrium equation (2.4) in V0 and natural boundary conditions vj n Z ti0 vui;j j Proc. R. Soc. A (2006)
on ST :
ð3:2Þ
1170
Z. F. Khisaeva and M. Ostoja-Starzewski
Here, nj denotes an outward normal vector to the boundary surface in the reference configuration. Expanding the quantity P{uiCdui, ui,jCdui,j} in Taylor’s series and ignoring terms of the third and higher order in dui, one can obtain the condition under which the functional P{Ui} assumes a local minimum for the actual solution ui for all non-zero dui such that duiZ0 on SU (Lee & Shield 1980): ð v2 j dui;j dup;q dV O 0: ð3:3Þ V0 vui;j vup;q If the condition (3.3) holds, the principle of stationary potential energy becomes the minimum principle. Assuming interface continuity of the displacement and traction field inside the body, the above variational principle becomes valid for multiphase material as well as for materials with continuously varying properties and can be used to estimate bounds on the effective material response. We now turn to the investigation of the principle of minimum complementary energy. To overcome the difficulties of inversion in formulating the complementary energy functional, Lee & Shield (1980) proposed to consider the complementary strain-energy function jc to be a function of the deformation gradient rather than stresses: jc Z
vj u Kj: vui;j i;j
ð3:4Þ
Then, the complementary energy functional can be written as ð ð vj vj QfUij g Z Uij Kj dV K nk ui0 dS; V0 vUij SU vUij where Uij is an admissible displacement gradient field satisfying v vj vj n Z ti0 ; on ST : Z 0; in V0 and vXj vUij vUij j
ð3:5Þ
ð3:6Þ
Using the Lagrange multiplier method, it can be shown that the functional (3.5) is stationary for UijZui,j , where ui is the actual solution of a given elasticity problem. In order for Q to assume a local minimum, the following condition should be fulfilled: ð v2 j duij dupq dV O 0; ð3:7Þ V0 vui;j vup;q for all non-zero duij satisfying v v2 j dupq Z 0; vxj vui;j vup;q
in V0 ;
v2 j du n Z 0; vui;j vup;q pq j
on ST :
ð3:8Þ
Equation (3.7) was obtained by applying Taylor series expansion up to the second order in duij to the quantity Q{ui,jCduij}, taking into account that equations (3.8) need to be satisfied by both actual and admissible solution. Proc. R. Soc. A (2006)
Bounds in finite elasticity and thermoelasticity
1171
An application of the divergence theorem leads to Pfui g C Qfui;j g Z 0;
ð3:9Þ
which is a well-known identity from the calculus of variations between total potential and complementary energies. Equations (3.1)–(3.9) define variational principles in finite elastostatics for elastic bodies not subjected to constraints. For the case when internal constraints such as incompressibility are present, the components of the deformation gradient tensor are not arbitrary and additional conditions need to be considered in formulating minimum potential and complementary energy theorems (Lee & Shield 1980). (b ) Variational principles in finite thermoelastostatics Here, we consider an uncoupled thermoelastic problem when the material is at rest and the heat flow is steady. In this case, the equation of heat conduction can be solved separately from the remaining field equations and the temperature field T is determined uniquely by thermal boundary conditions. Minimum energy principles in this case can be derived using the Helmholtz free energy function jZj(F, T ) and the Gibbs thermodynamic potential GZG(P, T ) in place of the isothermal strain energy and the complementary strain energy, respectively. G is associated with j by the transformation vj vj GðP; TÞ Z jKP : F or G ; T Z jK u : ð3:10Þ vui;j vui;j i;j The potential energy functional (3.1) can be rewritten as ð ð PfUi g Z jðUi;j ; q0 ÞdV K ti0 Ui dS; V0
ð3:11Þ
ST
where q0 is the prescribed temperature field. For simplicity, we will assume temperature to be uniform throughout the volume. The Helmholtz free energy in nonlinear elasticity has the form (Holzapfel 2000) jðUi;j ; TÞ Z j0 ðUi;j Þ
T DT ^ Ke0 ðUi;j Þ C TðTÞ; T0 T0
ð3:12Þ
where T0 is the temperature at reference state, DTZTKT0 is the temperature change, j0 is the free energy at reference temperature, e0 is the internal energy at ^ is a purely thermal contribution defined as reference temperature and T ð TZq0 dT ^ ZK T cF ðTÞðq0 KTÞ : ð3:13Þ T TZT0 Here, cF(T ) is specific heat capacity at constant deformation. For infinitesimal elastic deformations (3.12) can be easily reduced to a well-known Helmholtz free energy form by taking j0 Z ð1=2ÞCijkl 3ij 3kl , e0 Z ð1=2ÞCijkl 3ij 3kl KCijkl 3ij akl T0 and assuming cF to be a constant over a given small temperature change. Varying independently the quantities ui and ui,j, while keeping the temperature q0 constant, it can be proved by the same procedure as used in Proc. R. Soc. A (2006)
1172
Z. F. Khisaeva and M. Ostoja-Starzewski
§3a that the functional (3.11) is stationary for the actual solution UiZui and provides a local minimum if ð v2 j0 v2 e0 T KDT ð3:14Þ dui;j dup;q dV O 0; vui;j vup;q vui;j vup;q V0 for all non-zero dui vanishing over the surface upon which the displacement is prescribed. Applying the transformation (3.10), the complementary-type functional becomes ð ð vj QfUij g ZK GðUij ; q0 ÞdV K nk ui0 dS; ð3:15Þ V0 SU vUij where G can be written as q0 vj0 DT ve0 ^ 0 Þ: G ZK u Kj0 C u Ke0 C Tðq T0 vui;j i;j T0 vui;j i;j The functional (3.15) assumes a local minimum provided ð v2 j0 v2 e0 T KDT duij dupq dV O 0; vui;j vup;q vui;j vup;q V0
ð3:16Þ
ð3:17Þ
for all non-zero duij satisfying (3.8), in which j is replaced by the free energy function (3.12). The energetic contribution e0 to (3.12) appears only in the modified entropic theory of rubber thermoelasticity, and is equal to zero in the case of a purely entropic theory, for which conditions (3.14) and (3.17) simply reduce to the isothermal inequalities (3.3) and (3.7). Note that (3.3) and (3.7) are the convexity conditions on j. They play an important role in hyperelasticity by ensuring the stability of the material; the strain-energy models fitted to available experimental data are commonly designed so as to avoid the lack of convexity (Holzapfel et al. 2000). 4. Averaging theorems in finite deformation theory Averaging theorems play a key role in the estimation of overall material properties of composites. It is known that, in a small deformation theory, volume averages of both infinitesimal strain and stress fields can be fully determined from the surface data (e.g. Nemat-Nasser & Hori 1999). In finite elasticity theory, the relation between strain and displacement gradient is nonlinear, which, as a result, leads to difficulties in evaluation of the volume averages. Accordingly, different deformation measures should be considered when dealing with a problem involving finite strain. In this section, we will formulate averaging theorems along with the Hill condition for two conjugate quantities, namely, the deformation gradient tensor and the first Piola–Kirchhoff stress tensor. (a ) Average deformation gradient and stress theorems Deformation gradient tensor is the simplest kinematic variable and the only deformation measure for which the average theorem can be derived in finite Proc. R. Soc. A (2006)
Bounds in finite elasticity and thermoelasticity
1173
elasticity theory. Assuming no jumps in displacement and traction across the boundary between different phases of the composite, the averaging theorems for the deformation gradient tensor and the first Piola–Kirchhoff stress tensor take the form ð 1 F ij h Fij ðXk ÞdV Z Fij0 ; ð4:1Þ V 0 V0 ð 1 P ij h Pij ðXk ÞdV Z Pij0 ; ð4:2Þ V 0 V0 where Fij0 is the prescribed deformation gradient and ti0 Z Pij0 nj is a prescribed traction field acting on the boundary S0 in the reference configuration, accordingly. The superposed bar here and elsewhere denotes the volume average of the overall macroscopic quantities. Equations (4.1) and (4.2) can be readily proved by transforming the volume integral via the Green–Gauss theorem. One of the main consequences of these theorems is that they allow one to control the value of the effective deformation gradient tensor and the first Piola–Kirchhoff stress tensor of the composite under consideration. Equations (4.1) and (4.2) were first considered by Hill (1972) and then reexamined by many authors (Nemat-Nasser 1999; Lo ¨hnert & Wriggers 2003; Costanzo et al. 2005). In fact, in finite elasticity, the average stress theorem holds for both spatial and material descriptions (Hill 1972), i.e. for both Cauchy and first Piola–Kirchhoff stress tensors. This is not surprising since both tensors represent the same resultant force acting on a surface element of the deformed body. It is worth mentioning, that (4.1) holds only for a static process in the absence of body forces. If this assumption fails, one has to consider the effective value of a target quantity, but not the ‘true’ average over the volume (Costanzo et al. 2005). (b ) Hill condition for finite deformations A well-known criterion in a homogenization problem is the so-called Hill condition. It represents a condition under which the mechanical and energetic definitions of the effective properties are compatible. For the deformation gradient tensor and the first Piola–Kirchhoff stress tensor, it can be stated as follows (Hill 1972; Nemat-Nasser 1999): ij F ij : Pij Fij Z P ð4:3Þ This identity is trivially satisfied by a homogeneous body, but imposes certain restrictions on the boundary conditions if the body is heterogeneous. Applying the Green–Gauss theorem and noting that from the equilibrium equation in the absence of body forces Pij;j Z 0, we get ð ij F ij Z 1 ij nj Þðui KðF ij Kdij ÞXj ÞdV : Pij Fij KP ðt KP ð4:4Þ V0 S0 i When the boundary conditions are such that the right-hand side integral is vanishing, the average of the product of the deformation gradient tensor and the stress tensor is equal to the product of averages. This implies three types of Proc. R. Soc. A (2006)
1174
Z. F. Khisaeva and M. Ostoja-Starzewski
boundary conditions, which allow the use of the Hill condition when dealing with inhomogeneous bodies: (i) uniform kinematic boundary condition (prescribing a volume average deformation gradient): ui Z ðFij0 Kdij ÞXj ;
cXj 2S0 ;
ð4:5Þ
(ii) uniform static boundary condition (prescribing a volume average nominal stress): ti Z Pij0 nj ; cXj 2S0 ; ð4:6Þ (iii) uniform orthogonal-mixed boundary condition: ti KPij0 nj ui KðFij0 Kdij ÞXj Z 0; cXj 2S0 ;
ð4:7Þ
where the averaging theorems §4a have been used. In the following, we shall use boundary conditions (4.5) and (4.6) along with variational principles introduced in §3 for estimation of bounds on the effective response of nonlinear composites.
5. Hierarchies of mesoscale bounds (a ) Energy bounds in finite elasticity and thermoelasticity of random composites When dealing with a random microstructure one needs to impose certain restriction on the internal geometry of the composite in order to be able to estimate average properties of the material and the RVE size itself. It turns out that sufficient conditions for the bounds on the effective properties to converge with the increasing sample size are the assumptions of statistical homogeneity and mean-ergodicity of the material. In the present case, spatial homogeneity implies that all points in a given realization B(u) of the random composite BZ{B(u); u2U}, where U is a sample space, are equivalent, i.e. there is no preferred location in the structure. The ergodicity assumption states that any one realization of the composite is a representative of the entire ensemble, ð 1 hFi Z lim F dV ; ð5:1Þ V0/N V0 V0 where h$i denotes ensemble average and V0 is the material volume in reference state. Next, we introduce a non-dimensional scale parameter dZL/d, where d is the heterogeneity size (e.g. single grain) and L is the size of a mesoscale domain. Consider a partition of a body of size d into n smaller square elements of size d 0 Z d=n (nZ2 in figure 1). Defining two types of boundary conditions (e.g. Ostoja-Starzewski 2001)—restricted (the boundary condition specified on the boundary of each element) and unrestricted (the boundary condition specified on the boundary of the whole body)—one observes that the deformation of the material under the restricted boundary condition represents an admissible field for the unrestricted boundary condition, but not conversely. Proc. R. Soc. A (2006)
1175
Bounds in finite elasticity and thermoelasticity
d
d' Figure 1. Partition of a window Bd of scale d into four square-shaped sub-windows Bd0 .
Consider now the uniform displacement boundary condition (4.5) (i.e. when the volume average deformation gradient is being prescribed). In this case, ST is identically equal to zero and the functional (3.1) reduces to ð PfUi g Z jðUi;k ; Xj ÞdV ; ð5:2Þ V0
where the strain-energy function is now a function of both deformation gradient and coordinates. Under the assumption that the minimum theorem holds the energy stored in the body under the restricted boundary conditions is related to the unrestricted one as Jðu; F 0 Þ% Jr ðu; F 0 Þ; ð5:3Þ Ð where Jðu; F 0 ÞZ V0 jðu; X; FÞdV , is the elastic energy of any given realization B(u) of the composite, and the superscript ‘r’ denotes the effective properties obtained under restricted boundary conditions. Henceforth, in order to simplify the notation we shall write j instead of j(u, X, F). With D and 1 denoting the RVE size and inhomogeneity size, respectively, upon ensemble averaging we obtain for the upper bound JðF 0 Þ D % JðF 0 Þ d % JðF 0 Þ d0 % JðF 0 Þ 1 ; for 1! d 0 ! d! D: ð5:4Þ The hierarchy (5.4) can be generalized for the case of thermoelastostatics to be JðF 0 ; q0 Þ D % JðF 0 ; q0 Þ d % JðF 0 ; q0 Þ d0 % JðF 0 ; q0 Þ 1 ; for 1! d 0 ! d! D: ð5:5Þ We now turn to the investigation of an analogous reciprocal expression for the lower bounds. Under the uniform traction boundary condition (4.6), the functional Q reduces to ð vj QfUik g Z Uik Kj dV : ð5:6Þ V0 vUik Proc. R. Soc. A (2006)
1176
Z. F. Khisaeva and M. Ostoja-Starzewski
Thus, the following inequality between responses under restricted and unrestricted boundary conditions holds:
0
Ð
J ðu; P 0 Þ% Jr ðu; P 0 Þ;
ð5:7Þ
where J ðu; P ÞZ V0 fðvj=vUij ÞUij KjgdV : From this, upon ensemble averaging, we can derive a scale dependent hierarchy of a lower bound on the effective properties 0 J ðP Þ D % J ðP 0 Þ d % J ðP 0 Þ d0 % J ðP 0 Þ 1 ; for 1! d 0 ! d! D; ð5:8Þ which in thermoelastostatics takes the form GðP 0 ; q0 Þ D R GðP 0 ; q0 Þ d R GðP 0 ; q0 Þ d0 R GðP 0 ; q0 Þ 1 ; for 1! d 0 ! d! D: ð5:9Þ Noting that ðvj=vUik ÞUik can be equivalently expressed as P : F we can write J ðu; P 0 Þ Z V0 ðP : F ÞKJðu; P 0 Þ Z V0 ðP 0 : FÞKJðu; P 0 Þ; where averaging theorems §4a have been employed. At the macroscale, we will always recover the regular Legendre transformation J Z P : F KJ ;
ð5:11Þ
since for the RVE size composite material J ðu; P 0 ÞZ J ðu; F 0 Þ and Jðu; F 0 ÞZ Jðu; P 0 Þ. Equation (5.11) allows one to estimate effective strain energy function as d/N. Note that, strictly speaking, D (denoting the RVE size L relative to the microheterogeneity d) is infinite for a random medium lacking any spatial periodicity, but the convergence of upper and lower bounds is achieved within an accuracy of a few percent at finite scales; see e.g. Jiang et al. (2001) and Ostoja-Starzewski & Castro (2003) for case studies of two physically nonlinear composites. Now, suppose that we have a composite with two different phases (one stiffer than the other), it follows from our theory that, for the lower bound, C V0 P 0 : F ; for d! D: ð5:12Þ Jðu; P 0 Þ D R Jðu; P 0 Þ d KV0 P 0 : F d D Let us denote the right-hand side of the inequality (5.12) by jL. Under the in the body, the application of assumption of constant stress PðXÞZ P restricted boundary conditions does not alter the displacement field, so that Z F , which, as V0 /N, leads to F D d PÞ: JL Z jð
ð5:13Þ
Equation (5.13) is a generalization of the well-known Reuss bound of linear elasticity. It agrees with the derivation via a different route by Ponte Castan ˜eda (1989); see also Ogden (1978) for previous work on bounds. Generalization of the upper Voigt bound can be done in a similar way. Introducing a power-law type energy function into the inequalities (5.4) and (5.8), a hierarchy of bounds on elastic tensors can be obtained Proc. R. Soc. A (2006)
Bounds in finite elasticity and thermoelasticity
1177
in the same manner as in the infinitesimal strain theory (Hazanov 1999). In the case of more complex energy functions (such as those described by Ogden (1972)), it is more advantageous to work directly with the energy functions. Here, it is worth mentioning that the main assumption involved in the derivation of (5.4), (5.5), (5.8) and (5.9) is that minimum principles described in §3 hold for any kinematically admissible displacement. In reality, inequalities (3.3), (3.7), (3.14) and (3.17) provide a global minimizer criteria only if the strain energy function is convex, which is just a mathematical assumption. For materials described by non-convex strain-energy functions, bounds (5.4)–(5.9) hold only locally in the range of validity of inequalities (3.3)–(3.17), which places a limitation on their use. If the inversion of the constitutive relation (2.3) is not unique, a different complementary variational principle such as the one proposed by Gao (1999) might be of better use. (b ) Linear versus nonlinear bounding problems Effective properties of a random composite in linear thermoelasticity were recently considered by Du & Ostoja-Starzewski (2005); a scale-dependent hierarchy of bounds on the thermal expansion coefficient was derived in the way analogous to the one considered above. Now, contrary to the linear elasticity case, the behaviour of random nonlinear composites subjected to temperature changes is not governed by a thermal expansion coefficient aij alone but by a number of temperature dependant constants. Moreover, the only materials undergoing finite strains and temperature changes relative to equilibrium state are biological soft tissues and rubber-like materials (Holzapfel 2000), whose thermomechanical behaviour is almost entirely based on an entropy concept. Total stress in these materials is caused by a change in entropy with deformation, while internal energy does not change with deformation at all. Therefore, the behaviour of composites in finite thermoelasticity is expected to be very different from the linear elasticity one. The central aim of this section is to compare the above-derived hierarchies in nonlinear elasticity with the results in the linear elasticity theory. Consider an isothermal elasticity problem. In linear elasticity, the volume average of strain energy of the body can be fully determined from surface data if one of the ‘canonical’ boundary conditions (4.5) or (4.6) are applied. One of the main consequences is that, for a macroscopically isotropic composite with isotropic phases, the hierarchy on effective infinitesimal strain-energy function can be separated into volumetric and deviatoric parts, which cannot be done in nonlinear elasticity. Indeed, the isotropic strain-energy function for small deformation elasticity can be put in the following form: 1 j Z ke2 C m3ij0 3ij0 ; 2
ð5:14Þ
where k is the bulk modulus, m is the shear modulus, eZ3kk is volume strain and 3 0 ij Z 3ij Kð1=3Þ3kk dij is deviatoric part of the strain field. The strain-energy form (5.14) plays an important role in proving the existence and uniqueness of solution in linear elastostatics. The first and second terms of (5.14) are independently variable and represent volumetric and deviatoric parts of j, Proc. R. Soc. A (2006)
1178
Z. F. Khisaeva and M. Ostoja-Starzewski
respectively. Equation (5.14) can be rewritten as 1 1 j ZK pe C sij0 3ij0 : 2 2
ð5:15Þ
Here, pZKð1=3Þskk is the mean pressure and sij0 Z sij Kskk dij is the deviatoric part of the stress field. Since the deviatoric and volumetric stress and strain tensors are mutually orthogonal, it follows from the Hill condition that ð 1 1 1 1 1 j dV ZK pe C sij0 3ij0 ZK pe C sij0 3ij0 : ð5:16Þ V V 2 2 2 2 Hence, the average strain energy under uniform volumetric or isochoric boundary conditions is fully determined by the corresponding volumetric or isochoric stress and strain field on the boundary. It is important to notice that, defining an apparent bulk modulus as K ZK p= e and an apparent shear K1 modulus as M Z ð1=2Þ sij0 3ij0 , one can estimate bounds on effective volumetric and isochoric responses. Thus, applying uniform volumetric strain on the boundary after ensemble averaging we get k Z hK idD % hK idd % hK idd0 % hK id1 ;
for 1! d 0 ! d! D;
ð5:17Þ
whereas uniform isochoric kinematic boundary condition leads to m Z hM idD % hM idd % hM idd0 % hM id1 ;
for 1! d 0 ! d! D:
ð5:18Þ
The reciprocal expression for the lower bound on the effective response can be obtained from the complementary energy function, which gives t t t t 1 1 1 1 1 % % % ; for 1! d 0 ! d! D; ð5:19Þ Z k K D K d K d0 K 1 t t t t 1 1 1 1 1 % % % ; Z m M D M d M d0 M 1
for 1! d 0 ! d! D:
ð5:20Þ
In the nonlinear elasticity of compressible rubber-like materials, the strainenergy function j can be also split into a volumetric part jvol and isochoric part jiso (Holzapfel 2000). However, the application of purely volumetric boundary condition does not give a zero isochoric contribution to the average strain energy and vice versa. The reason for this is the complex nature of the strain-energy function, which in general cannot be expressed as a product of the nominal stress and deformation gradient just as it was done in (5.19). In the potential (3.12), the only term independent of deformation is the purely ^ thermal contribution TðTÞ. This term does not depend on the type of mechanical loading and, therefore, is the same under restricted and unrestricted boundary conditions: ^ ^ Tðu; q0 Þ Z Tðu; q0 Þ r : ð5:21Þ Hence, the purely thermal contribution can be ignored in hierarchies (5.5) and (5.9). Proc. R. Soc. A (2006)
Bounds in finite elasticity and thermoelasticity
1179
As pointed out in §3b, the energetic contribution e0 to the free energy function (3.12) is equal to zero in the case of a purely entropic theory, in which the thermoelastic bounding problem becomes identical to the purely elastic one. In the most general situation, e0 is assumed to be a function of the volume ratio e0 ðui;j ÞZ e0 ðJ Þ (Chadwick & Creasy 1984). Thus, in contrast to the linear elasticity theory, the hierarchy (5.5) cannot be separated into purely mechanical and thermal parts and one has to consider the first two terms of j in (3.12) jointly. 6. Conclusions In this paper, the variational principles of Lee & Shield (1980) have been extended to the case of finite thermoelastostatics. It was shown that the uniform displacement and traction boundary conditions can be used to obtain, respectively, the upper and lower bounds on the effective energy functions of a composite at finite deformations under isothermal and non-isothermal loading. The approach developed can be used to estimate the scale dependence (i.e. mesoscale bounds) on effective response of random microstructures in finite elasticity and thermoelasticity and, hence, the size of the RVE without unnecessary and unwieldy inversion of constitutive relations. Such results will be reported in the subsequent publication. Anonymous referees’ comments are gratefully appreciated. The work reported herein has been made possible through support by the Werner Graupe International Fellowship, the NSERC, and the Canada Research Chairs program.
References Chadwick, P. & Creasy, C. F. M. 1984 Modified entropic elasticity of rubberlike materials. J. Mech. Phys. Solids 32, 337–357. (doi:10.1016/0022-5096(84)90018-8) Costanzo, F., Gray, G. L. & Andia, P. C. 2005 On the definition of effective stress and deformation gradient for use in MD: Hill’s macro-homogeneity and the virial theorem. Int. J. Eng. Sci. 43, 533–555. (doi:10.1016/j.ijengsci.2004.12.002) de Veubeke, B. F. 1972 A new variational principle for finite elastic displacement. Int. J. Eng. Sci. 10, 745–763. (doi:10.1016/0020-7225(72)90079-1) Du, X. & Ostoja-Starzewski, M. 2005 Mesoscale bounds on effective thermal expansion of random composites. Sixth Int. Cong. Therm. Stresses, Vienna, Austria, pp. 755–758. Gao, D. Y. 1999 General analytic solution and complementary variational principles for large deformation nonsmooth mechanics. Meccanica 34, 169–198. Hazanov, S. 1999 On apparent properties of nonlinear heterogeneous bodies smaller than the representative volume. Acta Mech. 134, 123–134. (doi:10.1007/BF01312651) Hill, R. 1972 On constitutive macro-variables for heterogeneous solids at finite strain. Proc. R. Soc. A. 326, 131–147. Holzapfel, G. A. 2000 Nonlinear solid mechanics. A continuum approach for engineering, pp. 305–369. New York: Wiley. Holzapfel, G. A., Gasser, T. C. & Ogden, R. W. 2000 A new constitutive framework for arterial wall mechanics and a comparative study of material models. J. Elast. 61, 1–48. (doi:10.1023/ A:1010835316564) Huet, C. 1990 Application of variational concepts to size effects in elastic heterogeneous bodies. J. Mech. Phys. Solids 38, 813–841. (doi:10.1016/0022-5096(90)90041-2) Proc. R. Soc. A (2006)
1180
Z. F. Khisaeva and M. Ostoja-Starzewski
Jiang, M., Ostoja-Starzewski, M. & Jasiuk, I. 2001 Scale-dependent bounds on effective elastoplastic response of random composites. J. Mech. Phys. Solids 49(3), 655–673. (doi:10. 1016/S0022-5096(00)00034-X) Lee, S. J. & Shield, R. T. 1980 Variational principles in finite elastostatics. Z. Angew. Math. Phys. 31, 437–453. (doi:10.1007/BF01590856) Lo¨hnert, S. & Wriggers, P. 2003 Homogenisation of microheterogeneous materials considering interfacial delamination at finite strains. Tech. Mech. 23, 167–177. Nemat-Nasser, S. 1999 Averaging theorems in finite deformation plasticity. Mech. Mat. 31, 493–523. (doi:10.1016/S0167-6636(98)00073-8) Nemat-Nasser, S. & Hori, M. 1999 Micromechanics: overall properties of heterogeneous materials, pp. 27–35. Amsterdam: Elsevier. Ogden, R. W. 1972 Large deformation isotropic elasticity: on the correlation of theory and experiment for compressible rubberlike solids. Proc. R. Soc. A 328, 567–583. Ogden, R. W. 1978 Extremum principles in nonlinear elasticity and their application to composites—I. Int. J. Solids Struct. 14, 265–282. (doi:10.1016/0020-7683(78)90037-9) Ogden, R. W. 1984 Non-linear elastic deformations, pp. 312–321. New York: Halsted Press. Ostoja-Starzewski, M. 1999 Microstructural disorder, mesoscale finite elements, and macroscopic response. Proc. R. Soc. A 455, 3189–3199. (doi:10.1098/rspa.1999.0446) Ostoja-Starzewski, M. 2001 Mechanics of random materials: stochastics, scale effects, and computation. In Mechanics of random and multiscale microstructures, vol. 430 (ed. D. Jeulin & M. Ostoja-Starzewski) CISM Courses and Lectures, pp. 93–161. Wien-New York: Springer. Ostoja-Starzewski, M. & Wang, C. 1989 Linear elasticity of planar Delaunay networks: random field characterization of effective moduli. Acta Mech. 80, 61–80. (doi:10.1007/BF01178180) Ostoja-Starzewski, M. & Castro, J. 2003 Random formation, inelastic response, and scale effects in paper. Phil. Trans. R. Soc. A 361, 965–986. (doi:10.1098/rsta.2003.1175) Ponte Castan ˜eda, P. 1989 The overall constitutive behavior of nonlinearly elastic composites. Proc. R. Soc. A 422, 147–171. Sab, K. 1992 On the homogenization of random materials. Eur. J. Mech. A/Solids 11, 585–607. Zubov, L. M. 1971 Variational principles of the nonlinear theory of elasticity. J. Appl. Math. Phys. (ZAMP) 35, 369–373.
Proc. R. Soc. A (2006)