Abstract

Surfaces and interfaces can significantly influence the overall response of a solid body. Their behavior is well described by continuum theories that endow the surface and interface with their own energetic structures. Such theories are becoming increasingly important when modeling the response of structures at the nanoscale. The objectives of this review are as follows. The first is to summarize the key contributions in the literature. The second is to unify a select subset of these contributions using a systematic and thermodynamically consistent procedure to derive the governing equations. Contributions from the bulk and the lower-dimensional surface, interface, and curve are accounted for. The governing equations describe the fully nonlinear response (geometric and material). Expressions for the energy and entropy flux vectors, and the admissible constraints on the temperature field, all subject to the restriction of non-negative dissipation, are explored. A particular emphasis is placed on the structure of these relations at the interface. A weak formulation of the governing equations is then presented that serves as the basis for their approximation using the finite element method. Various forms for a Helmholtz energy that describes the fully coupled thermomechanical response of the system are given. They include the contribution from surface tension. The vast majority of the literature on surface elasticity is framed in the infinitesimal deformation setting. The finite deformation stress measures are, thus, linearized and the structure of the resulting stresses discussed. The final objective is to elucidate the theory using a series of numerical example problems.

Introduction

The surface of a solid body typically exhibits properties that differ from those of the encased bulk. These differences, caused by processes such as surface oxidation, ageing, coating, atomic rearrangement, and the termination of atomic bonds, are present in comparatively thin boundary layers. Similarly, interfaces within the bulk can be viewed as two-sided internal surfaces. The mechanical and thermal properties of the interface can also differ significantly from the surrounding bulk. Surface and interface effects are especially significant for nanomaterials due to their large surface-to-volume ratio. These effects were appreciated already in the early 1900s by Pawlow [1], for example, who predicted the decrease in the melting temperature of a particle as its size reduces.

There exist two principal approaches to study the thermodynamics of surfaces and interfaces:

  • the zero-thickness layer or Gibbs (geometrical) method wherein a mathematical surface with zero thickness is introduced to capture excess quantities on the surface [2]

  • the finite-thickness layer method that dates back to van der Waals wherein a layer of finite thickness can be imagined instead of the interface

The reader is referred to Guggenheim [3] for further details and a comparison of the two approaches.

Following Gibbsean thermodynamics, various models have been proposed to endow the surface and interface with their own distinct properties, e.g., Adam [4], Shuttleworth [5], Herring [6], Bikerman [7], Orowan [8], and Cahn [9,10]. The physical chemistry of surfaces and interfaces for liquids and solids is studied in Refs. [4,11–15], among others.

The thermodynamic fundamentals of surface science were reviewed by Rusanov in Refs. [16,17]. Müller and Saul [18] presented a review on the importance of stress and strain effects on surface physics. The role of stress at solid surfaces was critically examined by Ibach [19]. Leo and Sekerka [20] investigated the equilibrium conditions for interfaces using a variational approach wherein the excess energy associated with the interface is allowed to depend on both the deformation of the interface and the crystallographic normal to the interface. Cammarata et al. [21,22,23,24] highlighted the surface and interface stress effects on thin films and nanoscaled structures. Fischer et al. [25] studied the role of surface energy and surface stress in phase-transforming nanoparticles and reported on the thermodynamics of a moving surface.

Gutman [26] recognized an inconsistency between the notion of the surface stress defined by, e.g., Shuttleworth, and that of the continuum elasticity theory. Bottomley and Ogino [27] examined an alternative to the Shuttleworth formulation of surface stress in solids (see also Ref. [28]). Kramer and Weissmüller [29] criticized Refs. [26–28] and clarified various misleading statements in the literature. Fischer et al. [25] circumvented such issues by a careful handling of the actual and reference configurations. Furthermore, Fischer et al. utilized the concept of configurational forces [30,31] as a convenient framework for distinguishing between surface energy, surface tension, and surface stress.

A widely adopted model, proposed by Gurtin and Murdoch [32,33], gives the surface its own tensorial stress measures (see, e.g., Refs. [22,34–36], for applications in nanomaterials). Murdoch [37], Gurtin and Struthers [38], and Gurtin et al. [39] extended this approach to consider interfaces within a solid. An important extension of the Gurtin and Murdoch model to account for the flexural resistance of the surface was developed by Steigmann and Ogden [40] and further studied in Refs. [41,42]. Moeckel [43] followed a different approach to Gurtin and Murdoch [32] for a moving interface within a thermomechanical solid. A master-balance equation is derived and applied to various key fields on the interface (see also Ref. [44]). The resulting equations are then restricted using the entropy principles [45] and constitutive relations for material interfaces derived. An alternative approach to develop general governing equations for the interface is to integrate the known equations for the bulk over the thickness of the interfacial layers (see, e.g., Ref. [46]). Daher and Maugin [47] used the method of virtual power [48,49] to derive the governing equations for an interface within a thermomechanical solid. This flexible approach is well suited to describing complex phenomena. For example, Daher and Maugin [50] applied the method to a model of deformable semiconductors with interfaces (see also Ref. [51]). Murdoch [52] addressed various aspects of surface modeling. Šilhavý [53] proved the existence of equilibrium of a two-phase state with an elastic solid bulk and deformation dependent interfacial energy. Park et al. [54], Park and Klein [55] developed an alternative continuum framework based on the surface Cauchy–Born model, an extension of the classical Cauchy–Born model, to include surface stresses (see also Refs. [56,57]).

The effect of surface energetics for ellipsoidal inclusions and the size-dependent elastic state of embedded inhomogeneities was investigated by Sharma et al. [58], Sharma and Ganti [59], and Sharma and Wheeler [60]. They utilized the classical formulation of Eshelby [61,62] for embedded inclusions and modified it by incorporating surface energies. Duan et al. [63] extended the Eshelby formalism for inclusion/inhomogeneity problems to the nanoscale. Effective mechanical and thermal properties of heterogeneous materials containing nanoinhomogeneities based on the generalized Eshelby formalism is investigated in Refs. [63–65]; see also related works Refs. [66–72].

Zöllner et al. [73] recently reported that to explore the biomechanical interaction during tissue expansion, one could model skin growth using a boundary energy. The thermomechanical behavior of low-dimensional systems was reviewed by Sun [74] together with a theoretical analysis elaborating on existing approximations in continuum mechanics and quantum computations. Johnson [75] studied the thermodynamics of a coherent interface separating two nonhydrostatically stressed crystals.

The objective of this presentation is to derive the equations governing the fully nonlinear coupled transient thermomechanical response of a body composed of a bulk, intersected by an interface, and encased by a surface, all of which are assumed to be energetic. Furthermore, an energetic curve that accounts for the interaction between the interface and the surface is also present. The importance of capturing such an interaction when modeling ionic nanowires was recently observed in Ref. [76].

A particular emphasis is placed here on the thermomechanical response of the interface. The governing equations for the various parts of the body are obtained from the balances of several key properties, namely linear and angular momentum, energy, and entropy. The corresponding constitutive relations then arise as thermodynamic restrictions on the dissipation. The diffusion of mass or chemical species is not accounted for in this contribution (see Ref. [77] for the case of species diffusion in thermomechanical solids with energetic surfaces).

Interfaces can be classified as follows. A material interface is one that does not move independently of the surrounding bulk material. An energetic interface is understood here to imply that the interface possesses its own thermomechanical structure in the form of an internal energy, entropy, constitutive relations, and dissipation. Such an interface is termed a thermodynamical singular surface by Daher and Maugin [47]. A thermal interface is defined as one that allows for heat conduction along the interface but possesses no energetic structure. Finally, a standard interface does not allow heat conduction along the interface and possesses no energetic structure. Daher and Maugin [47] term such an interface a free singular surface.

An interface can be classified further according to its thermal properties as follows:

  • Thermally perfect interface: The jump in the temperature and in the normal heat flux across the interface is zero.

  • Generic imperfect interface: In this general case neither the jump in the temperature nor the jump in the normal heat flux across the interface need be zero. A thermal interface is an example of a generic imperfect interface. Özdemir et al. [78], for example, developed a thermomechanical cohesive zone model for generic imperfect interfaces. The following two models are specializations of the generic imperfect interface model (see Refs. [79–81] and references therein).

    • (i)

      Weakly conducting (Kapitza) interface: This type of imperfect interface is modeled using Kapitza's concept of thermal resistance.1 The model allows for a temperature jump across the interface. The normal heat flux is, however, continuous across the interface. The Kapitza model for thermal interfaces has been widely investigated (see, e.g., Ref. [83] and references therein).

    • (ii)

      Highly conducting (HC) interface: This type of imperfect interface models the temperature as continuous across the interface, while allowing a jump in the normal component of the heat flux. Yvonnet et al. [84] have used the HC model to numerically investigate the effective conductivity of nanocomposites containing interfaces. For nonenergetic interfaces, a jump in the heat flux across the interface is only possible if there is an associated heat flux along the interface (see, e.g., Ref. [85]). This restriction does not hold for energetic interfaces [86].

Mosler and Scheider [87] recently described the admissible assumptions for the motion of the interface for a class of damage-type cohesive models. In a similar spirit, the relation of the interface temperature to that in the surrounding bulk is discussed here in detail. In particular, it is shown that the interface temperature is not, in general, equal to the average of the temperature across the interface.

Steinmann and Häsner [88] considered an energetic interface (with zero dissipation across the interface) within a thermomechanical solid subject to the assumption of infinitesimal deformation. Steinmann [89] presented extensive details on energetic surfaces using both the deformational and configurational frameworks of continuum mechanics. That investigation did not consider thermal contributions, interfaces, or numerical aspects. Javili and Steinmann [90,91,92,93] extended the work to account for the thermomechanics of energetic surfaces and the implementation of the model using the finite element method. The current review includes an extension of these works to account for the role of the interface and curve. Related works by Yvonnet et al. [83,84] focus on stationary heat conduction across thermal interfaces.

The focus of this work is on energetic elasticity as compared to entropic elasticity. For entropic elastic materials, e.g., polymers, the temperature increases due to stretching. On the contrary, energetic elastic materials, e.g., metals, cool down when stretched. The thermomechanical coupling, whether for energetic or entropic elasticity, is termed, henceforth, the Gough–Joule effect. The Gough–Joule effect is often used for entropic elastic materials as it was originally realized for rubbers (see Gough [94] and Joule [95]). It is noteworthy, that the Gough–Joule effect is not solely due to the deformation, but it is rather the rate of a deformation that induces a change in the temperature field.

In summary, the key objectives and contributions of this work are as follow:

  • To systematically derive the equations governing the response of a thermomechanical body, intersected by a material interface, and enclosed by a surface, undergoing finite deformations coupled to transient heat conduction. The procedure used to derive the equations is general and could be applied to a range of complex phenomena.

  • To account for the interaction between the interface and the surface.

  • To derive the thermodynamically consistent constitutive relations, dissipation inequalities, and temperature evolution equations in the various parts of the body (bulk, surface, interface, and curve).

  • To briefly present the weak formulation of the governing equations for the general case of a thermomechanical solid possessing an energetic surface, interface, and curve. Various other models are shown to arise as restrictions of this general case.

  • To elaborate various aspects of material (constitutive) modeling and discuss the admissible range of material parameters.

  • To elucidate the theory with numerical examples based on a finite element approximation of the governing equations.

This manuscript is organized as follows. The notation and certain key concepts are introduced briefly in Sec. 2. The equations governing the response of the various parts of the body are derived from fundamental balance principles in Sec. 3. A thermodynamic framework is then utilized in Sec. 4 to determine the form of the thermodynamically consistent constitutive relations. The nature of the coupling between the primary fields (i.e., the displacement and the temperature) and the various parts of the body are made clear. The dissipation inequality for the interface is analyzed in detail to show the possible relations between the interface temperature and that of the surrounding bulk. The weak formulation of the governing equations is then given in Sec. 5. Various important problems, which result as restrictions of this general weak formulation, are highlighted. Aspects of material modeling are elaborated in Sec. 6. A surface Helmholtz energy is introduced and the relationship between surface material parameters is discussed. The admissibility of surface negative material parameters is investigated. A series of numerical example problems, based on a finite element approximation of the weak formulation, is presented in Sec. 7 to elucidate the theory. Section 8 concludes this work.

Preliminaries

The purpose of this section is to summarize certain key concepts in nonlinear continuum mechanics and to introduce the notation adopted here. Detailed expositions on nonlinear continuum mechanics can be found in Refs. [96–101], among others. For further details concerning the continuum mechanics of deformable surfaces and interfaces the reader is referred to Refs. [38,43,47] and the references therein.

Notation and Definitions

Direct notation is adopted throughout. Occasional use is made of index notation, the summation convention for repeated indices being implied. The three-dimensional Euclidean space is denoted E3. The scalar product of two vectors a and b is denoted a·b = [a]i[b]i. The scalar product of two second-order tensors A and B is denoted A:B=[A]ij[B]ij. The composition of two second-order tensors A and B, denoted A·B, is a second-order tensor with components [A·B]ij = [A]im[B]mj. The vector product of two vectors a and b is denoted a × b with [a × b]k = [ε]ijk[a]i[b]j, where ε denotes the third-order permutation (Levi–Civita) tensor. Analogously, the vector product of a vector a and a second-order tensor B is denoted a × B with [a × B]kl = [ε]ijk[a]i[B]jl. The nonstandard products of a fourth-order tensor C and a vector b are defined by [b·¯C]ikl=[C]ijkl[b]j and [C·¯b]ijl=[C]ijkl[b]k. Other nonstandard products of a fourth-order tensor C and a vector b are defined by [b·¯C]ijl=[C]ijkl[b]k and [C·¯b]ikl=[C]ijkl[b]j. The nonstandard product of a fourth-order tensor C and a second-order tensor A is defined by [C:¯A]ik=[C]ijkl[A]jl. The nonstandard minor transposes of a fourth-order tensor C are defined by [tC]ijkl=[C]jikland[Ct]ijkl=[C]ijlk. The action of a second-order tensor A on a vector a is given by [A·a]i = [A]ij[a]j. The tensor product of two vectors a and b is a second-order tensor D = ab with [D]ij = [a]i[b]j. The tensor product of two second-order tensors A and B is a fourth-order tensor D = AB with [D]ijkl = [A]ij[B]kl. The two nonstandard tensor products of two second-order tensors A and B are the fourth-order tensors [A¯B]ijkl=[A]ik[B]jl and [A¯B]ijkl=[A]il[B]jk.

Quantities or operators corresponding to the bulk, surface, interface, and curve are denoted as {•}, {}, {¯}, and {˜}, respectively, unless specified otherwise. Note, {} denotes a surface-quantity that is not necessarily tangent to the surface. Similarly, a bulk quantity {•} on the surface implies the evaluation of {•} on the surface and is not, in general, equivalent to the corresponding surface quantity {}.

The expression the various parts of the body implies the bulk, the surface, the interface, and the curve.

Key definitions are given in Tables 1 and 2. For the derivations thereof and extensive technical details see classical manuscripts on differential geometry [102,103,104], among others. Basic concepts and terminologies of surfaces and curves can be found in the fundamentals of differential geometry on surfaces and curves briefly reviewed in Appendix  A.

Table 1

Summary of the notation used to denote key quantities and operators in the bulk, on the surface and interface, and on the curve. Quantities and operators corresponding to the material and spatial configurations are distinguished by a semicolon.

BulkSurfaceInterfaceCurve
DomainB0 ; BtS0 ; StI0 ; ItC0 ; Ct
 NormalN ; nN¯, N¯ ; n¯, n¯N˜c, N˜i, N˜s, N˜s± ; n˜c, n˜i, n˜s, n˜s±
 TangentN˜s, N ; n˜s, nN˜i ; n˜iN˜ ; n˜
Unit tensorI ; iI ; iI¯ ; i¯I˜ ; i˜
Differential elementdV ; dvdA ; dadA ; dadL ; dl
Oriented elementdA ; dadA ; dadL ; dl
Control regions
B
0 ;
B
t
S
0 ;
S
t
I0 ; It
C
0 ;
C
t
CurvatureC¯ ; c¯C¯ ; c¯C˜ ; c˜
Covariant basisG1, G2, G3 ; g1, g2, g3G1, G2 ; g1, g2G¯1, G¯2 ; g¯1, g¯2G˜1 ; g˜1
Contravariant basisG1, G2, G3 ; g1, g2, g3G1, G2 ; g1, g2G¯1, G¯2 ; g¯1, g¯2G˜1 ; g˜1
PlacementX ; xX ; xX¯ ; x¯X˜ ; x˜
Tangent mapF ; fF ; fF¯ ; f¯F˜ ; f˜
Normal mapCof F ; cof fCofF;coffCof¯F¯;cof¯f¯Cof˜F˜;cof˜f˜
JacobianJ ; jJ ; jJ¯ ; j¯J˜ ; j˜
VelocityV ; vV ; vV¯ ; v¯V˜ ; v˜
DivergenceDiv ; divDiv ; divDiv¯ ; div¯Div˜ ; div˜
GradientGrad ; gradGrad ; gradGrad¯ ; grad¯Grad˜ ; grad˜
DeterminantDet ; detDet ; detDet¯ ; det¯Det˜ ; det˜
CofactorCof ; cofCof ; cofCof¯ ; cof¯Cof˜ ; cof˜
BulkSurfaceInterfaceCurve
DomainB0 ; BtS0 ; StI0 ; ItC0 ; Ct
 NormalN ; nN¯, N¯ ; n¯, n¯N˜c, N˜i, N˜s, N˜s± ; n˜c, n˜i, n˜s, n˜s±
 TangentN˜s, N ; n˜s, nN˜i ; n˜iN˜ ; n˜
Unit tensorI ; iI ; iI¯ ; i¯I˜ ; i˜
Differential elementdV ; dvdA ; dadA ; dadL ; dl
Oriented elementdA ; dadA ; dadL ; dl
Control regions
B
0 ;
B
t
S
0 ;
S
t
I0 ; It
C
0 ;
C
t
CurvatureC¯ ; c¯C¯ ; c¯C˜ ; c˜
Covariant basisG1, G2, G3 ; g1, g2, g3G1, G2 ; g1, g2G¯1, G¯2 ; g¯1, g¯2G˜1 ; g˜1
Contravariant basisG1, G2, G3 ; g1, g2, g3G1, G2 ; g1, g2G¯1, G¯2 ; g¯1, g¯2G˜1 ; g˜1
PlacementX ; xX ; xX¯ ; x¯X˜ ; x˜
Tangent mapF ; fF ; fF¯ ; f¯F˜ ; f˜
Normal mapCof F ; cof fCofF;coffCof¯F¯;cof¯f¯Cof˜F˜;cof˜f˜
JacobianJ ; jJ ; jJ¯ ; j¯J˜ ; j˜
VelocityV ; vV ; vV¯ ; v¯V˜ ; v˜
DivergenceDiv ; divDiv ; divDiv¯ ; div¯Div˜ ; div˜
GradientGrad ; gradGrad ; gradGrad¯ ; grad¯Grad˜ ; grad˜
DeterminantDet ; detDet ; detDet¯ ; det¯Det˜ ; det˜
CofactorCof ; cofCof ; cofCof¯ ; cof¯Cof˜ ; cof˜
Table 2

Summary of the notation used to denote key quantities and operators in the bulk, on the surface and interface, and on the curve corresponding to the material and spatial configurations

Material configurationSpatial configuration
I:=G1G1+G2G2+G3G3i:=g1g1+g2g2+g3g3
Grad{}:={}/XDiv{}:=Grad{}:Igrad{}:={}/xdiv{}:=grad{}:i
Det{}:=[{}·G1]·[[{}·G2]×[{}·G3]]/[G1·[G2×G3]]det{}:=[{}·g1]·[[{}·g2]×[{}·g3]]/[g1·[g2×g3]]
Cof{}:=Det{}{}-tdv=JdVda=CofF·dAcof{}:=det{}{}-tdV=jdvdA=coff·da
I:=G1G1+G2G2=I-NNi:=g1g1+g2g2=i-nn
Grad{}:=Grad{}·IDiv{}:=Grad{}:Igrad{}:=grad{}·idiv{}:=grad{}:i
Det{}:=|[{}·G1]×[{}·G2]|/|G1×G2|det{}:=|[{}·g1]×[{}·g2]|/|g1×g2|
Cof{}:=Det{}{}-tda=JdAdl=CofF·dLcof{}:=det{}{}-tdA=jdadL=coff·dl
I¯:=G¯1G¯1+G¯2G¯2=I-N¯N¯i¯:=g¯1g¯1+g¯2g¯2=i-n¯n¯
Grad¯{¯}:=Grad{¯}·I¯Div¯{¯}:=Grad¯{¯}:I¯grad¯{¯}:=grad{¯}·i¯div¯{¯}:=grad¯{¯}:i¯
Det¯{¯}:=|[{¯}·G¯1]×[{¯}·G¯2]|/|G¯1×G¯2|det¯{¯}:=|[{¯}·g¯1]×[{¯}·g¯2]|/|g¯1×g¯2|
Cof¯{¯}:=Det¯{¯}{¯}-tda=J¯dAdl=Cof¯F¯·dLcof¯{¯}:=det¯{¯}{¯}-tdA=j¯dadL=cof¯f¯·dl
I˜:=G1˜G˜1=N˜N˜i˜:=g1˜g˜1=n˜n˜
Grad˜{˜}:=Grad{˜}·I˜Div˜{˜}:=Grad˜{˜}:I˜grad˜{˜}:=grad{˜}·i˜div˜{˜}:=grad˜{˜}:i˜
Det˜{˜}:=|{˜}·G˜1|/|G˜1|det˜{˜}:=|{˜}·g˜1|/|g˜1|
Cof˜{˜}:=Det˜{˜}{˜}-tdl=J˜dLn˜=Cof˜F˜·N˜cof˜{˜}:=det˜{˜}{˜}-tdL=j˜dlN˜=cof˜f˜·n˜
Material configurationSpatial configuration
I:=G1G1+G2G2+G3G3i:=g1g1+g2g2+g3g3
Grad{}:={}/XDiv{}:=Grad{}:Igrad{}:={}/xdiv{}:=grad{}:i
Det{}:=[{}·G1]·[[{}·G2]×[{}·G3]]/[G1·[G2×G3]]det{}:=[{}·g1]·[[{}·g2]×[{}·g3]]/[g1·[g2×g3]]
Cof{}:=Det{}{}-tdv=JdVda=CofF·dAcof{}:=det{}{}-tdV=jdvdA=coff·da
I:=G1G1+G2G2=I-NNi:=g1g1+g2g2=i-nn
Grad{}:=Grad{}·IDiv{}:=Grad{}:Igrad{}:=grad{}·idiv{}:=grad{}:i
Det{}:=|[{}·G1]×[{}·G2]|/|G1×G2|det{}:=|[{}·g1]×[{}·g2]|/|g1×g2|
Cof{}:=Det{}{}-tda=JdAdl=CofF·dLcof{}:=det{}{}-tdA=jdadL=coff·dl
I¯:=G¯1G¯1+G¯2G¯2=I-N¯N¯i¯:=g¯1g¯1+g¯2g¯2=i-n¯n¯
Grad¯{¯}:=Grad{¯}·I¯Div¯{¯}:=Grad¯{¯}:I¯grad¯{¯}:=grad{¯}·i¯div¯{¯}:=grad¯{¯}:i¯
Det¯{¯}:=|[{¯}·G¯1]×[{¯}·G¯2]|/|G¯1×G¯2|det¯{¯}:=|[{¯}·g¯1]×[{¯}·g¯2]|/|g¯1×g¯2|
Cof¯{¯}:=Det¯{¯}{¯}-tda=J¯dAdl=Cof¯F¯·dLcof¯{¯}:=det¯{¯}{¯}-tdA=j¯dadL=cof¯f¯·dl
I˜:=G1˜G˜1=N˜N˜i˜:=g1˜g˜1=n˜n˜
Grad˜{˜}:=Grad{˜}·I˜Div˜{˜}:=Grad˜{˜}:I˜grad˜{˜}:=grad{˜}·i˜div˜{˜}:=grad˜{˜}:i˜
Det˜{˜}:=|{˜}·G˜1|/|G˜1|det˜{˜}:=|{˜}·g˜1|/|g˜1|
Cof˜{˜}:=Det˜{˜}{˜}-tdl=J˜dLn˜=Cof˜F˜·N˜cof˜{˜}:=det˜{˜}{˜}-tdL=j˜dlN˜=cof˜f˜·n˜

Kinematics

Consider a continuum body that takes the open set B0R3 at the time t = 0, as depicted in Fig. 1. The boundary of the body B0 is denoted by S0 := ∂B0. The body B0 is partitioned into two disjoint subdomains, denoted by the open sets B0+ and B0-, by a two-sided interface I0. The interface I0 is an open set; thus, B0=B0+I0B0-. The boundary of the interface, a two-sided curve, is denoted as C0 := I0. In a similar fashion to the interface, the curve C0 partitions the surface S0 into two open sets S0+ and S0- such that S0+S0-=∅⁣ and S0 = S0+C0S0-.

Fig. 1
The domains B0, S0, I0, and C0 and the various unit normals
Fig. 1
The domains B0, S0, I0, and C0 and the various unit normals
Close modal

Remark. The boundary S0 is closed and is assumed to be smooth. For the more general case where the boundary S0 is nonsmooth, the curve C0 can be understood as a geometrical entity. For the restricted case considered here where the boundary is smooth, the curve C0 accounts for the interaction between the interface and the surface and should be understood as a physical entity. The extension of this work to nonsmooth boundaries is straightforward. Nevertheless, it involves introducing additional geometric concepts and further notation while providing little additional insight into the fundamental concepts.

The bulk is defined by B0 := B0+B0- and is the reference placement of material particles labeled XB0. The two sides of the interface I0 are denoted I0+:=B0+I0 and I0-:=B0-I0, with I0=I0+I0-. Material particles on the interface are labeled X¯I0 and, by definition, X¯:=X|I0. The outward unit normals to I0+ and I0- are denoted N¯+ and N¯-, respectively, with N¯+(X¯) = -N¯-(X¯). The unit normal to I0 is denoted N¯(X¯):=N-¯(X¯).

The surface is defined by S0:=S0+S0- with outward unit normal N. Material particles on the surface are labeled XS0 and, by definition, X=X|I0. The two sides of the curve C0 are denoted C0+:=I0+ and C0-:=I0-, with C0=C0+C0-. Material particles on the curve C0 are labeled X˜C0, where X˜:=X|C0. The outward unit normals to C0+ and C0-, tangential to I0 and identical at a point X˜, are denoted N˜i. The unit normals to C0+ and C0-, tangential to S0, are denoted N˜s+ and N˜s-, respectively, with N˜s+(X˜)=-N˜s-(X˜). The unit normal to C0, tangential to S0, is defined by N˜s(X˜):=N˜s-(X˜). The unit normals to C0 in the sense of the Frénet–Serret formulae and identical at a point X˜ are denoted N˜c. Note that N˜c does not necessarily coincide with either N˜i or N˜s.

Let T = [0, T] R+ denote the time domain. A motion of the reference placement for a time tT is denoted by the orientation-preserving map φ: B0×TR3. The current placement of the bulk associated with the motion φ is denoted Bt=ϕ(B0(X),t) with particles designated as x = φ (X, t) Bt (see Fig. 2).2

Fig. 2
The material and spatial configurations of a continuum body and the associated motions, deformation gradients, and velocity measures in the various parts of the body
Fig. 2
The material and spatial configurations of a continuum body and the associated motions, deformation gradients, and velocity measures in the various parts of the body
Close modal

The restriction of the motion φ to the surface S0 is denoted ϕ. The current placement of the surface is denoted St=ϕ(S0(X),t) with particles x=ϕ(X,t)St. It is assumed that particles on the surface of the body S0 constitute the surface for all times tT and ϕ=ϕ|S0, consequently x=x|St. This assumption, referred to as kinematic slavery by Steinmann and Häsner [88], mimics coherence in interfaces.

The restriction of the motion φ to the positive and negative sides of the interface I0 is denoted ϕ¯+ and ϕ¯-, respectively. The placements of the positive and negative sides of the interface at time t are denoted It+=ϕ¯+(I0+(X¯),t) and It-=ϕ¯-(I0-(X¯),t), respectively, with particles x¯+=ϕ¯+(X¯,t)It+ and x¯-=ϕ¯-(X¯,t)It-. The material interface is assumed to be coherent: The motion is not necessarily smooth across the interface, however, it is continuous across the interface, smooth away from the interface, and smooth up to the interface from either side [105]. Thus, ϕ¯+=ϕ¯-=ϕ|I0 and, consequently, x¯+=x¯-=x|It. The restriction of the motion φ to the interface is defined by ϕ¯:=ϕ|I0 with particles designated as x¯=ϕ¯(X¯,t)It.

Remark. Given that this work focuses on coherent material interfaces it may seem superfluous to distinguish between It+ and It-. However, coherence concerns only the continuity of the motion across the interface and jumps in other fields, e.g., temperature, are permitted.□

Following from the aforementioned assumptions, the restriction of the motion φ to the curve C0 at the junction of the interface and the surface is denoted ϕ˜. The current placement of the curve is denoted Ct=ϕ˜(C0(X˜),t) with particles x˜=ϕ˜(X˜,t)Ct. Finally, ϕ˜=ϕ|C0 and x˜=x|Ct=x¯|Ct=x^|Ct.

For the normal vectors to the surface, interface, and curve in the spatial configuration the same convention as in the material configuration is used, however, with lowercase n instead of uppercase N.

Deformation Gradients

The deformation gradient F:TB0TBt, that is the (invertible) linear tangent map between spatial and material line elements dxTBt (tangent space to Bt) and dXTB0, and the associated (material) velocity V are defined by
F(X,t):=Gradϕ(X,t)andV:=Dtϕ(X,t)
Similarly, the various deformation gradients for the remainder of the body are given by
F:TS0TSt,F(X,t):=Gradϕ(X,t),F¯:TI0TIt,F¯(X¯,t):=Grad¯ϕ¯(X¯,t),F˜:TC0TCt,F˜(X˜,t):=Grad˜ϕ˜(X˜,t)
The surface, interface, and curve deformation gradients (F,F,¯andF˜) are noninvertible linear tangent maps between the respective spatial and material elements (see Table 1 for further information). Although these deformation gradients are noninvertible, due to rank deficiency, they each possess an inverse in the following generalized sense:
F·F-1=iandF-1·F=I,F¯·F¯-1=i¯andF¯-1·F¯=I¯,F˜·F˜-1=i˜andF˜-1·F˜=I˜
The volume elements in the bulk, the area elements on the surface, and interface and the line elements on the curve in the material and spatial configurations can be mapped into each other by means of the Jacobian determinants of their associated deformation gradients and their inverses in the various parts of the body as
dv=JdVwithJ:=DetF,da=JdAwithJ:=DetF,da=J¯dAwithJ¯:=Det¯F¯,dl=J˜dLwithJ˜:=Det˜F˜,dV=jdvwithj:=detF-1,dA=jdawithj:=detF-1,dA=j¯dawithj¯:=det¯F¯-1,dL=j˜dlwithj˜:=det˜F˜-1

Control Regions

The equations governing the coupled response of the various parts of the body are obtained from the balances of linear and angular momentum, energy, and entropy performed over a control region. These balances are then localized at arbitrary points in the bulk B0, on the surface S0, on the interface I0, or on the curve C0.

The canonical control region is defined as one that possibly has as part of its boundary the surface S0 and is denoted

B0B0
with boundary
B0
(see Fig. 3). The orientable external surface of the control region is defined by
S0ext:=B0S0
while the interior surface is
S0int:=B0\S0ext
. The outward unit normals to
S0int
and
S0ext
are denoted M and N, respectively. The surface of the control region is defined by
S0:=S0extS0int
. The boundary of
S0ext
, a curve, is defined by
L0:=S0ext
. The unit normal to the curve
L0
is denoted N and is tangent to the surface
S0ext
.

Fig. 3
The bulk B0 and the canonical control region B0 and its representation in the bulk, and on the surface, interface, and curve
Fig. 3
The bulk B0 and the canonical control region B0 and its representation in the bulk, and on the surface, interface, and curve
Close modal

The control region

B0
can include a segment of the interface I0 defined by
I0:=B0I0
. Thus,
B0
can be partitioned into two subdomains
B0+
and
B0-
. The interface I0 has two sides
I0+:=I0B0+
and
I0-:=I0B0-
with I0=I0+I0-.

The boundary of the interface I0 is a curve denoted as

C0
⁠. The, possibly empty, intersection of the curves
C0
and C0 is denoted as
C0ext:=C0C0
. The interior boundary of the interface I0 is
C0int:=C0\C0ext
. The outward unit normal to the curve
C0int
is denoted M¯ and is tangent to the interface I0. The unit normal to the curve
C0ext
, tangent to the interface I0, is denoted N˜i. The unit normal to the curve
C0ext
, in the sense of Frénet–Serret formulae, is denoted N˜c. The unit normal to the curve
C0ext
, tangent to the surface
S0
, is denoted N˜s. The unit normal to the boundary of the curve
C0ext
, tangent to the curve
C0ext
is denoted N˜. The curve
C0ext
has two sides
+C0ext:=C0extI0+
and
-C0ext:=C0extI0-
with
C0ext=+C0ext-C0ext
. Analogously,
+C0int:=C0intI0+
and
-C0int:=C0intI0-
with
C0int=+C0int-C0int
.

Key Definitions and Identities

Various key definitions and identities that are required in the remainder of the presentation are now introduced. Several other useful relations are recorded in Appendix  B.

Jump and Average Relations

The following identity relates the jump of the product of two quantities (scalars or tensors, arbitrarily denoted a and b) to the products of their jumps and averages:
ab=a{{b}}+{{b}}a
(1)
where the operator denotes either scalar multiplication or a contraction. The average and the jump of a quantity over the interface I0 is defined as
{{{¯}}}:=12[{¯}|I0++{¯}|I0]and{¯}:={¯}|I0+{¯}|I0

Integral Relations

The extended form of the divergence theorem in the material configuration for the various parts of the body is now given. The (bulk) divergence theorem relates the material divergence of a quantity over the control volume
B0
entirely within the bulk, possibly containing an interface I0, to the flux of the quantity over the boundary
B0
, and the jump of the flux in the quantity over the interface I0 (see, e.g., Ref. [96] for further details). For a tensor field {}
B0Div{}dV=B0{}·MdA-I0[[{}]]·N¯dA
(2)
Similarly, the corresponding surface, interface, and curve divergence theorems for tensorial quantities on the surface {}, interface {¯}, and curve {˜} are, respectively, given by
S0extDiv{}dA=L0{}·NdL-S0extC{}·NdA-C0ext[[{}]]·Ns˜dLwhereC:=-DivN
(3)
I0Div¯{¯}dA=C0int{¯}·M¯dL+C0ext{¯}·N˜idL-I0C¯{¯}·N¯dAwhereC¯:=-Div¯N¯
(4)
C0extDiv˜{˜}dL=C0ext{˜}·N˜-C0extC˜{˜}·Nc˜dLwhereC˜:=-Div˜N˜c
(5)

Piola Identity and Piola Transform

The local balance equations expressed in spatial quantities can be obtained from the material versions using the standard (bulk) Piola identity
Div(CofF)=0
(6)
The surface, interface, and curve divergence theorems together with their corresponding Nanson's formulae result in the following surface, interface, and curve Piola identities:
Div(CofF)=Jcn,Div¯(Cof¯F¯)=J¯c¯n¯,Div˜(Cof˜F˜)=J˜c˜n˜c
(7)
A bulk Cauchy stress, heat flux, or entropy, collectively denoted as a Cauchy quantity γ, transforms into its corresponding Piola quantity Γ via the (bulk) Piola transform
Γ=γ·CofF
(8)
Analogously, a surface, interface, and curve Cauchy quantity denoted as γ, γ¯, and γ˜, respectively, transform into their corresponding Piola quantities Γ, Γ¯, and Γ˜ via the Piola transforms
Γ=γ·CofF,Γ¯=γ¯·Cof¯F¯,Γ˜=γ˜·Cof˜F˜
(9)

Superficial and Tangential Tensor Fields

Material second-order tensors and vectors on the surface, interface, and curve can be, respectively, classified as superficial (in their tangent spaces) or tangential. Superficial material second-order tensors on the surface, interface, and curve, respectively, possess the orthogonality properties
{}·N=0,{¯}·N¯=0,{˜}·N˜c=0
If the arbitrary quantities in the preceding relation are vectors, they are termed tangential. Tangential material second-order tensors on the surface, interface, and curve, respectively, possess the orthogonality properties
{}·N=0andN·{}=0,{¯}·N¯=0andN¯·{¯}=0,{˜}·N˜c=0andN˜c·{˜}=0
or, alternatively,
I·{}·I={},I¯·{¯}·I¯={¯},I˜·{˜}·I˜={˜}

Spatial second-order tensors and vectors can be defined in the same fashion by replacing the material quantities with the spatial ones. Note that the set of all tangential second-order tensors is contained within the set of all superficial second-order tensors.

Governing Equations

The coupled governing equations for the thermomechanical problem in the various parts of the body are now derived using an arbitrary (canonical) control region

B0B0
⁠. Thereafter, the balances of momentum, energy, and entropy, i.e., the governing equations in integral form, are given. Finally, these integral balance expressions are localized at arbitrary points in the bulk B0, on the surface S0, or interface I0, or on the curve C0, thereby giving the local (strong) form of the governing equations.3

The integral form of the governing equations over the control region, in its most general form, consists of eight terms that account for the contributions from the various parts of the body; that is,
0(or0)=B0{}dV+S0ext{}dA+I0{¯}dA+C0ext{˜}dL+S0int{°}·MdA+L0{°}·NdL+C0int{°¯}·M¯dL+C0ext{°˜}·N˜

where {} denotes a source-like term and {°} is the corresponding flux. The corresponding integral expressions for the balance of linear and angular momentum are derived and, subsequently, localized in Sec. 3.1. The localization process is explained in detail. In Sec. 3.2 the integral and resulting local forms of the energy balance are given. Finally, in Sec. 3.3 the process is repeated for the entropy balance.

Mechanical Power

The global working, i.e., the mechanical power, in the material configuration is denoted
W0
and given by
W0=W0(V,V,V¯,V˜):=B0V·BpdV+S0extV·BpdA+I0V¯·B¯pdA+C0extV˜·B˜pdL+S0intV·[P·M]dA+L0V·[P·N]dL+C0intV¯·[P¯·M¯]dL+C0extV˜·[P˜·N˜]
(10)

The prescribed bulk (body) force per unit reference volume of

B0
is denoted Bp (N/m3) where the superscript p indicates a prescribed quantity. The prescribed tractions per unit reference area of the surface
S0ext
and the interface I0 are denoted Bp (N/m2) and B¯p (N/m2), respectively. The prescribed traction per unit reference length of the curve
C0ext
is denoted B˜p (N/m). As with the bulk force Bp, the prescribed tractions on the surface, interface, and curve can be understood as prescribed surface, interface, and curve forces, respectively. The bulk, surface, interface, and curve Piola stresses are denoted P, P, P¯, and P˜, respectively. The terms P·M,P·N,P¯·M¯, and P˜·N˜ denote the tractions acting on
S0int,L0,C0int,
and
C0ext
, respectively.

Remark. It assumed that the Cauchy theorem holds for the bulk, surface, interface, and curve Piola stresses. Furthermore, P, P¯, and P˜ are superficial tensor fields in their tangent spaces.□

Invoking arguments concerning the invariance of the working under superposed rigid-body motions (see Ref. [92] for details concerning the case of thermomechanical solids with surface energy only) yields the global balances of momentum and angular momentum
B0BpdV+S0extBpdA+I0B¯pdA+C0extB˜pdL+S0intP·MdA+L0P·NdLC0intP¯·M¯dL+C0extP˜·N˜=0
(11)
B0R×BpdV+S0extR×BpdA+I0R¯×B¯pdA+C0extR˜×B˜pdL+S0intR×[P·M]dA+L0R×[P·N]dL+C0intR¯×[P¯·M¯]dL+C0extR˜×[P˜·N˜]=0(12)
(12)

where R, R,R¯, and R˜ are vectors to the points in the bulk, surface, interface, and curve, respectively, relative to a fixed point in R3, e.g., the origin.

Bulk

The global statements (11) and (12) are now localized at arbitrary XB0 by considering a control region
B0
such that
B0
does not contain any portion of the surface or interface, i.e.,
B0S0=B0I0=∅⁣
. Under such restrictions the limits of the integrals in Eqs. (11) and (12) are
B0∅⁣,S0ext=∅⁣,I0=∅⁣,C0ext=∅⁣,S0int=B0∅⁣,L0=∅⁣,C0int=∅⁣,C0ext=∅⁣
From the arbitrariness of the control region
B0
, and taking into account the corresponding extended (bulk) divergence theorem (2) together with Eq. (B1)1, one obtains the familiar local balances of linear and angular momentum in the bulk B0
DivP+Bp=0
(13)
F·Pt=P·Ft
(14)

Surface

The global statements (11) and (12) are now localized at arbitrary XS0 by considering a control region
B0
where
B0
contains a portion of the surface, i.e.,
B0S0∅⁣
but does not contain any portion of the interface, i.e.,
B0I0=∅⁣
. Under such restrictions the limits of the integrals in Eqs. (11) and (12) are
B0∅⁣,S0ext∅⁣,I0=∅⁣,C0ext=∅⁣,S0int=B0/S0ext∅⁣,L0∅⁣,C0int=∅⁣,C0ext=∅⁣
The volume of the control region is then decreased to zero, i.e., the limit
B0∅⁣
is taken and; consequently,
S0int=S0ext
with M = -N and R = R. From the arbitrariness of the control region
B0
and taking into account the corresponding extended surface divergence theorem (3) together with Eq. (B1)2, one obtains the local balances of linear and angular momentum on the surface S0 as
DivP+BP-P·N=0
(15)
F·Pt=P·Ft
(16)

Interface

Following the general approach outlined for the bulk and the surface, the global statements (11) and (12) are now localized at arbitrary X¯I0. Thus, a control region
B0
is considered such that
B0
contains a portion of the interface, i.e.,
B0I0∅⁣
, but
B0
does not contain any portion of the surface, i.e.,
B0S0=∅⁣
. Under such restrictions the limits of the integrals in Eqs. (11) and (12) are
B0∅⁣,S0ext=∅⁣,I0∅⁣,C0ext=∅⁣,S0int=B0/I0∅⁣,L0=∅⁣,C0int=I0∅⁣,C0ext=∅⁣
The volume of the control region is then decreased to zero, i.e., the limit
B0∅⁣
is taken and, consequently,
S0int=I0
with M = N¯ on I0+, M = -N¯ on I0-, and R = R¯. Once again, from the arbitrariness of the control region
B0
and taking into account the corresponding extended interface divergence theorem (4) together with Eq. (B1)3, one obtains the local balance of linear and angular momentum on the interface I0 as
Div¯P¯+B¯p+[[P]]·N¯=0
(17)
F¯·P¯t=P¯·F¯t
(18)

Curve

Finally, the global statements (11) and (12) are localized at arbitrary X˜C0. Thus, a control region
S0ext
is considered such that
S0ext
contains a portion of the curve, i.e.,
S0extC0∅⁣
. Under such restrictions the limits of the integrals in Eqs. (11) and (12) are
B0=∅⁣,S0ext∅⁣,I0=∅⁣,C0ext∅⁣,S0int=S0ext∅⁣,L0∅⁣,C0int=C0ext∅⁣,C0ext∅⁣
The limit
S0ext∅⁣
is then taken such that
S0ext=C0ext
with N=N˜s on
+C0ext,N=-N˜s
on
-C0ext
and R=R˜. From the arbitrariness of the control region
S0ext
, and taking into account the corresponding extended divergence theorem (5) together with Eq. (B1)4, one obtains the local balance of linear and angular momentum on the curve C0 as
Div˜P˜+B˜p-P¯·N˜i+[[P]]·Ns˜=0
(19)
F˜·P˜t=P˜·F˜t
(20)

Remark. The local balance of momentum on the interface Eq. (17) relates the interface Piola stress to the jump of the bulk Piola stress across the interface. The curve Piola stress, however, is coupled not only to the jump of the surface Piola stress across the curve but also to a contribution from the interface Piola stress according to Eq. (19).□

Material and Spatial Formulations

In summary, the local balances of linear and angular momentum for the various parts of the body in material quantities read
DivP+B=0                  andF·Pt=P·FtinB0      whereB=Bp,DivP+B=0                  andF·Pt=P·FtinS0      whereB=Bp-P·N,Div¯P¯+B¯=0                  andF¯·P¯t=P¯·F¯tinI0      whereB¯=B¯p+[[P]]·N¯Div˜P˜+B˜=0                  andF˜·P˜t=P˜·F˜tinC0      whereB˜=B˜p-P¯·N˜i+[[P]]·N˜s
(21)
Invoking the balances of momentum in the various parts of the body (21), the expression (10) for the mechanical power can be reformulated in terms of the stresses (the proof is given in Appendix  C) as
W0=B0P:GradVdV+I0P¯:Grad¯V¯dA+S0extP:GradVdA+C0extP˜:Grad˜V˜dL
(22)
The (bulk) Cauchy stress σ and the surface, interface, and curve tangential Cauchy stresses, denoted σ, σ¯, and σ˜, respectively, are related to their corresponding Piola stresses via the Piola transforms (8) and (9). The working in Eq. (22) can, thus, be reformulated in terms of spatial quantities as
Wt=Btσ:gradvdv+Itσ¯:grad¯v¯da+Stextσ:gradvda+Ctextσ˜:grad˜v˜dl

where the spatial velocity fields in the bulk on the surface, interface, and curve are denoted v(x,t),v(x,t),v¯(x¯,t), and v˜(x˜,t), respectively.

Using the Piola identities (6) and (7), the local momentum balance Eq. (21) can be expressed in terms of spatial
divσ+b=0whereb=bp,divσ+b=0whereb=bp-σ·n,div¯σ¯+b¯=0whereb¯=b¯p+[[σ]]·n¯,div˜σ˜+b˜=0whereb˜=b˜p-σ¯·n˜i+[[σ]]·n˜sandσ=σtinBtandσ=σtinStandσ=σtinItandσ˜=σ˜tinCt
(23)

where the prescribed bulk force per unit current volume of Bt is denoted bp := jBp. In an identical fashion, the prescribed surface, interface, and curve forces are denoted bp:=jBp,b¯p:=j¯B¯pandb˜p:=j˜B˜p, respectively.

Thermal Power

In order to derive the internal energy balance equation one needs a relation for the global heating. The global heating, i.e., the global external thermal power, expressed in the material configuration is denoted
Q0
and takes the form
Q0:=B0QpdV+S0extQpdA+I0Q¯pdA+C0extQ˜pdL+S0int[-Q·M]dA+L0[-Q·N]dL+C0int[-Q¯·M¯]dL+C0ext[-Q˜·N˜]
(24)

The prescribed bulk (body) heat source per unit reference volume (and per unit time) in

B0
is denoted Qp (N/s m2). The prescribed heat sources per unit reference area (and per unit time) of the surface
S0ext
and the interface I0 are denoted Qp(N/s m) and Q¯p(N/s m), respectively. The prescribed heat source per unit length (and per unit time) of the curve
C0ext
is denoted Q˜p(N/s). The bulk, surface, interface, and curve Piola heat flux vectors are denoted Q, Q, Q¯, and Q˜, respectively. The terms Q·M, Q·N, Q¯·M¯, and Q˜·N˜ denote the heat fluxes on
S0int
,
L0
C0int
, and
C0ext
, respectively. It is assumed that the Cauchy theorem holds for the heat flux vectors in all parts of the body. Note that Q, Q¯, and Q˜ are tangential vectors fields.

The rate of change of total internal energy expressed in the material configuration
E
0 is given by
E0:=B0DtEdV+S0DtEdA+I0DtE¯dA+C0DtE˜dL

where the bulk internal energy per unit reference volume, the surface and interface internal energies per unit reference area, and the curve internal energy per unit reference length are denoted E (N/m2), E (N/m), E¯ (N/m), and E˜ (N), respectively.

The internal energy balance states that working and heating cause a change in the internal energy, that is,
W0+Q0=E0
(25)
The global statement (25) is now localized at arbitrary XB0, XS0, X¯I0, and X˜C0. The resulting local balances of energy for the various parts of the body expressed in material quantities are
P:GradV-DivQ+Q=DtEinB0whereQ=Qp,P:GradV-DivQ+Q=DtEinS0whereQ=Qp+Q·N,P¯:Grad¯V¯-Div¯Q¯+Q¯=DtE¯inI0whereQ¯=Q¯p-[[Q]]·N,¯P˜:Grad˜V˜-Div˜Q˜+Q˜=DtE˜inC0whereQ˜=Q˜p+Q¯·N˜i-[[Q]]·N˜s
(26)

The Piola identities for the various parts of the body allow Eq. (26) to be expressed solely in terms of spatial quantities. The (bulk) Cauchy heat flux vector q and the surface, interface, and curve tangential Cauchy heat flux vectors, denoted q, q¯, and q˜, respectively, are related to their corresponding Piola heat flux vectors via the Piola transforms (8) and (9).

Thus, the local balances of energy for the various parts of the body in terms of spatial quantities are
σ:gradv-div(q+ev)+q=dteinBtwhereq=qp,σ:gradv-div(q+ev)+q=dteinStwhereq=qp+q·n,σ¯:grad¯v¯-div¯(q¯+e¯v¯)+q¯=dte¯inItwhereq¯=q¯p-[[q]]·n¯,σ˜:grad˜v˜-div˜(q˜+e˜v˜)+q˜=dte˜inCtwhereq˜=q˜p+q¯·n˜i-[[q]]·n˜s
(27)

where the prescribed heat source per unit current volume (and unit time) in Bt is denoted qp := jQp. The bulk internal energy per unit current volume is denoted e := jE. In an identical fashion to the bulk, the prescribed surface, interface, and curve heat sources are denoted qp:=jQp,q¯p:=j¯Q¯p, and q˜p:=j˜Q˜p respectively. Similarly, e:=jE,e¯:=j¯E¯,ande˜:=j˜E˜ denote the surface, interface, and curve internal energies, respectively. Recall that dt denotes the spatial time derivative. Thus, convective contributions of the type ev appear in the corresponding divergence theorems.

Entropy Power

The global entropy input rate in the material configuration, denoted
H
0, is given by
H0:=B0HpdV+S0extHpdA+I0H¯pdA+C0extH˜pdL+S0int[-H·M]dA+L0[-H·N]dL+C0int[-H¯·M¯]dL+C0ext[-H˜·N˜]
(28)

The prescribed bulk (body) entropy source per unit reference volume (and unit time) in

B0
is denoted Hp (N/sKm2). The prescribed entropy sources on the surface
S0ext
, and the interface I0 are denoted Hp (N/sKm) and H¯p (N/sKm), respectively. The prescribed entropy source on the curve
C0ext
is denoted H˜p (N/sK). In the global entropy input (28) the terms H·M, H·N, H¯·M¯, and H˜·N˜ denote the entropy fluxes on
S0int
,
L0
,
C0int
, and
C0ext
, respectively. The vectors H, H, H¯, and H˜ are the bulk, surface, interface, and curve Piola entropy flux vectors, respectively. The vectors H, H¯, and H˜ are tangential tensor fields in their tangent spaces.

The total entropy production rate in the material configuration, denoted
P0
, is defined by
P0:=B0ΠdV+S0extΠdA+I0Π¯dA+C0extΠ˜dL0

where II ≥ 0, Π ≥ 0, Π¯ ≥ 0, and Π˜ ≥ 0 are the entropy production rates per unit reference volume of the bulk, per unit reference area of the surface and interface, and per unit reference length of the curve, respectively, all constrained to be positive.

The rate of the change of total entropy in the material configuration, denoted
N
0, is defined by
N0:=B0DtΞdV+S0extDtΞdA+I0DtΞ¯dA+C0extDtΞ˜dL

where the bulk entropy per unit reference volume is denoted Ξ (N/Km2). The surface and interface entropies per unit reference area are denoted Ξ¯ (N/Km) and Ξ (N/Km), respectively. The curve entropy is denoted Ξ˜ (N/K).

The second law of thermodynamics imposes the physical restriction that the rate of increase of entropy in a body is not less than the total entropy supplied to the body or, equivalently, that the entropy increase is due to positive entropy production and entropy input, that is,
N0=P0+H0withN0H0sincePo0
(29)
The global statement of balance of entropy (29) is then localized at arbitrary XB0, XS0, X¯I0, and X˜C0. The resulting local balances of entropy for the various parts of the body in material quantities are
-DivH+Π+H=DtΞandΠ0inB0whereH=Hp,-DivH+Π+H=DtΞandΠ0inS0      whereH=Hp+H·N-Div¯H¯+Π¯+H¯=DtΞ¯andΠ¯0inI0            whereH¯=H¯p-[[H]]·N,¯-Div˜H˜+Π˜+H˜=DtΞ˜andΠ˜0inC0                        whereH˜=H˜p+H¯·Ni˜-[[H]]·Ns˜
(30)

The (bulk) Cauchy entropy flux vector h and the surface, interface, and curve tangential Cauchy entropy flux vectors, denoted h, h¯, and h˜, respectively, are related to their corresponding Piola entropy flux vectors via the Piola transforms (8) and (9).

Following the same procedures employed in the previous sections, the local balances of entropy for the various parts of the body in terms of spatial quantities are
-div(h+ξv)+π+h=dtξandπ0inBtwhereh=hp,-div(h+ξv)+π+h=dtξandπ0inStwhereh=hp+h·n,-div¯(h¯+ξ¯v¯)+π¯+h¯=dtξ¯andπ¯0inItwhereh¯=h¯p-[[h]]·n,¯-div˜(h˜+ξ¯v¯)+π˜+h˜=dtξ˜andπ˜0inCt      whereh˜=hp˜+h¯·n˜i-[[h]]·n˜s
(31)

where the bulk prescribed entropy source, entropy and entropy production per unit current volume in Bt are denoted hp:=jHp,ξ:=jΞ,andπ:=jΠ, respectively. In an identical fashion to the bulk, the prescribed surface, interface, and curve entropy sources are denoted hp:=jHp,h¯p:=j¯H¯p,andh˜p:=j˜H˜p, respectively. Similarly, ξ:=jΞ,ξ¯:=j¯Ξ¯, and ξ˜:=j˜Ξ˜ denote the surface, interface, and curve entropies, respectively. Finally, the prescribed surface, interface, and curve entropy productions are denoted π:=jΠ,π¯:=j¯Π¯,andπ˜:=j˜Π˜, respectively.

To proceed, the commonly made assumption that the entropy fluxes are the corresponding heat fluxes divided by the corresponding absolute temperatures is used, that is:
H:=QΘ-1,H:=QΘ-1,H¯:=Q¯Θ¯-1,H˜:=Q˜Θ˜-1,Hp:=QpΘ-1,Hp:=QpΘ-1,H¯p:=Q¯pΘ¯-1,H˜p:=Q˜pΘ˜-1
(32)

The variables Θ > 0, Θ > 0, Θ¯ > 0, and Θ˜ > 0 denote the absolute temperatures in the bulk, on the surface, interface, and curve, respectively. Note that the assumption (32) is not true in general but holds for simple thermodynamic processes (see, e.g., Refs. [30,106,107]). The validity and necessity of this assumption should be explored using the procedure outlined in Müller [108] and later works.

The dissipation power for the various parts of the body are denoted
D,D
,
D¯
, and
D˜
and are defined by
D:=ΘΠ0,D:=ΘΠ0,D¯:=Θ¯Π¯0,D˜:=Θ˜Π˜0
Substituting these relations into the balances of entropy for the various parts of the body (30) yields the localized dissipation inequalities
D=ΘDtΞ+DivQ-Q·GradlnΘ-Qp0inB0,D=ΘDtΞ+DivQ-Q·GradlnΘ-[Qp+ΘΘQ·N]0inS0,D¯=Θ¯DtΞ¯+Div¯Q¯-Q¯·Grad¯lnΘ¯-[Q¯p-Θ¯[[QΘ]]·N¯]0inI0,D˜=Θ˜DtΞ˜+Div˜Q˜-Q˜·Grad˜lnΘ˜-[Q˜p+Θ˜Θ¯Q¯·N˜i-Θ˜[[QΘ]]·N˜s]0inC0
(33)
The localized dissipation inequalities (33) can be reformulated using the energy balance Eq. (26) as
D=ΘDtΞDtE+P:GradVQGradlnΘ0inB0,D^=Θ^DtΞ^DtE^+P^:Grad^V^Q^Grad^lnΘ+[1Θ^Θ]QN0inS0,D¯=Θ¯DtΞ¯DtE¯+P¯:Grad¯V¯Q¯Grad¯lnΘ¯+[Θ¯QΘQ]N¯0inI0,D˜=Θ˜DtΞ˜DtE˜+P˜:Grad˜V˜Q˜Grad˜lnΘ˜+[Θ˜Q^Θ^Q^]Ns˜+[1Θ˜Θ¯]Q¯Ni˜0inC0
(34)
Next, the Helmholtz energy, obtained as a Legendre transformation of the internal energy in terms of entropy and temperature, is introduced. Separate Helmholtz energies in the bulk, on the surface, interface, and curve, denoted Ψ, Ψ, Ψ¯, and Ψ˜, respectively, are permitted and are defined by
Ψ:=E-ΘΞ,Ψ:=E-ΘΞ,Ψ¯:=E¯-Θ¯Ξ¯,Ψ˜:=E˜-Θ˜Ξ˜
(35)
From the definitions of the Helmholtz energies, the extended Clausius–Duhem dissipation inequalities can be obtained as
D=P:GradVDtΨΞDtΘQGradlnΘ0inB0,D^=P^:Grad^V^DtΨ^Ξ^DtΘ^Q^Grad^lnΘ^+[1Θ^Θ]QN0inS0,D¯=P¯:Grad¯V¯DtΨ¯Ξ¯DtΘ¯Q¯Grad¯lnΘ¯+Q[Θ¯QΘQ]N¯0inI0,D˜=P˜:Grad˜V˜DtΨ˜Ξ˜DtΘ˜Q˜Grad˜lnΘ˜+[Θ˜Q^Θ^Q^]Ns˜+[1Θ˜Θ¯]Q¯Ni˜0inC0
(36)

Thermohyperelastic Helmholtz Energy

To exploit the dissipation inequalities further, a thermohyperelastic constitutive model is specified by selecting a particular list of arguments for the Helmholtz energies. For the case of thermohyperelastic behavior, the arguments of the Helmholtz energies are chosen as follows:
Ψ:=Ψ(F,Θ),Ψ:=Ψ(F,Θ,n),Ψ¯:=Ψ¯(F¯,Θ¯,n¯),Ψ˜:=Ψ˜(F˜,Θ˜,n˜)
(37)

Note that the surface and interface Helmholtz energies are allowed to depend on their corresponding normals in the spatial configuration to capture possible anisotropic behavior. The anisotropy of the curve, however, is modeled by dependence of the curve Helmholtz energy on its tangent in the spatial configuration. It should be emphasized that in deriving the balance equations in Sec. 3, no isotropy assumptions were made.

Constitutive Equations

Exploiting the dissipation inequalities in the various parts of the body (36) using a Coleman–Noll-like procedure yields the constitutive laws
P=ΨF,Ξ=-ΨΘinB0,P=ΨF-nSwithS:=ψn·i·cofF,Ξ=-ΨΘinS0,P¯=Ψ¯F¯-n¯S¯withS¯:=ψ¯n¯·i¯·cof¯F¯,Ξ¯=-Ψ¯Θ¯inI0,P˜=Ψ˜F˜-s˜N˜withs˜:=ψ˜n˜·[i-i˜],Ξ˜=-Ψ˜Θ˜inC0
(38)

where the Helmholtz energies in the spatial configuration, i.e., per unit deformed area for the surface and interface and per unit deformed length for the curve, are defined by ψ:=jΨ,ψ¯:=j¯Ψ¯, and ψ˜:=j˜Ψ˜, respectively. The quantities S, S¯, and s˜ denote the surface shear, interface shear, and curve shear, respectively (see Refs. [105,109]). Note that the terms nS, n¯S¯, and s˜N˜ map from the tangent spaces to the material surface, interface, and curve, to the normals to the spatial surface, interface, and curve, respectively.

Reduced Dissipation Inequalities

Inserting the constitutive relations derived in Sec. 4.1 in the dissipation inequalities (36), results in the following reduced dissipation inequalities:
D=QGrad lnΘ0inB0,D^=Q^Grad^lnΘ^D^+[1Θ^Θ]QND^0inS0,D¯=Q¯Grad¯lnΘ¯D¯+[Θ¯QΘQ]N¯D¯0inI0,D˜=Q˜Grad˜lnΘ˜D˜+[Θ˜Q^Θ^Q^]Ns˜+[1Θ˜Θ¯]Q¯Ni˜D˜0inC0
(39)

Note that the first terms in Eq. (39)2–4 represent the dissipation along the surface, interface and curve and are denoted

D
⁠,
D¯,
and
D˜
, respectively. In contrast, the remaining terms represent the dissipation across the surface, interface, and curve, and are denoted as
D
,
D¯
, and
D˜
, respectively.

Dissipation Inequalities in the Bulk and Along the Surface, Interface, and Curve

In order to partially ensure thermodynamic consistency by satisfying the dissipation inequalities along the surface, interface, and curve, and in the spirit of Fourier's law, the heat flux vectors are assumed to be of the form
Q=-JF-1·k·F-T·GradΘinB0,Q=-JF-1·k·F-T·GradΘinS0,Q¯=-J¯F¯-1·k¯·F¯-T·Grad¯Θ¯inI0,Q˜=-J˜F˜-1·k˜·F˜-T·Grad˜Θ˜inC0
(40)

where k, k, k¯, and k˜ denote the bulk, surface, interface, and curve positive semidefinite spatial thermal conductivity tensors, respectively.

Dissipation Inequalities Across the Surface, Interface, and Curve

Assuming that the dissipation inequalities in the bulk and along the surface, interface, and curve are satisfied, the further reduced dissipation inequalities across the surface, interface, and curve are given by
D=Θ[1Θ-1Θ]Q·N0,D¯=Θ¯[[[QΘ]]-[[Q]]Θ¯]·N¯0,D˜=Θ˜[[[QΘ]]-[[Q]]Θ˜]·N˜s+Θ˜[1Θ˜-1Θ¯]Q¯·N˜i0
(41)

The requirements for these inequalities to be satisfied are now discussed.

Surface.
The most obvious choice to sufficiently satisfy the further reduced dissipation inequality on the surface (41)1 is
Θ=Θ|S0
(42)

that is, to impose so-called thermal slavery on the surface. It should be emphasized that the thermal slavery condition on the surface is not an a priori assumption but arises, rather, as a natural consequence of the thermodynamics. Alternatively McBride et al. [77] propose a Robin-type boundary condition between the bulk and surface.

Interface.
The conditions to satisfy the further reduced dissipation inequality on the interface Eq. (41)2 are now presented. Equation (41)2 can be reformulated4 using either of the two identities
[[QΘ]]=[[Q]]{{1Θ}}+{{Q}}[[1Θ]],[[Q]]Θ¯=[[QΘ]]{{Θ}}Θ¯+{{QΘ}}[[Θ]]Θ¯
(43)
as
D¯=Θ¯[{{1Θ}}-1Θ¯][[Q]]·N¯            +Θ¯[[1Θ]]{{Q}}·N¯      0   (using Eq. (43)1),D¯=[Θ¯-{{Θ}}][[QΘ]]·N¯            -[[Θ]]{{QΘ}}·N¯      0   (using Eq. (43)2)
(44)

The further reduced dissipation inequality on the interface (41)2 is satisfied if either of the two equivalent forms in Eq. (44) is satisfied.

By defining the coldness ϑ as the temperature inverse, i.e., ϑ:=1/Θ, Eq. (44) can be expressed in an alternative form as
-D¯=[ϑ¯-{{ϑ}}]-[[Q]]ϑ¯·N¯-[[ϑ]]{{Q}}ϑ¯·N¯0   (corresponding to Eq. (44)1),D¯=[Θ¯-{{Θ}}][[QΘ]]·N¯-[[Θ]]{{QΘ}}·N¯0   (corresponding to Eq. (44)2)
(45)
Coldness-based discussion.
The sufficient conditions to satisfy Eq. (45)1 are examined first. The inequality is satisfied if the following conditions hold:
[ϑ¯-{{ϑ}}][[Q]]·N¯0and[[ϑ]]{{Q}}·N¯0
(46)
A natural choice to satisfy the first relation (46)1 is a Fourier-type law for the jump of the heat flux across the interface, that is,
ϑ¯-{{ϑ}}=-α¯ϑ[[Q]]·N¯
(47)
where α¯ϑ ≥ 0 is a coldness sensitivity coefficient across the interface. To satisfy the second condition (46)2 a Fourier-type law for the average heat flux across the interface can be chosen, that is,
[[ϑ]]=r¯ϑ{{Q}}·N¯
(48)

where r¯ϑ ≥ 0 is a heat flux resistance coefficient across the interface.

Remark. For a thermal interface, the jump in the normal component of the heat flux is given by Eq. (26)3 as
[[Q]]·N¯=-Div¯Q¯
(49)

which is the thermal Young–Laplace equation on the interface. Equation (49) states that the normal jump of heat flux is equal to the negative of the divergence of the heat flux along the interface. In particular, in the absence of heat flux along the thermal interface the normal jump of the heat flux, i.e., the left-hand side of the Eq. (49) vanishes, i.e., the standard interface, and, therefore, Eq. (46)1 is trivially satisfied.□

Remark. A sufficient condition for a thermal interface to satisfy Eq. (46)1 is to choose the interface coldness equal to the average of the coldness across the interface. This conclusion shall be compared with Daher and Maugin [47] who viewed it as a necessity.□

Remark. Using the relation [[ϑ]]=-{{ϑ}}[[Θ]]/{{Θ}}, the condition (46)2 can be reformulated as [[Θ]]{{Q}}·N¯0 that can be satisfied by a Fourier-type law for the average heat flux across the interface [[Θ]]=-r¯Q{{Q}}·N¯ where r¯Q0 denotes the Kapitza heat flux resistance coefficient.□

Temperature-based discussion
The sufficient conditions to satisfy Eq. (45)2 are now examined. The inequality is satisfied if the following conditions hold:
[Θ¯-{{Θ}}][[QΘ]]·N¯0and[[Θ]]{{QΘ}}·N¯0
(50)
A natural choice to satisfy condition (50)1 is a Fourier-type law for the jump in the entropy flux across the interface, that is,
Θ¯-{{Θ}}=α¯Θ[[H]]·N¯
(51)
where α¯Θ0 is a temperature sensitivity coefficient across the interface. Likewise, to satisfy the condition (50)2 a Fourier-type law for the average entropy flux across the interface can be chosen, that is,
[[Θ]]=-r¯Θ{{H}}·N¯
(52)

where r¯Θ0 is the entropy flux resistance coefficient across the interface.

Remark. For a standard interface, the jump of the entropy flux; in contrast to the jump of the heat flux, does not necessarily vanish according to Eq. (30)3. That is, for a standard interface Eq. (50)1 is not trivially satisfied. Thus, a sufficient condition to satisfy Eq. (50)1 is to choose the interface temperature equal to the average of the temperature across the interface.

Remark. For a highly conducting interface the jump of the temperature across the interface vanishes, which corresponds to zero normal dissipation across the interface, i.e.,

D=0
⁠. For this case Θ¯:={{Θ}} or 1/Θ¯:={{1/Θ}}. Therefore, for a highly conducting interface, the interface temperature (resp. coldness) is trivially the average of the temperature (resp. coldness) across the interface sinceΘ|I0+=Θ|I0-=Θ.□

Remark. Using the relation [[Θ]]=-{{Θ}}[[ϑ]]/{{ϑ}}, the condition (50)2 can be reformulated as, [[ϑ]]{{H}}·N¯0, which can be satisfied by a Fourier-type law for the average entropy flux across the interface [[ϑ]]=r¯H{{H}}·N¯ where r¯H0 denotes an entropy flux resistance coefficient.□

Summary.
Thus, in summary, the sufficient conditions to satisfy the further reduced dissipation inequality on the interface (41)2 are
ϑ¯-{{ϑ}}=-α¯ϑ[[Q]]·N¯and([[ϑ]]=r¯ϑ{{Q}}·N¯or[[Θ]]=-r¯Q{{Q}}·N¯)
or
Θ¯-{{Θ}}=α¯Θ[[H]]·N¯and([[Θ]]=-r¯Θ{{H}}·N¯or[[ϑ]]=r¯H{{H}}·N¯)
Curve.
The further reduced dissipation inequality on the curve (41)3 has two parts. The first term involves jumps and formally resembles the further reduced dissipation inequality on the interface (41)2. The second term formally resembles the further reduced dissipation inequality on the surface (41)1. The sufficient condition to satisfy the further reduced dissipation inequality on the curve is to ensure that each of these two terms are non-negative. Analogous to the interface, we have two sets of conditions and one of them has to be satisfied to ensure that the first term in Eq. (41)3 is non-negative, that is,
1Θ˜={{1Θ}}and[[Θ]]=-r˜Q{{Q}}·N˜s
(53)
or
Θ˜={{Θ}}and[[1Θ]]=r˜H{{H}}·N˜s
(54)

where r˜Q0 and r˜H0 are heat flux and entropy flux resistance coefficients, respectively. Analogous to the surface, we can impose the thermal slavery on the curve, i.e., Θ˜=Θ¯|C0, as the most obvious way to guarantee that the second term in Eq. (41)3 is non-negative. A second novel option is to impose a constraint between the temperatures on the curve Θ˜ and interface Θ¯ and the heat flux Q¯·N˜i that resembles a Robin-type boundary condition between the curve and the interface.

Temperature Evolution Equations

The thermohyperelastic constitutive equations together with the local energy balance equations give the temperature evolution equations in the bulk, on the surface, interface, and curve as
cFDtΘ=-DivQ+Q+ΘPΘ:DtFinB0,cFDtΘ=-DivQ+Q+ΘPΘ:DtFinS0,c¯F¯DtΘ¯=-Div¯Q¯+Q¯+Θ¯P¯Θ¯:DtF¯inI0,c˜F˜DtΘ˜=-Div˜Q˜+Q˜+Θ˜P˜Θ˜:DtF˜inC0
(55)
whereby cF,cF, c¯F¯, and c˜F˜ denote the heat capacity coefficients in the bulk, on the surface, interface, and curve, respectively, defined as
cF:=-Θ2ΨΘΘ,cF:=-Θ2ΨΘΘ,c¯F¯:=-Θ¯2Ψ¯Θ¯Θ¯,c˜F˜:=-Θ˜2Ψ˜Θ˜Θ˜
(56)

The derivation of the interface temperature evolution Eq. (55)3 is given in Appendix  D.

Weak Formulation

In order to establish a principle of virtual work like statement and as a prerequisite for finite element method approximations (to be presented in a subsequent contribution), the mechanical weak form, i.e., the weak form of the linear momentum balance Eq. (21) and the thermal weak form, i.e., the weak form of the temperature evolution Eq. (56), are required.

Weak Formulation—Mechanical

To derive the mechanical weak form, the local linear momentum balance equations for the various parts of the body (21)1–4 are tested from left with vector-valued test functions

δϕH01(B0)
⁠,
δϕH01(S0)
,
δϕ¯H01(I0)
, and
δϕ˜H01(C0)
, respectively. The result is integrated over the corresponding domains in the material configuration to give the global weak form of the balance of linear momentum as

Find
ϕH1(B0)
,
ϕH1(S0)
,
ϕ¯H1(I0)
, and
ϕ˜H1(C0)
such that
B0δϕ·[DivP+Bp]dV+S0δϕ·[DivP+[Bp-P·N]]dA+I0δϕ¯·[Div¯P¯+[B¯p+[[P]]·N¯]]dA+C0δϕ˜·[Div˜P˜+[B˜p-P¯·N˜i+[[P]]·N˜s]]dL=0,δϕH01(B0),δϕH01(S0),δϕ¯H01(I0),δϕ˜H01(C0)
Using Eq. (B3)1–4 and the extended divergence theorems (2), the orthogonality properties of the Piola stress measures in the various parts of the body (which cause the integrals containing the curvature terms to vanish) and from the kinematic slavery condition
{{δϕ}}|I0=δϕ¯,{{δϕ}}|C0=δϕ˜,δϕ|S0=δϕ,δϕ¯|C0=δϕ˜
(57)
the weak form of the balance of linear momentum becomes
B0P:GradδϕdV+S0P:GradδϕdA+I0P¯:Grad¯δϕ¯dA+C0P˜:Grad˜δϕ˜dL-B0δϕ·BpdV-S0δϕ·BpdA-I0δϕ¯·B¯pdA-C0δϕ˜·B˜pdL=0(58)
(58)

where the test functions satisfy the conditions given in Eq. (57).

Weak Formulation—Thermal

In order to derive the thermal weak form we proceed formally. The local temperature evolution equations for the various parts of the body (55)1–4 are tested from left with scalar test functions

δΘH01(B0)
⁠,
δΘH01(S0)
,
δΘ¯H01(I0)
, and
δΘ˜H01(C0)
, respectively. The result is integrated over the corresponding domains in the material configuration to give the global weak form of the temperature evolution equation as:

Find
ΘH1(B0)
,
ΘH1(S0)
,
Θ¯H1(I0)
, and
Θ˜H1(C0)
such that
B0δΘ[cFDtΘ+DivQ-Qp-ΘP,Θ:DtF]dV+S0δΘ[cFDtΘ+DivQ-[Qp+Q·N]-ΘP,Θ:DtF]dA+I0δΘ¯[c¯F¯DtΘ¯+Div¯Q¯-[Q¯p-[[Q]]·N¯]-Θ¯P¯,Θ¯:DtF¯]dA+C0δΘ˜[c˜F˜DtΘ˜+Div˜Q˜-[Q˜p+Q¯·N˜i-[[Q]]·N˜s]-Θ˜P˜,Θ˜:DtF˜]dL=0,δΘH01(B0),δΘH01(S0),δΘ¯H01(I0),δΘ˜H01(C0)
(59)
Using Eq. (B2)1–4 and the extended divergence theorems given in Eqs. (2)–(5), the orthogonality properties of the heat flux vectors for the various parts of the body (which cause the integrals containing the curvatures to vanish), Eq. (1), and the assumption of thermal slavery for the surface and the curve, i.e., Θ=Θ|S0 and Θ˜=Θ¯|C0 so that
δΘ|S0=δΘ,δΘ¯|C0=δΘ˜
(60)
the global weak form of the temperature evolution equation becomes
-I0[[δΘ]]{{Q}}·N¯+[{{δΘ}}-δΘ¯][[Q]]·N¯dA-C0[[δΘ]]{{Q}}·N˜s+[{{δΘ}}-δΘ˜][[Q]]·N˜sdL=B0Q·GradδΘ-δΘcFDtΘ+δΘQp+δΘΘP,Θ:DtFdV      +S0Q·GradδΘ-δΘcFDtΘ+δΘQp+δΘΘP,Θ:DtFdA      +I0Q¯·Grad¯δΘ¯-δΘ¯c¯F¯DtΘ¯+δΘQ¯p+δΘΘP¯,Θ¯:DtF¯dA      +C0Q˜·Grad˜δΘ˜-δΘ˜c˜F˜DtΘ˜+δΘ˜Q˜p+δΘ˜Θ˜P˜,Θ˜:DtF˜dL
(61)

where the test functions satisfy the conditions given in Eq. (60).

Comparison With Several Other Models

The objective of this section is to briefly derive several models given in the literature by simplifying the weak forms (58) and (61). In particular, the focus is on the thermal weak form (61). The following cases are examined:

  • standard thermomechanical solids [110]

  • highly-conducting interfaces [84]

  • Kapitza interfaces [83]

  • thermomechanical surfaces [93]

  • thermomechanical HC interfaces [86]

  • generalized Kapitza interfaces [111]

Note that none of the aforementioned models account for the contributions from the curve and, therefore, all the corresponding integrals vanish identically.

Standard Thermomechanical Solids

For standard thermomechanical solids with neither energetic nor thermal interfaces, surfaces, or curves, the left-hand side of Eq. (61) and the integrals over the interface, surface, and curve vanish. Therefore, the thermal weak form reads
B0Q·GradδΘ-δΘcFDtΘ+δΘQp+δΘΘP,Θ:DtFdV=-S0δΘQpdA

The numerical implementation of finite strain thermoelasticity is well documented in the literature (see, e.g., Ref. [110] and references therein).

Highly Conducting Interfaces

Thermal solids with highly conducting interfaces do not allow for a jump in the temperature field across the interface, i.e., [[Θ]]=0 and [[δΘ]]=0. Thus, {{Θ}}=Θ¯ and {{δΘ}}=δΘ¯. Therefore, the stationary thermal weak form in the absence of (bulk) heat sources reads
B0Q·GradδΘdV+I0Q¯·Grad¯δΘ¯dA=-S0δΘQpdA
(62)

For details of the numerical implementation see, e.g., Yvonnet et al. [84] and the references therein.

Kapitza Interfaces

For the case of thermal solids containing Kapitza interfaces, the normal heat flux is continuous across the interface, i.e., [[Q]]·N¯=0, but a jump in the temperature field is admissible via the relation [[Θ]]=-r¯Q{{Q}}·N¯. Therefore, the stationary thermal weak form reads
B0Q·GradδΘdV-I01r¯Q[[δΘ]][[Θ]]dA=-B0δΘQpdV
(63)

The last term on the left-hand side vanishes for a standard interface; this assumption has been made by Yvonnet et al. [83].

Thermomechanical Surfaces

For thermomechanical solids with energetic surfaces, the left-hand side of Eq. (61) and the integrals over the interface and curve vanish. Thus, the thermal weak form reads
B0Q·GradδΘ-δΘcFDtΘ+δΘQp+δΘΘP,Θ:DtFdV+S0Q·GradδΘ-δΘcFDtΘ+δΘQp+δΘΘP,Θ:DtFdA=0

which is identical to the one given in Javili and Steinmann [93].

Thermomechanical Highly Conducting Interfaces

For thermomechanical solids with HC energetic interfaces, the left-hand side of Eq. (61) vanishes since [[Θ]]=0, [[δΘ]]=0, {{Θ}}=Θ¯, and {{δΘ}}=δΘ¯. Thus, the thermal weak form reads
B0Q·GradδΘ-δΘcFDtΘ+δΘQp+δΘΘP,Θ:DtFdV+S0δΘQpdA+I0Q¯·Grad¯δΘ¯-δΘ¯c¯F¯DtΘ¯+δΘ¯Q¯p+δΘ¯Θ¯P¯,Θ¯:DtF¯dA=0

This model, as compared to Eq. (62), accounts for thermomechanical coupling in the bulk and on the interface and is not limited to the stationary case; see Javili et al. [86] for further details.

Generalized Kapitza Interfaces

The Kapitza interface is termed generalized in the sense that the interface is mechanically energetic in contrast to classical Kapitza interfaces. For the case of thermomechanical solids containing generalized Kapitza interfaces, a jump in the temperature field is admissible via the relation [[Θ]]=-r¯Q{{Q}}·N¯ while [[Q]]·N¯=0. Therefore, the thermal weak form reads
B0Q·GradδΘ-δΘcFDtΘ+δΘQp+δΘΘP,Θ:DtFdV+S0δΘQpdA-I01r¯Q[[δΘ]][[Θ]]dA=0

This model, as compared to Eq. (63), not only endows the interface with its own mechanically energetic structure but also accounts for thermomechanical coupling in the bulk and is not limited to the stationary case; see Javili et al. [111] for further details.

Material Modeling

The objective of this section is to elaborate various aspects of material (constitutive) modeling. From the perspective of material modeling, the surface and the interface are essentially identical. Furthermore, the derivations on the surface and interface capture all the key features of the material model for the curve. Therefore, without any loss of generality, we shall focus the majority of the subsequent discussion on the material modeling of the bulk and the surface. However, in order to provide sufficient background material to explain the numerical examples in Sec. 7, we also briefly present details on the thermomechanical modeling of interfaces. The influence of the surface on the overall response of a structure at the nanoscale is one of the key applications for models of surface elasticity in the literature. We also restrict attention to thermoelastic materials. Inelastic behavior at the nanoscale is recently reported in Ref. [112].

The material modeling of bulk materials is a mature field with many standard references (see, e.g., Ref. [100]). This is not the case for the surface. Modeling the constitutive response of the surface surrounding the bulk introduces additional complexity for the following reasons:

  • The surface deformation gradient F is rank deficient. Thus, the derivations of the surface stress measures and their linearizations are more complicated.

  • The surface is a two-dimensional manifold embedded in three-dimensional space. Thus, the admissible range for the surface material parameters differs from those of the bulk (see Refs. [92,113] for further details).

  • Surface tension, which corresponds to a constant energy per unit area of the spatial configuration, introduces a nonstandard term into the surface Helmholtz energy. The surface Helmholtz energy for the strain-free configuration no longer vanishes. Hence, the classical (bulk) growth conditions [114,115] are no longer valid.

  • The notions of the surface energy, surface tension and surface stress are often used inappropriately in the literature (for a detailed discussion and clarification see Ref. [25] and references therein).

  • The definition of the surface elasticity theory in the seminal work of Gurtin and Murdoch [32] has a serious defect. This error was reported by the authors in Ref. [116] but has been overlooked by many others.

  • Gurtin and Murdoch used the letter S to denote the Piola stress and called it the Piola–Kirchhoff stress. The letter S is widely used in the literature to denote the (second) Piola–Kirchhoff stress.

For these reasons, there exists considerable confusion and inconsistency in the literature on surface elasticity (e.g., Refs. [25,26,29,117] aim to report and clarify such issues). In order to clarify the theory and to present a self-contained and consistent framework, a Helmholtz energy for the surface is introduced based on measures of the finite deformation, and the resulting stresses derived. The vast majority of the literature on surface elasticity theory considers the small-strain setting. Thus, the stress measures are linearized at the material configuration to obtain their small-strain counterparts. The result of the linearization of the finite deformation stress measures needs to be interpreted carefully. It is shown here that the linearization of the surface Piola stress P at the material configuration is different from that of the surface Piola–Kirchhoff stress S. Recall that for the bulk the linearizations of the Piola stress P and the Piola–Kirchhoff stress S at the material configuration are identical.

First a standard hyperelastic material model for the bulk is introduced and corresponding Piola and Piola–Kirchhoff stresses given. Thereafter, in Sec. 6.2, the derivations of the surface Piola and Piola–Kirchhoff stresses corresponding to a hyperelastic material model for the surface are performed. The surface material parameters and their admissible ranges for the linear elastic case are also presented. Furthermore, in order to elucidate the role of surface tension and the dependence of the surface operators upon the curvature within a simplified setting, an example involving the Young–Laplace equation is given. Finally, the hyperelastic material models are extended to the case of thermohyperelasticity in Sec. 6.3.

Remark. The form of the Helmholtz energy on the interface, and indeed in the bulk, can be obtained from fundamental reasoning or from atomistic modeling (see, e.g., Refs. [25,118]). Dingreville and Qu [119] developed a semianalytic method to compute the surface elastic properties of crystalline materials. Moreover, an interface energy can be constructed using the surface Cauchy–Born hypothesis [56]. In Ref. [76] the surface elastic parameters are extracted from ab initio calculations. Similar strategies can be employed to extract the interface parameters. Thus, in general, thermal and mechanical constants can be obtained using inverse parameter identification.

Hyperelastic Material in the Bulk

The bulk response is assumed to be hyperelastic. Specifically, a neo-Hookean-type Helmholtz energy Ψ of the following form is chosen:
Ψ(F)=12λln2J+12μ[F:F-3-2lnJ]withJ=DetF>0
(64)
where the Helmholtz energy is characterized by the two Lamé constants μ and λ. Alternatively, and in order to guarantee objectivity, the Helmholtz energy is expressed in terms of the right Cauchy–Green stretch tensor C:=Ft·F as
Ψ(C)=12λln2J+12μ[TrC-3-2lnJ]withJ=DetF=[DetC]12
where Tr {•} = {•}: I denotes the trace operator. The Piola stress P and Piola–Kirchhoff stress S follow from Eq. (38) as
P(F):=FΨ(F)=λlnJF-t+μ[F-F-t]
(65)
S(C):=2CΨ(C)=λlnJC-1+μ[I-C-1]
(66)

It is of particular interest to linearize these two stress measures at the material configuration in order to obtain their small strain counterparts. In doing so, recall that F − I = Grad u where u denotes the displacement vector and the infinitesimal strain is defined by ε := [Grad u]sym. The symmetric part of a second-order tensor {} can be obtained via {}sym=Isym:{} where Isym:=(1/2)[I¯I+I¯I] denotes the fourth-order symmetric identity tensor. The fourth-order volumetric identity tensor is defined by Ivol:=(1/3)II. Furthermore, {}vol:={}:Ivol represents the volumetric part of a second-order tensor {}.

The linearization of the Piola stress P in Eq. (65) at the material configuration is given by
LinP|F=I=P|F=I+A|F=I:[F-I]=3λɛvol+2μɛ
(67)
where the tangent A is
A:=FP=λ[F-tF-t+lnJD]+μ[I-D]
with
D:=FF-t=-F-t¯F-1andI:=FF=i¯I
Next, the Piola–Kirchhoff stress S given in Eq. (66) is linearized at the material configuration
LinS|C=I=S|C=I+12C|C=I:[C-I]=3λɛvol+2μɛ
(68)
where
C:=2CS=λ[12C-1C-1+lnJH]-μH
with
H:=CC-1=-12[C-1¯C-1+C-1¯C-1]
Comparing Eqs. (67) and (68), it is obvious that
LinP|F=I=LinS|c=I=:σLinorσLin=E:ɛ
(69)
with
E:=[3λIvol+2μIsym]

where σLin denotes the linearized stress tensor and E is the fourth-order tensor of elastic moduli.

It is common to decompose the linearized stress σLin and infinitesimal strain ε into their volumetric and deviatoric contributions as
σLin=σLinvol+σLindevandɛ=ɛvol+ɛdev
It follows from Eq. (69) that
σLinvol=3κεvol.andσLindev=2μεdev

where κ := λ + 2/3μ is the bulk compression modulus. For the sake of completeness, the elastic modulus and Poisson ratio are denoted as E and v, respectively. The relation between the various material parameters in the bulk are summarized in Table 3 (see, e.g., the standard Ref. [120] for further details).

Table 3

Classical relations between material parameters in the bulk where Δ:=E2+9λ2+2Eλ

λμ=GEνκ
λμλμμ(3λ+2μ)λ+μλ2(λ+μ)λ+23μ
λEλE-3λ+Δ4E2λE+λ+ΔE+3λ+Δ6
λνλλ(1-2ν)2νλ(1+ν)(1-2ν)ννλ(1+ν)3ν
λκλ32(κ-λ)9κ(κ-λ)3κ-λλ3κ-λκ
μEμ(E-2μ)3μ-EμEE-2μ2μμE3(3μ-E)
μν2μν1-2νμ2μ(1+ν)ν2μ(1+ν)3(1-2ν)
μκκ-23μμ9κμ3κ+μ3κ-2μ6κ+2μκ
EνEν(1+ν)(1-2ν)E2(1+ν)EνE3(1-2ν)
Eκ3κ(3κ-E)9κ-E3κE9κ-EE3κ-E6κκ
νκ3κν1+ν3κ(1-2ν)2(1+ν)3κ(1-2ν)νκ
λμ=GEνκ
λμλμμ(3λ+2μ)λ+μλ2(λ+μ)λ+23μ
λEλE-3λ+Δ4E2λE+λ+ΔE+3λ+Δ6
λνλλ(1-2ν)2νλ(1+ν)(1-2ν)ννλ(1+ν)3ν
λκλ32(κ-λ)9κ(κ-λ)3κ-λλ3κ-λκ
μEμ(E-2μ)3μ-EμEE-2μ2μμE3(3μ-E)
μν2μν1-2νμ2μ(1+ν)ν2μ(1+ν)3(1-2ν)
μκκ-23μμ9κμ3κ+μ3κ-2μ6κ+2μκ
EνEν(1+ν)(1-2ν)E2(1+ν)EνE3(1-2ν)
Eκ3κ(3κ-E)9κ-E3κE9κ-EE3κ-E6κκ
νκ3κν1+ν3κ(1-2ν)2(1+ν)3κ(1-2ν)νκ

Hyperelastic Material on the Surface

As in the bulk, we assume a hyperelastic response for the surface. The surface Helmholtz energy Ψ is chosen as follows [121]:
Ψ(F)=γJ+12λln2J+12μ[F:F-2-2lnJ]withJ=DetF>0
(70)
The surface Helmholtz energy consists of a surface tension part characterized by the surface tension γ, and a neo-Hookean part, formally similar to that of the bulk given, characterized by the two surface Lamé constants μ and λ. The surface tension part represents a constant Helmholtz energy per unit area of the spatial configuration and accounts for possible fluid-like behavior of the surface. The neo-Hookean part represents the elastic, solid-like response of the surface. Alternatively, the surface Helmholtz energy can be expressed in terms of the surface right Cauchy–Green stretch tensor C:=Ft·F as follows:
Ψ(C)=γJ+12λln2J+12μ[TrC-2-2lnJ]withJ=DetF=[DetC]12>0

where Tr{}={}:I denotes the surface trace operator.

The surface Piola stress P and the surface Piola–Kirchho stress S are, respectively, given by
P(F)=FΨ(F)=γJF-t+λlnJF-t+μ[F-F-t]
(71)
S(C):=2CΨ(C)=γJC-1+λlnJC-1+μ[I-C-1]
(72)

In an identical fashion to the bulk, these stress measures are linearized at the material configuration to obtain their small-strain counterparts. Recall that F-I=Gradu where u denotes the surface displacement vector and the infinitesimal surface strain is defined by ɛ:=[Gradu]sym The symmetric part of a surface second-order tensor {} can be obtained as {}sym=Isym:{} where Isym:=(1/2)[I¯I+I¯I] is the surface fourth-order symmetric identity tensor. The surface fourth-order volumetric, or rather spherical5 identity tensor is defined by Ivol:=12II. Furthermore, {}vol:={}:Ivol denotes the volumetric part of a surface second-order tensor {}.

The linearization of the surface Piola stress P given in Eq. (71) at the material configuration reads
LinP|F=I=P|F=I+A|F=I:[F-I]=γI+[γI¯I+2[λ+γ]Ivol+2[μ-γ]Isym]:Gradu=γI+γGradu+2[λ+γ]ɛvol+2[μ-γ]ɛ(73)
(73)
where
A:=FP=γJ[F-tF-t+D]+λ[F-tF-t+lnJD]+μ[I-D]
with
D:=FF-t=-F-t¯F-1+[i-i]¯[F-1·F-t]andI:=FF=i¯I
Next, the surface Piola–Kirchho stress S given in Eq. (72) is linearized at the material configuration
LinS|C=I=S|C=I+12C|C=I:[C-I]=γI+2[λ+γ]ɛvol+2[μ-γ]ɛ
(74)
where
C:=2CS=γJ[12C-1C-1+H]+λ[12C-1C-1+lnJH]-μH
with
H:=CC-1=-12[C-1¯C-1+C-1¯C-1]
Comparing the Eqs. (73) and (74), it is obvious that on the surface, and in contrast to the bulk, LinP|F=ILinS|C=I but rather
LinP|F=I=LinS|C=I+γGradu
Therefore, the two surface stress measures corresponding to the small-strain assumption are
PσLin:=LinP|F=I=γI+E:ɛ+γGradu(nonsymmetric)SσLin:=LinS|C=I=γI+E:ɛ(symmetric)
(75)
with
E:=[2λeffIvol+2μeffIsym]andλeff:=λ+γwhereμeff:=λ-γ

The mistake made by Gurtin and Murdoch [32] in their widely cited paper was to omit the surface tension in the definition of the effective quantities λeff and μeff. This error was rectified in an addendum to their original work [116]. Nonetheless, the error has been reproduced in numerous subsequent publications by many others.

Both of the linearized stress measures given in Eq. (75) are used in the literature. In the absence of surface tension they are obviously equivalent. In the presence of surface tension care must be taken to ensure that they are used in combination with their corresponding balance equation. In the authors' opinion the nonsymmetric version should be used in conjunction with Eq. (21)2.

Remark. Note that
Div(PσLin)=Div(SσLin)+γDiv(Gradu)

which shall be compared with Eq. (13) of Ref. [117] where the author explains geometrically the displacement-gradient term. Ru [117] argues that the controversial displacement-gradient term appears due to the fact that the infinitesimal elastic deformation is indeed an incremental deformation superposed on the initial (finite) deformation due to the surface tension. In this manuscript, the reference configuration is assumed to be strain-free as well as stress-free. The surface tension γ introduces a residual stress in the body and deforms it into a state from which the infinitesimal deformation is externally applied. In this sense, the final deformed configuration may be understood as a hybrid formulation combining the linearized deformation of the bulk and finite (second-order) deformation of the surface (see Ref. [117] for further details of this approach). Such justifications seem to be unavoidable when explaining Eq. (75), nevertheless, the mathematical derivations of this section show rigorously the nature of each term.□

For certain solids the influence of surface tension is negligible when compared to that of surface elasticity (see, e.g., Ref. [122]). Thus, in order to study the elastic effects of the surface as compared to the bulk, it is assumed, for the majority of the remainder of this work, that γ0. The role of surface tension will be explained further in the subsequent Young–Laplace example.

In analogy to the bulk, the symmetric linearized surface stress denoted as σLin=SσLin and infinitesimal surface strain ɛ can be decomposed into their volumetric and deviatoric contributions via
σLin=σLinvol+σLinvolandɛ=ɛvol+ɛdev
It follows from Eq. (75)2 that
σLinvol=2κɛvolandσLindev=2μɛdev

with κ:=λ+μ as the surface compression modulus. Note that the definition of the surface modulus κ follows in the spirit of κ in the bulk. This parameter has been defined differently in the literature by incorporating a factor of two (see, e.g., Refs. [58,123]). Nevertheless, it seems that the definition here is more stringent. For the sake of completeness, the surface elastic modulus and surface Poisson ratio are denoted as E and v, respectively. The relation between the different surface material parameters are summarized in Table 4 (see Javili and Steinmann [92] for further details). Note that the surface incompressibility limit, i.e., the upper bound for the surface Poisson ratio v, takes the value of 1 and not 12 as in the bulk. This can be justified geometrically.

Table 4

Relations between material parameters at the surface where Δ is defined by Δ := E2+4λ2

λμ=GEνκ
λμλμ4μ(λ+μ)λ+2μλλ+2μλ+μ
λEλE-2λ±Δ4E2λE±ΔE+2λ±Δ4
λνλλ(1-ν)2νλ(1-ν2)ννλ(1+ν)2ν
λκλκ-λ4κ(κ-λ)2κ-λλ2κ-λκ
μE2μ(E-2μ)4μ-EμE(E-2μ)2μμE4μ-E
μν2μν1-νμ2μ(1+ν)νμ(1+ν)1-ν
μκκ-μμ4μκμ+κκ-μκ+μκ
EνEν1-ν2E2(1+ν)EνE2(1-v)
Eκ2κ(2κ-E)4κ-EEκ4κ-EE2κ-E2κκ
νκ2κν1+νκ(1-ν)1+ν2κ(1-ν)νκ
λμ=GEνκ
λμλμ4μ(λ+μ)λ+2μλλ+2μλ+μ
λEλE-2λ±Δ4E2λE±ΔE+2λ±Δ4
λνλλ(1-ν)2νλ(1-ν2)ννλ(1+ν)2ν
λκλκ-λ4κ(κ-λ)2κ-λλ2κ-λκ
μE2μ(E-2μ)4μ-EμE(E-2μ)2μμE4μ-E
μν2μν1-νμ2μ(1+ν)νμ(1+ν)1-ν
μκκ-μμ4μκμ+κκ-μκ+μκ
EνEν1-ν2E2(1+ν)EνE2(1-v)
Eκ2κ(2κ-E)4κ-EEκ4κ-EE2κ-E2κκ
νκ2κν1+νκ(1-ν)1+ν2κ(1-ν)νκ

Finally, the admissible range for the surface material parameters is considered. This range can be determined by analyzing the well-posedness of the boundary value problem. The necessary and sufficient conditions for the loss of well-posedness of the boundary value problem governing linear elastic, homogeneous solids are well known (see, e.g., Refs. [125–134]). They are the loss of ellipticity of the governing equations and the boundary complementing condition, respectively.

Javili et al. [113] show that the sufficient conditions for the well-posedness of the boundary value problem governing a bulk surrounded by a surface are pointwise stability both in the bulk and on the surface. The results are summarized in Table 5, which shows that the strong ellipticity of the surface is necessary but not sufficient for the well-posedness. Shenoy [134] argues that, within the context of an atomistic model, the calculated surface elasticity tensor need not be pointwise stable as the “surface cannot exist independent of the bulk” and that it is the “total energy (bulk + surface)” that needs to “satisfy the positive definiteness condition,” where positive definiteness refers to pointwise stability. This is not entirely correct. The definition of pointwise stability is a local concept. An argument similar to that of Shenoy [134] is applicable when discussing the stability of the weak approximation to the governing equations, for example, as the basis for finite element computations (see Ref. [113] for further details).

Table 5

Summary of the conditions for the strong ellipticity and pointwise stability of the elasticity tensors in the bulk and on the surface in terms of the Lamé moduli

Strong ellipticityμ > 0 and λ + 2μ > 0μ>0 and λ+2μ>0
Pointwise stabilityμ > 0 and 3λ + 2μ > 0μ>0 and λ+μ>0
Strong ellipticityμ > 0 and λ + 2μ > 0μ>0 and λ+2μ>0
Pointwise stabilityμ > 0 and 3λ + 2μ > 0μ>0 and λ+μ>0

Example: Young–Laplace Equation

In order to further clarify the theory of surface elasticity and to demonstrate its generality, the classical example of the Young–Laplace equation is investigated. This well-known equation states that for static fluids the balance equation on the surface is given by
p=-γc
(76)
where p is the fluid pressure,6 and c is twice the mean curvature7 in the spatial configuration defined in Eq. (A1). The spatial balance of linear momentum on the surface (23)2 in the absence of external tractions reads
divσ=σ·n
(77)
Using the definition of the hydrostatic pressure p in terms of the Cauchy stress σ for inviscid fluids σ = − pi on the right-hand side of the previous equation, gives
divσ=-pn
(78)
In order to compare Eq. (78) with the Young–Laplace Eq. (76), the surface material model proposed in Eq. (70) with vanishing elastic parameters is selected, i.e.,
Ψ=γJP=FΨ=γJF-tσ=jP·Ft=γi
(79)
Inserting the σ=γi in Eq. (78) yields
div(γi)=-pnγdivi=-pn
(80)

and finally using the relation divi=cn yields γc=-p. Thus, it is clear that the curvature is embedded in the surface divergence operator.

Thermohyperelastic Material

The material models proposed in Secs. 6.1 and 6.2 describes the mechanical response of materials. They do not, however, possess enough structure to capture the thermal behavior and thermomechanically coupled phenomena. The objective of this section is to extend the discussion in the previous sections to account for thermomechanical contributions.

Bulk

For the simple case of thermohyperelasticity, the Helmholtz energy of the bulk is chosen as (see, e.g., Ref. [115])
Ψ(F,Θ):=12λln2J+12μ[F:F-3-2lnJ]-3ακ[Θ-Θ0]lnJJ+cF[Θ-Θ0-ΘlnΘΘ0]-Ξ0[Θ-Θ0]
(81)

The term 12λln2J+12μ[F:F-3-2lnJ] is the classical Helmholtz energy function of neo-Hookean-type presented in Eq. (64). Thermomechanical coupling is captured by the term −3αk[Θ-Θ0] ln J/J, where the thermal expansion coefficient α weights the product of the bulk compression modulus k and the difference between the current and the reference temperatures [Θ-Θ0]. The term cF[Θ-Θ0-Θ1nΘ/Θ0] represents the purely thermal behavior in terms of the specific heat capacity cF. The final term Ξ0 accounts for the material specific absolute entropy.

From the bulk Helmholtz energy in Eq. (81), the Piola stress and the entropy in the bulk follow in the familiar format of thermohyperelasticity from the constitutive laws (38)1 as
P=ΨF=[λlnJ-μ]F-t+μF-3ακ1J[Θ-Θ0][1-lnJ]F-t
and
Ξ=-ΨΘ=3ακlnJJ+cFlnΘΘ0+Ξ0
Next the corresponding contributions to the tangents required for a nonlinear finite element implementation using the Newton–Raphson scheme are derived. The partial derivatives of the Piola stress P with respect to deformation gradient F and temperature Θ are given by
PF=λF-tF-t+μI+[λlnJ-μ]D      +3ακ[Θ-Θ0][2-lnJJF-tF-t-1-lnJJD],PΘ=-3ακ1-lnJJF-t
The partial derivatives of the heat flux vector Q with respect to the temperature gradient GradΘ and the deformation gradient F are, respectively, given by
QGradΘ=-JkGandQF=-JkG·¯GradΘ
where, for the case of isotropic thermal conduction in the spatial configuration,8
GF-1·F-tandGF-1·F-tF-t+B
with
B:=F-1·F-tF=t[F-1·D]+F-1·D=-[F-1¯F-1]·F-t-F-1·[F-t¯F-1]
Finally, for the partial derivatives of the Gough–Joule contribution, one makes use of the identity
F-t:DtF=F-t:GradV=divV
to obtain
(ΘP,Θ:DtF)Θ=-3ακ1-lnJJdivV,(ΘP,Θ:DtF)F=-3ακΘ[lnJ-2JdivVF-t+1-lnJJD]
where
D:=divVF=F-t:DtFF=DtF:F-tF+F-t:DtFF

Surface

As in the bulk, a simple thermohyperelastic surface Helmholtz energy of the following form is chosen:
Ψ(F,Θ):=12λln2J+12μ[F:F-2-2lnJ]+γJ-2ακ[Θ-Θ0]lnJJ+cF[Θ-Θ0-ΘlnΘΘ0]-Ξ0[Θ-Θ0]
(82)
The term 12λln2J+12μ[F:F-2-2lnJ] is the classical Helmholtz energy function of neo-Hookean-type given in Eq. (70). The thermomechanical coupling is captured by the term -2ακ[Θ-Θ0]lnJ/J where α is the surface thermal expansion coefficient. The surface temperatures in the reference and current configurations are denoted as Θ0 and Θ, respectively. The surface specific heat capacity cF represents the purely thermal behavior. Finally, the surface specific entropy Ξ0 is also included. The surface Piola stress and entropy corresponding to the surface Helmholtz energy given in Eq. (82) follow from the constitutive laws (38)2 as
P=ΨF=[λlnJ-μ+γJ]F-t+μF-2ακ1J[Θ-Θ0][1-lnJ]F-t
and
Ξ=-ΨΘ=2ακlnJJ+cFlnΘΘ0+Ξ0
The surfaces contributions to the tangent are formally identical to those in the bulk. However, due to the rank-deficiency, subtle modifications need to be taken into account. Thus,
PF=[λ+γJ]F-tF-t+μI+[λlnJ-μ+γJ]D      +2ακ[Θ-Θ0][2-lnJJF-tF-t-1-lnJJD],PΘ=-2ακ1-lnJJF-t
The partial derivatives of the interface heat flux vector Q with respect to the temperature gradient GradΘ and the interface deformation gradient F are
QGradΘ=-JkGandQF=-JkG·¯GradΘ
where for isotropic thermal conduction along the interface in the spatial configuration9
GF-1·F-tandGF-1·F-tF-t+B
with
B:=F-1·F-tF=t[F-1·D]+F-1·D
Finally, for the partial derivatives of the Gough–Joule contribution, one makes use of the identity
F-t:DtF=F-t:GradV=divV
to obtain
(ΘP,Θ:DtF)Θ=-2ακ1-lnJJdivV,(ΘP,Θ:DtF)F=-2ακΘ[lnJ-2JdivVF-t+1-lnJJD]
where
D:=divVF=F-t:DtFF

Interface

In an identical fashion to the surface, a simple thermohyperelastic interface Helmholtz energy of the following form is chosen:
Ψ¯(F¯,Θ¯):=12λ¯ln2J¯+12μ¯[F¯:F¯-2-2lnJ¯]+γ¯J¯-2α¯κ¯[Θ¯-Θ¯0]lnJ¯J+c¯F¯[Θ¯-Θ¯0-Θ¯lnΘ¯Θ¯0]-Ξ¯0[Θ¯-Θ¯0]
(83)

The term 12λ¯ln2J¯+12μ¯[F¯:F¯-2-2lnJ¯] mimics the classical Helmholtz-energy function of neo-Hookean-type characterized through the two interface Lamé constants λ¯ and μ¯. The term γ¯ represents surface-tension-like behavior on the interface. Thermomechanical coupling is captured by the term −2α¯κ¯[Θ¯-Θ¯0]lnJ¯/J where α¯ is the interface thermal expansion coefficient and κ¯:=μ¯+λ¯ the interface compression modulus. The interface temperatures in the reference and current configurations are denoted as Θ¯0 and Θ¯, respectively. The interface specific heat capacity c¯F¯ represents the purely thermal behavior. Finally, the interface specific entropy Ξ¯0 is included.

From the constitutive laws (38)3 the interface Piola stress and entropy corresponding to the interface Helmholtz energy in Eq. (83) are obtained as
P¯=Ψ¯F¯=[λ¯lnJ¯-μ¯+γ¯J¯]F¯-t+μ¯F¯-2α¯κ¯1J¯[Θ¯-Θ¯0][1-lnJ¯]F¯-t
and
Ξ¯=-Ψ¯Θ¯=2α¯κ¯lnJ¯J¯+c¯F¯lnΘ¯Θ¯0+Ξ¯0
Remark. The interface Helmholtz energy given in Eq. (83) accounts for thermomechanical coupling and the interface heat capacity. For a Kapitza interface these two contributions, as well as the heat conduction along the interface and the interface prescribed heat source, should vanish to satisfy the Kapitza assumption of continuity of the normal heat flux across the interface (see Eq. (55)3). Therefore, the form of the Helmholtz energy for the generalized Kapitza interface simplifies to
Ψ¯(F¯,Θ¯)=12λ¯ln2J¯+12μ¯[F¯:F¯-2-2lnJ¯]+γ¯J¯-Ξ¯0[Θ¯-Θ¯0]
(84)

Curve

Finally, for the sake of completeness, and in a very similar fashion to the surface and interface, a simple thermohyperelastic curve Helmholtz energy of the following form is chosen:
Ψ˜(F˜,Θ˜):=12μ˜[F˜:F˜-1-2lnJ˜]+γ˜J˜-α˜μ˜[Θ˜-Θ˜0]lnJ˜J+c˜F˜[Θ˜-Θ˜0-Θ˜lnΘ˜Θ˜0]-Ξ˜0[Θ˜-Θ˜0]
(85)

The term 12μ˜[F˜:F˜-1-2lnJ˜] mimics the classical Helmholtz-energy function of neo-Hookean-type characterized through a constant μ˜. Due to the one-dimensional nature of the curve, one mechanical parameter suffices to model the elastic behavior of the curve. In this sense, the curve parameter μ˜ can be understood as the elastic-modulus or the compression modulus of the curve and shear modulus is no longer meaningful. A surface-tension-like term is introduced by γ˜. The thermomechanical coupling is captured by the term -α˜μ˜[Θ˜-Θ˜0]lnJ˜/J where α˜ is the curve thermal expansion coefficient. The curve temperatures in the reference and current configurations are denoted as Θ˜0 and Θ˜, respectively. The curve specific heat capacity c˜F˜ represents the purely thermal behavior. Finally, the curve specific entropy Ξ˜0 is included.

From the constitutive laws (38)4 the curve Piola stress and entropy corresponding to the curve Helmholtz energy Eq. (85) are obtained as
P˜=Ψ˜F˜=[-μ˜+γ˜J˜]F˜-t+μ˜F˜-α˜μ˜1J˜[Θ˜-Θ˜0][1-lnJ˜]F˜-t
and
Ξ˜=-Ψ˜Θ˜=α˜μ˜lnJ˜J˜+c˜F˜lnΘ˜Θ˜0+Ξ˜0

Numerical Examples

The objective of this section is to briefly elucidate the theory presented in Secs. 2–6 using a series of numerical examples. The roles of the surface and interface on the overall response of a body are studied.10 Specific attention is paid to the influence of the material parameters.

These examples are of particular relevance to problems involving materials at the nanoscale. The influence of decreasing specimen size on the overall response can be captured by fixing the bulk material parameters and increasing those of the surface and interface. This is equivalent to fixing the bulk Helmholtz energy and increasing the Helmholtz energy of the surface and interface.

The material behavior in the bulk is characterized by the thermohyperelastic Helmholtz energy function given in Eq. (81). Heat conduction in the bulk is assumed to be spatially isotropic, i.e., k = k i.

The material parameters for the bulk are stated in Table 6 and hold unless specified otherwise.

Table 6

Bulk material properties

Lamé constantμ80193.8N/mm2
Lamé constantλ110743.5N/mm2
Thermal expansion coefficientα1.0 × 10−51/K
Conductivityk45.0N/(s K)
Specific capacitycF3.588N/(mm2 K)
Reference temperatureΘ0298.0K
Lamé constantμ80193.8N/mm2
Lamé constantλ110743.5N/mm2
Thermal expansion coefficientα1.0 × 10−51/K
Conductivityk45.0N/(s K)
Specific capacitycF3.588N/(mm2 K)
Reference temperatureΘ0298.0K

Surface

The following set of examples illustrate the role of the surface on the overall response. Thus, the surface Helmholtz energy presented in Sec. 6.3.2 is chosen. The role of the various terms in the energy is then studied by varying their corresponding constitutive coefficients. In Sec. 7.1.1 the influence of surface tension γ is demonstrated. In contrast to all of the other terms in the surface Helmholtz energy, it does not vanish at the reference (strain-free) configuration and acts in a fashion similar to an external force field. In this sense, it should be treated as a prescribed force and, therefore, is applied incrementally [90].

The surface theory presented here is not restricted to the isotropic case and anisotropic behavior can be captured if the form of the energy includes the anisotropy via a dependence on the spatial surface normal; see Eq. (38). For the case of surface tension a simple anisotropic model is proposed. The numerical results demonstrate the flexibility and generality of the proposed scheme.

In Sec. 7.1.2 the focus is on the thermomechanical behavior of the surface and its coupling with the bulk. The influence of surface tension is ignored. Furthermore heat conduction on the surface is assumed to be spatially isotropic, i.e., k=ki.

Surface Tension

Consider the unit-cube depicted in Fig. 4 (top-left). The cube is discretized using 1728 trilinear hexahedral elements and 2197 nodes. The cube is covered by an energetic surface with a surface Helmholtz energy containing only the surface tension terms in Eq. (82). The surface tension of γ = μ is prescribed in 20 equal steps.

Fig. 4
Transformation of a cube into a sphere due to an incrementally prescribed surface tension of γ∧/μ = 1.0 mm
Fig. 4
Transformation of a cube into a sphere due to an incrementally prescribed surface tension of γ∧/μ = 1.0 mm
Close modal

For a nonvanishing surface tension the reference configuration does not coincide with that of an equilibrium state. Thus, the cube deforms in order to reach an equilibrium state, i.e., to minimize the total energy of the bulk and the surface. Clearly, any deformation introduces an increase in the bulk energy as it is zero at the reference configuration. Nevertheless, the total energy of the system can still decrease by decreasing the surface energy and increasing the bulk energy. Recall that surface tension is equivalent to a constant energy per unit deformed area. Therefore, the surface energy can decrease by decreasing the surface area. As a result, the cube tends to transform into a shape where the surface-to-volume ratio is minimal, i.e., a sphere. The surface also tends to shrink. The transformation of the cube into a spherical shape is illustrated in Fig. 4.

The role of anisotropy is now investigated. The example is identical to the previous one except that the surface tension γ is no longer constant. It depends on the surface normal and the surface Jacobian via the relation
γ=μ[1+βJ[n·e]2]

where the material parameter β controls the anisotropic contribution and the vector e is a given unit vector characterizing a preferred direction. The anisotropic effects will be greatest where the normal to the surface is parallel to e and will vanish where the normal to the surface is orthogonal to e. The vector e = [0,0,1] together with the reference and deformed configurations are shown in Fig. 5. The anisotropic effect will be maximum on the top and the bottom of the model and will vanish on the lateral walls of the model (elsewhere the effect is between these two extremes). The model is expected, therefore, to shrink more on the top and the bottom and less at the lateral walls of the cube. Consequently, the cube will transform into an ellipsoidal shape instead of the spherical one shown in Fig. 4. The resulting ellipsoidal shapes for different values of β are shown in Fig. 5.

Fig. 5
Transformation of a unit-cube into an ellipsoid due to a prescribed anisotropic surface tension γ∧/μ = [1 + β∧J∧[n·e]2] mm
Fig. 5
Transformation of a unit-cube into an ellipsoid due to a prescribed anisotropic surface tension γ∧/μ = [1 + β∧J∧[n·e]2] mm
Close modal

Thermomechanical Surface

Consider the three-dimensional strip shown in Fig. 6. The strip is subjected to plane-strain-like conditions across the thickness. In the middle of the specimen is a hole of radius 0.2 mm. Lateral deformations are prevented, i.e., the width of the strip cannot change. A displacement of 0.5 mm is prescribed at the edges resulting in a tensile loading. Thermally homogeneous Neumann boundary conditions are applied to all boundaries of the bulk. The resulting amount of deformation is clearly in the finite strain regime. The displacement loading is applied in 20 equal steps and the total time is 10 ms. The strip is discretized using 576 trilinear hexahedral elements and 1008 nodes. Assume that the surface of the hole in the specimen is energetic. The surface Helmholtz energy is chosen to be of the form given in Eq. (82) with vanishing γ. The response of the material without the energetic surface is studied first. Thereafter, mechanical and thermomechanical surface effects are detailed, respectively.

Fig. 6
Strip with surface: geometry (left) and applied boundary conditions (right). Dimensions are in mm. The thickness is 0.1 mm.
Fig. 6
Strip with surface: geometry (left) and applied boundary conditions (right). Dimensions are in mm. The thickness is 0.1 mm.
Close modal
Bulk effect.

Consider first a thermomechanical bulk without an energetic surface. Under the prescribed boundary conditions the tensile stress increases. This increase of the tensile stress, or more precisely the divergence of the strain rate, cools the specimen due to the Gough–Joule effect. The presence of the hole leads to a nonuniform stress distribution in the specimen as illustrated in Fig. 7. Consequently, the nonuniform temperature profile depicted in Fig. 8 is obtained.

Fig. 7
The stress distribution for a thermomechanical bulk. The results (a)–(f) correspond to the reference configuration and 20%, 40%, 60%, 80%, and 100% of the final deformation, respectively. The stress depicted is the xx-component of the Cauchy stress tensor. The reference configuration is indicated using a dashed (white) line.
Fig. 7
The stress distribution for a thermomechanical bulk. The results (a)–(f) correspond to the reference configuration and 20%, 40%, 60%, 80%, and 100% of the final deformation, respectively. The stress depicted is the xx-component of the Cauchy stress tensor. The reference configuration is indicated using a dashed (white) line.
Close modal
Fig. 8
The temperature distribution for a thermomechanical bulk. The results (a)–(f) correspond to the reference configuration and 20%, 40%, 60%, 80%, and 100% of the final deformation, respectively. The reference configuration is indicated using a dashed (white) line.
Fig. 8
The temperature distribution for a thermomechanical bulk. The results (a)–(f) correspond to the reference configuration and 20%, 40%, 60%, 80%, and 100% of the final deformation, respectively. The reference configuration is indicated using a dashed (white) line.
Close modal
Mechanical surface effect.

Consider now a purely mechanical surface. That is k=cF=α=0. The mechanical material parameters for the surface are assumed to be μ/μ=λ/λ=2mm. The mechanical resistance of the surface influences the bulk, resulting in a different stress distribution when compared to that shown Fig. 7 obtained in the absence of surface effects. The evolution of the stress field is illustrated in Fig. 9. The increase in the tensile stress cools the specimen due to the Gough–Joule effect. Thus, the temperature decreases as the applied deformation increases. The nonhomogeneous stress distribution produces the nonhomogeneous temperature distribution shown in Fig. 10.

Fig. 9
The influence a purely mechanical surface on the stress distribution for μ∧/μ=λ∧/λ = 2 mm. The results (a)–(f) correspond to the reference configuration and 20%, 40%, 60%, 80%, and 100% of the final deformation, respectively. The stress depicted is the xx-component of the Cauchy stress tensor. The reference configuration is indicated using a dashed (white) line.
Fig. 9
The influence a purely mechanical surface on the stress distribution for μ∧/μ=λ∧/λ = 2 mm. The results (a)–(f) correspond to the reference configuration and 20%, 40%, 60%, 80%, and 100% of the final deformation, respectively. The stress depicted is the xx-component of the Cauchy stress tensor. The reference configuration is indicated using a dashed (white) line.
Close modal
Fig. 10
The influence of a purely mechanical surface on the temperature distribution for μ∧/μ=λ∧/λ = 2 mm. The results (a)–(f) correspond to the reference configuration and 20%, 40%, 60%, 80%, and 100% of the final deformation, respectively. The reference configuration is indicated using a dashed (white) line.
Fig. 10
The influence of a purely mechanical surface on the temperature distribution for μ∧/μ=λ∧/λ = 2 mm. The results (a)–(f) correspond to the reference configuration and 20%, 40%, 60%, 80%, and 100% of the final deformation, respectively. The reference configuration is indicated using a dashed (white) line.
Close modal
Thermomechanical surface effect.

The surface is now endowed with thermal properties. The mechanical properties of the surface are fixed as before by setting μ/μ=λ/λ=2mm. In order to better understand the role of the surface thermal properties, the surface conductivity k, heat capacity cF, and heat expansion coefficient α are investigated separately.

The effect of surface conductivity.

The surface heat conduction coefficient k is increased from zero to k/k = 100 mm. The remaining surface thermal parameters are set to zero. Figure 11 shows the temperature distributions at the end of the applied loading for a range of k/k. Increasing the surface conductivity clearly produces an increasingly uniform temperature distribution along the surface. Note that the result shown in Fig. 11 for k/k = 0 is the same as shown in Fig. 10(f).

Fig. 11
The influence of interface conduction on the temperature distribution for μ∧/μ=λ∧/λ = 2 mm and k∧/k = 0–100 mm
Fig. 11
The influence of interface conduction on the temperature distribution for μ∧/μ=λ∧/λ = 2 mm and k∧/k = 0–100 mm
Close modal
The effect of surface heat capacity.

The surface heat capacity cF is now varied from zero to cF/cF=1mm. The remaining surface thermal parameters are set as zero. Increasing the surface heat capacity causes the surface to conserve its initial temperature and resist temperature change. A higher value of the surface heat capacity results in less temperature decrease at the surface. This can be observed in Fig. 12, which shows the final temperature distributions for different surface heat capacities.

Fig. 12
The influence of the interface heat capacity on the temperature distribution for μ∧/μ = λ∧/λ = 2 mm and c∧F∧/cF = 0–1 mm
Fig. 12
The influence of the interface heat capacity on the temperature distribution for μ∧/μ = λ∧/λ = 2 mm and c∧F∧/cF = 0–1 mm
Close modal
The effect of the surface Gough–Joule contribution.

Finally, the effect of the surface heat expansion coefficient α on the temperature distribution is studied. The surface heat expansion coefficient α is varied from zero to α/α=0.75 while the remaining surface thermal parameters are specified as zero. According to Eq. (82), increasing the surface heat expansion coefficient increases the surface thermomechanical coupling. As a result of the Gough–Joule effect, the surface cools due to the positive strain rate applied to the surface from the surrounding bulk material. The surface cooling is proportional to α/α. This explains the temperature distributions shown in Fig. 13.

Fig. 13
The influence of the interface heat expansion coefficient on temperature distribution for μ∧/μ=λ∧/λ = 2 mm and α∧/α = 0–0.75
Fig. 13
The influence of the interface heat expansion coefficient on temperature distribution for μ∧/μ=λ∧/λ = 2 mm and α∧/α = 0–0.75
Close modal

Interface

The “toy” problem is a cube with an internal energetic interface, as illustrated in Fig. 14. The material interior and exterior to the interface is identical. For both the bulk and the interface a thermohyperelastic material model is assumed. The choice of the Kapitza or highly conducting assumption on the interface's behavior is examined.

Fig. 14
A cube containing an internal energetic interface. The geometry is show on the left and the two different load cases on the right. The first load case corresponds to a thermal load while the second is a mechanical one.
Fig. 14
A cube containing an internal energetic interface. The geometry is show on the left and the two different load cases on the right. The first load case corresponds to a thermal load while the second is a mechanical one.
Close modal

Kapitza Interface

The interface is initially defined to be a Kapitza interface. Hence, while discontinuities in the temperature field are permitted across the interface, the normal component of the heat flux vector is continuous. The domain is loaded according to the thermal load case depicted in Fig. 14. The top and the lateral sides of the cube are thermally isolated, i.e., Qp = 0. A temperature increase of ΔΘp = 40 K relative to the initial temperature of Θ0 = 298 K is applied to the lower face of the specimen. The resulting temperature distributions at the same instant in time for different Kapitza resistance coefficients r¯Q are illustrated in Fig. 15. As expected, the jump in the temperature across the interface increases with increasing Kapitza resistance.

Fig. 15
The temperature distribution for varying values of the Kapitza resistance r¯Q on the interface
Fig. 15
The temperature distribution for varying values of the Kapitza resistance r¯Q on the interface
Close modal

Highly Conducting Interface

The interface is now assumed to be highly conducting. Hence, while discontinuities in the normal component of the heat flux vector are permitted across the interface, the temperature field is continuous. The domain is subjected to the mechanical loading conditions illustrated in Fig. 14. The cube is extended by 100% by prescribing equal displacements to the top and bottom faces. Lateral contractions are prevented. The loading is applied quasi-statically and the cube is thermally isolated.

Consider first the response without an interface as shown in Fig. 16(a). In the absence of the interface, the stress state in the cube is homogeneous. The inclusion of the energetic interface results in a change in the distribution of the stress field, as shown in Fig. 16(b). The mechanical resistance of the interface causes the stress to increase above and below the region enclosed by the interface. Half of the midplane of the cube, corresponding to the cases where the interface was omitted and included, is shown alongside one another in Fig. 16(c) to facilitate the comparison of the mechanical response. The mechanical resistance of the energetic interface causes the material surrounded by the interface to deform less than the corresponding region in the absence of an interface (the spatial discretizations are identical).

Fig. 16
The stress distribution without and with an HC interface are shown in (a) and (b), respectively, and compared in (c). The resulting temperature distribution is shown in (d).
Fig. 16
The stress distribution without and with an HC interface are shown in (a) and (b), respectively, and compared in (c). The resulting temperature distribution is shown in (d).
Close modal

The nonhomogeneous stress distribution that arises due the mechanical resistance of the interface results in a nonhomogeneous temperature field as shown in Fig. 16(d). The Gough–Joule effect causes the temperature in the body to decrease. The temperature field is clearly continuous across the interface.

Conclusion

A procedure for deriving the equations governing the response of a solid bulk, surrounded by a surface, and intersected by an interface, all of which are energetic, has been presented. The interaction between the surface and the interface has also been accounted for. The structure of the constitutive relations in the various parts of the body was made clear by assigning these regions distinct Helmholtz energies and following a Coleman–Noll-like procedure. The remaining reduced dissipation inequalities suggest the structure of the heat and entropy flux vectors and permissible restrictions on the temperature field in the various parts of the body; particular attention was paid to the interface. The weak formulation of the governing equations was then given and compared to six commonly used restrictions thereof. The theory was elucidated using several numerical examples.

The theory proposed here is often motivated by applications in nanomechanics where size-effects play a critical role. The ratio of the bulk Helmholtz energy to those in the remaining parts of the body implicitly accounts for the size-effects. However, the theory lacks a physical length scale related to the thickness of the interface or surface. Steigmann and Ogden [40] incorporate local elastic resistance to flexure into the Gurtin and Murdoch [32] theory, thereby introducing an inherent length scale and regularizing the associated variational problem. Recent work by Forest et al. [135], comparing first and second strain gradient theories for capillary effects in elastic fluids at small length scales, emphasizes the importance of nonlocality and the associated issue of length scale. An extension to account for the flexural resistance of the surface or interface is, therefore, suggested (see also Refs. [41,42,136]).

Acknowledgment

The support of this work by the European Research Council Advanced Grant MOCOPOLY is gratefully acknowledged. The second author thanks the National Research Foundation of South Africa for their support.

Geometry of Surfaces and Curves

It is enlightening to briefly review some basic terminologies and results on surfaces and curves. For further details the reader is referred to Refs. [89,102–104] among others. Here, some technicalities are borrowed from Steinmann [89].

Surfaces
A two-dimensional (smooth) surface S in the three-dimensional, embedding Euclidean space with coordinates x is parameterized by two surface coordinates ηα with α = 1, 2 as
x=x(ηα)
The corresponding tangent vectors aαTS to the surface coordinate lines ηα, i.e., the covariant (natural) surface basis vectors are given by
aα=ηαx
The associated contravariant (dual) surface basis vectors aα are defined by the Kronecker property δβα=aα·aβ and are explicitly related to the covariant surface basis vectors aα by the co- and contravariant surface metric coefficients aαβ (first fundamental form of the surface) and aαβ, respectively, as
aα=aαβaβwithaαβ=aα·aβ=[aαβ]-1,aα=aαβaβwithaαβ=aα·aβ=[aαβ]-1
The contra- and covariant base vectors a3 and a3, normal to TS, are defined by
a3:=a1×a2anda3:=[a33]-1a3sothata3·a3=1
Thereby, the corresponding contra- and covariant metric coefficients, respectively, [a33] and [a33] follow as
[a33]=|a1×a2|2=det[aαβ]=[det[aαβ]]-1=[a33]-1
Accordingly, the surface area element ds and the surface normal n are computed as
ds=|a1×a2|dη1dη2=[a33]1/2dη1dη2and   n=[a33]1/2a3=[a33]1/2a3
Moreover, with i denoting the ordinary mixed-variant unit tensor of the three-dimensional, embedding Euclidian space, the mixed-variant surface unit tensor i is defined as
i:=δβαaαaβ=aαaα=i-a3a3=i-nn

Clearly the mixed-variant surface unit tensor acts as a surface (idempotent) projection tensor.

The surface gradient and surface divergence operators for vector fields are defined by
grads{}:=ηα{}aαanddivs:=ηα{}·aα
As a consequence, observe that grads{}·n=0 holds by definition. For fields that are smooth in a neighborhood of the surface, the surface gradient, and surface divergence operators are alternatively defined as
grads{}:=grad{}·ianddivs:=grads{}:i=grad{}:i
Finally, the Gauss formulae for the derivatives of the co- and contravariant surface basis vectors read
ηβaα=Γαβγaγ+kαβnandηβaα=-Γβγαaγ+kβαn
where Γαβγ=ηβaα·aγ denote the surface Christoffel symbols and kαβ are the coefficients of the curvature tensor. The curvature tensor k=kαβaαaβ and twice the mean curvature k=kαα of the surface S are defined as the negative surface gradient and surface divergence of the surface normal n, respectively,
k:=-gradsn=-ηβnaβandk:=-divsn=-ηβn·aβ
(A1)

The covariant coefficients of the curvature tensor (second fundamental form of the surface) are computed by kαβ=aα·k·aβ=-aα·ηβn.

Curves
A one-dimensional (smooth) curve C in the three-dimensional, embedding Euclidian space with coordinates x is parameterized by the arc length η as
x=x(η)
The corresponding tangent vector tTC to the curve, together with the (principal) normal and binormal vectors m and b, orthogonal to TC, are defined by
t:=ηxandm:=ηt/|ηt|andb:=t×m
Due to the parametrization of the curve in its arc length η, the tangent vector t has unit length and the curve line element dc is computed as
dc=|ηx|dη=|t|dη=dη
Moreover, we define the mixed-variant curve unit tensor i˜ as
i˜:=tt=i-mm-bb

Clearly the mixed-variant curve unit tensor acts as a curve (idempotent) projection tensor.

The curve gradient and curve divergence operators for vector fields are defined by
gradc{}:=η{}tand   divc:=η{}·t
As a consequence, observe that gradc{}·m=0 and gradc{}·b=0 hold by definition. For fields that are smooth in a of the curve, the curve gradient and curve divergence operators are alternatively defined as
gradc{}:=grad{}·i˜and   divc:=gradc{}:i˜=grad{}:i˜
Finally, the Frénet–Serret formulae for the derivatives of the curve basis vectors read
ηt=k˜mandηm=-k˜t+k'˜bandηb=-k'˜m
where k˜ and k˜' denote the scalar valued curvature and torsion of the curve. Based on the Frénet–Serret formulae, the curve divergence of the curve unit tensor i˜ renders
divci˜=k˜m
Likewise, the curvature tensors k˜||,k˜, and the scalar valued curvature k˜ are defined as the curve gradient of the curve tangent t and the negative curve gradient and curve divergence of the curve normal m, respectively,
k˜||:=gradct=ηttandk˜:=-gradcm=-ηmt,k˜:=-divcm=-ηm·t

The curvature tensors are computed as k˜||=k˜mt and k˜=k˜tt-k'˜bt.

Useful Relations

In order to derive the balance equations presented in Sec. 3, a number of mathematical identities are needed. They are listed here without proof. For scalars R,R,R¯, and R˜ vectors Q, Q,Q¯, and Q˜, and second-order tensors P, P,P,¯andP˜ the following relations hold:
Div(Q×P)=Q×DivP+ε:[GradQ·Pt]inB0Div(Q×P)=Q×DivP+ε:[GradQ·Pt]inS0Div¯(Q¯×P¯)=Q¯×Div¯P¯+ε:[Grad¯Q¯·P¯t]inI0Div˜(Q˜×P˜)=Q˜×Div˜P˜+ε:[Grad˜Q˜·P˜t]inC0
(B1)
Div(RQ)=RDivQ+Q·GradRinB0Div(RQ)=RDivQ+Q·GradRinS0Div¯(R¯Q¯)=R¯Div¯Q¯+Q¯·Grad¯R¯inI0Div˜(R˜Q˜)=R˜Div˜Q˜+Q˜·Grad˜R˜inC0
(B2)
Div(Q·P)=Q·DivP+P:GradQinB0Div(Q·P)=Q·DivP+P:GradQinS0Div¯(Q¯·P¯)=Q¯·Div¯P¯+P¯:Grad¯Q¯inI0Div˜(Q˜·P˜)=Q˜·Div˜P˜+P˜:Grad˜Q˜inC0
(B3)

Reformulation of the Global Working in Terms of the Various Stress Measures

The steps to reformulate the expression for the global working (10) in terms of the stress measures in the various parts of the body, as given in Eq. (22), are as follows. The global working
W
0 is defined in Eq. (10) by
W0:=B0V·BpdV+S0extV·BpdA+I0V¯·B¯pdA+C0extV˜·B˜pdL+S0intV·[P·M]dA+L0V·[P·N]dL+C0intV¯·[P¯·M¯]dL+C0extV˜·[P˜·N˜]
Using the balances of linear momentum (21) to express the various stress measures in terms of their associated bulk, surface, interface, and curve forces, denoted Bp,Bp,B¯p,andB˜p, respectively, the global working becomes
W0:=B0V·[-DivP]dV+S0extV·[-DivP+P·N]dA+I0V¯·[-Div¯P¯-[[P]]·N¯]dA+C0extV˜·[-Div˜P˜+P¯·Ni˜-[[P]]·N˜S]dL+S0intV·[P·M]dA+L0V·[P·N]dL+C0intV¯·[P¯·M¯]dL+C0extV˜·[P˜·N˜]
It follows from the definition of the control regions, given in Sec. 2.3, that
S0int{}dA=B0{}dA-S0ext{}dA,C0int{¯}dL=I0{¯}dL-C0ext{¯}dL
Furthermore, the outward normal to
B0
on the surface is the same as the outward normal to the surface, i.e.,
M|S0ext
 = N. In an identical fashion to the surface
M¯|C0ext=N˜i
. Due to kinematic slavery on the surface
V|S0ext=V
and
V¯|C0ext=V˜
. From these observations, the following two terms in the global working can be restated as follows:
S0intV·[P·M]dA=B0V·[P·M]dA-S0extV·[P·N]dA,L0V·[P·N]dL=I0V¯·[P¯·M¯]dL-C0extV˜·[P¯·N˜i]dL
Thus, the global working becomes
W0=B0V·[-DivP]dV+S0extV·[-DivP]dA+I0V¯·[-Div¯P¯-[[P]]·N¯]dA+C0extV˜·[-Div˜P˜-[[P]]·N˜S]dL+B0V·[P·M]dA+I0V¯·[P¯·M¯]dL+L0V·[P·N]dL+C0extV˜·[P˜·N˜]
and using the identities (B3) yields
=B0P:GradVdV-B0Div(V·P)dV+S0extP:GradVdA-S0extDiv(V·P)dA+I0P¯:Grad¯V¯dA-I0Div¯(V¯·P¯)dA-I0V¯·[[P]]·N¯dA+C0extP˜:Grad˜V˜dL-C0extDiv˜(V˜·P˜)dL-C0extV˜·[[P]]·N˜sdL+B0V·P·MdA+I0V¯·P¯·M¯dL+L0V·P·NdL+C0extV˜·P˜·N˜
Finally, using the extended divergence theorems for the various parts of the body (2)–(5) gives
W0=B0P:GradVdV+S0extP:GradVdA+I0P¯:GradV¯dA+C0extP˜:Grad˜V˜dL-B0V·P·MdA+I0[[V·P]]·N¯dA-L0V·P·NdL+S0extCV·P·NdA+C0ext[[V·P]]·N˜sdL-I0V¯·P¯·M¯dL+I0C¯V¯·P¯·N¯dA-C0extV˜·P˜·N˜+C0extC˜V˜·P˜·N˜cdL-I0V¯·[[P]]·N¯dA-C0extV˜·[[P]]·N˜sdL+B0V·P·MdA+I0V¯·P¯·M¯dL+L0V·P·NdL+C0extV˜·P˜·N˜=B0P:GradVdV+S0extP:GradVdA+I0P¯:Grad¯V¯dA+C0extP˜:Grad˜V˜dL+I0[[V·P]]·N¯dA+C0ext[[V·P]]·N˜sdL-I0V¯·[[P]]·N¯dA-C0extV˜·[[P]]·N˜sdL

The interface is assumed to be coherent; thus [[V]] = 0 and [[V]] = 0. Therefore, V¯·[[P]] = [[V·P]] and V˜·[[P]] = [[V·P]] and the last four terms vanish. What remains is the global working given in Eq. (22).

Derivation of the Interface Temperature Evolution Equation

It is enlightening to review the procedure to obtain the nonstandard interface temperature evolution Eq. (55)3.

The balance of internal energy on the interface (26)3 reads
P¯:Grad¯V¯-Div¯Q¯+Q¯=DtE¯withQ¯=Q¯p-[[Q]]·N¯onI0
The rate of change of the internal energy on the interface DtE¯ is expanded using the definition of the Helmholtz energy (35)3 as follows:
DtE¯=Dt(Ψ¯+Θ¯Ξ¯)
and from the parametrization of the Helmholtz energy given in Eq. (37)3,
=[Ψ¯F¯+Ψ¯n¯·n¯F¯]:DtF¯+Ψ¯Θ¯DtΘ¯+Ξ¯DtΘ¯+Θ¯DtΞ¯,
and from the constitutive relations (38)3,
=P¯:DtF¯+Θ¯DtΞ¯
Inserting this result into balance of energy on the interface (26)3 yields
P¯:Grad¯V¯-Div¯Q¯+Q¯=P¯:DtF¯+Θ¯DtΞ¯
and recalling that
Grad¯V¯=X¯(V¯)=X¯(Dtx¯)=Dt(x¯X¯)=DtF¯
allowing the balance of energy to be expressed as
-Div¯Q¯+Q¯=Θ¯DtΞ¯
(D1)
The right-hand side of Eq. (D1) can be simplified as follows:
Θ¯DtΞ¯=Θ¯Ξ¯Θ¯DtΘ¯+Θ¯Ξ¯F¯:DtF¯+Θ¯Ξ¯n¯·Dtn¯
and using the constitutive relation Ξ¯=-(Ψ¯/Θ¯),
=-Θ¯2Ψ¯Θ¯Θ¯DtΘ¯-Θ¯2Ψ¯F¯Θ¯:DtF¯+Θ¯Ξ¯n¯·n¯F¯:Dtn¯=-Θ¯2Ψ¯Θ¯Θ¯c¯F¯DtΘ¯-Θ¯Θ¯[Ψ¯F¯+Ψ¯n¯·n¯F¯P¯]:DtF¯

where c¯F¯ is the interface heat capacity. Finally, inserting this result into Eq. (D1) gives the equation for the evolution of the interface temperature as in Eq. (55)3.

1

Pyotr Leonidovich Kapitza (1894–1984) was a physicist and Nobel laureate who first proposed the presence of thermal interface resistance corresponding to a discontinuous temperature field across an interface for liquid helium; see Kapitza [82].

2

Here and henceforth, the subscripts t and 0 shall designate spatial and material quantities, respectively, unless specified otherwise.

3

Note that inertial forces are omitted. Furthermore, in the absence of mass flux, mass is conserved according to the standard relations given in Table 1.

4
The reduced dissipation inequality on the interface is alternatively expressed in the literature (see, e.g., Ref. [47]) in one of the following forms:
D¯=Θ¯[[Q·[1Θ-1Θ¯]]]·N¯=Θ¯[[QΘ·[1-ΘΘ¯]]]·N¯=[[QΘ·[Θ¯-Θ]]]·N¯0
5

The term volumetric is used in analogy to the bulk. Nevertheless, it has a different meaning on the surface. A volumetric deformation in the bulk is a deformation mode that changes the volume uniformly. A volumetric surface deformation, however, is a deformation mode that changes the area uniformly. In this sense the term spherical seems more appropriate. Nonetheless, for the sake of consistency, the term volumetric is used henceforth.

6

For the sake of simplicity, it is assumed that the surface is an interface between the fluid and an outside vacuum.

7

Often in the literature c is written as sum of the principal curvatures c = −[1/r1+ 1/r2] where r1 and r2 denote the principal radii of curvature. Based on this definition, the curvature is negative if the surface curves away from the normal.

8

For isotropic thermal conduction in the material configuration GI and GIF-t.

9

For isotropic thermal conduction in the material configuration GIandGIF-t.

10

Numerical investigations of a more general case where the surface, interface, and curve are all considered involve additional complications while providing little additional insight into the problem and, therefore, are not studied here.

References

1.
Pawlow
,
P.
,
1908
, “
Über die Abhängigkeit des Schmelzpunktes von der Oberflächenenergie eines festen Körpers
,”
Z. physik. Chemie
,
65
, pp.
1
35
.
2.
Gibbs
,
J. W.
,
1961
,
The Scientific Papers of J.W. Gibbs
,
Dover Publications, New York
.
3.
Guggenheim
,
E. A.
,
1940
, “
The Thermodynamics of Interfaces in Systems of Several Components
,”
Trans. Faraday Soc.
,
35
, pp.
397
412
.10.1039/tf9403500397
4.
Adam
,
N. K.
,
1941
,
The Physics and Chemistry of Surfaces
,
Oxford University Press
,
London
.
5.
Shuttleworth
,
R.
,
1950
The Surface Tension of Solids
,”
Proc. Phys. Soc. A
,
63
(
5
), pp.
444
457
.10.1088/0370-1298/63/5/302
6.
Herring
,
C.
,
1951
, “
Some Theorems on the Free Energies of Crystal Surfaces
,”
Phys. Rev.
,
82
(
1
), pp.
87
93
.10.1103/PhysRev.82.87
7.
Bikerman
,
J. J.
,
1965
, “
Surface Energy of Solids
,”
Phys. Status Solid. B
,
10
(
1
), pp.
3
16
.10.1002/pssb.19650100102
8.
Orowan
,
E.
,
1970
, “
Surface Energy and Surface Tension in Solids and Liquids
,”
Proc. R. Soc. A
,
316
(
1527
), pp.
473
491
.10.1098/rspa.1970.0091
9.
Cahn
,
J. W.
,
1978
, “
Thermodynamics of Solid and Fluid Surfaces
,”
Segregation to Interfaces
(
ASM Seminar Series
), Cleveland, OH, pp.
3
23
.
10.
Cahn
,
J. W.
,
1989
, “
Interfacial Free Energy and Interfacial Stress: The Case of an Internal Interface in a Solid
,”
Acta Metall. Mater.
,
37
(
3
), pp.
773
776
.10.1016/0001-6160(89)90004-7
11.
Murr
,
L. E.
,
1975
,
Interfacial Phenomena in Metals and Alloys
,
Addison-Wesley
, Boston, MA.
12.
Rottman
,
C.
,
1988
, “
Landau Theory of Coherent Interphase Interfaces
,”
Phys. Rev. B
,
38
, pp.
12031
12034
.10.1103/PhysRevB.38.12031
13.
Adamson
,
A. W.
,
1990
,
Physical Chemistry of Surfaces
,
John Wiley & Sons
, New York.
14.
Howe
,
J. M.
,
1997
,
Interfaces in Materials
,
John Wiley & Sons
, New York.
15.
Marichev
,
V. A.
,
2010
, “
General Thermodynamic Equations for the Surface Tension of Liquids and Solids
,”
Surf. Sci.
,
604
(
3–4
), pp.
458
463
.10.1016/j.susc.2009.12.020
16.
Rusanov
,
A. I.
,
2005
, “
Surface Thermodynamics Revisited
,”
Surf. Sci. Rep.
,
58
, pp.
111
239
.10.1016/j.surfrep.2005.08.002
17.
Rusanov
,
A. I.
,
1996
, “
Thermodynamics of Solid Surfaces
,”
Surf. Sci. Rep.
,
23
, pp.
173
247
.10.1016/0167-5729(95)00007-0
18.
Müller
,
P.
, and
Saul
,
A.
,
2004
, “
Elastic Effects on Surface Physics
,”
Surf. Sci. Rep.
,
54
(
5–8
), pp.
157
258
.10.1016/j.surfrep.2004.05.001
19.
Ibach
,
H.
,
1997
, “
The Role of Surface Stress in Reconstruction, Epitaxial Growth and Stabilization of Mesoscopic Structures
,”
Surf. Sci. Rep.
,
29
(
5–6
), pp.
195
263
.10.1016/S0167-5729(97)00010-1
20.
Leo
,
P. H.
, and
Sekerka
,
R. F.
,
1989
, “
Overview No. 86: The Effect of Surface Stress on Crystal-Melt and Crystal-Crystal Equilibrium
,”
Acta Metall. Mater.
,
37
(
12
), pp.
3119
3138
.10.1016/0001-6160(89)90184-3
21.
Cammarata
,
R. C.
,
1997
, “
Surface and Interface Stress Effects on Interfacial and Nanostructured Materials
,”
Mater. Sci. Eng. A,
,
237
(
2
), pp.
180
184
.10.1016/S0921-5093(97)00128-7
22.
Cammarata
,
R. C.
,
1994
, “
Surface and Interface Stress Effects in Thin Films
,”
Prog. Surf. Sci.
,
46
(
1
), pp.
1
38
.10.1016/0079-6816(94)90005-1
23.
Cammarata
,
R. C.
,
Sieradzki
,
K.
, and
Spaepen
,
F.
,
2000
, “
Simple Model for Interface Stresses With Application to Misfit Dislocation Generation in Epitaxial Thin Films
,”
J. Appl. Phys.
,
87
(
3
), pp.
1227
1234
.10.1063/1.372001
24.
Cammarata
,
R. C.
,
2009
, “
Generalized Thermodynamics of Surfaces With Applications to Small Solid Systems
,”
Solid State Phys.
,
61
, pp.
1
75
.10.1016/S0081-1947(09)00001-0
25.
Fischer
,
F. D.
,
Waitz
,
T.
,
Vollath
,
D.
, and
Simha
,
N. K.
,
2008
, “
On the Role of Surface Energy and Surface Stress in Phase-Transforming Nanoparticles
,”
Prog. Mater. Sci.
,
53
(
3
), pp.
481
527
.10.1016/j.pmatsci.2007.09.001
26.
Gutman
,
E. M.
,
1995
, “
On the Thermodynamic Definition of Surface Stress
,”
J. Phys. Condens. Mat.
,
7
(
48
), pp.
L663
L667
.10.1088/0953-8984/7/48/001
27.
Bottomley
,
D. J.
, and
Ogino
,
T.
,
2001
, “
Alternative to the Shuttleworth Formulation of Solid Surface Stress
,”
Phys. Rev. B
,
63
, p.
165412
.10.1103/PhysRevB.63.165412
28.
Marichev
,
V. A.
,
2006
, “
Structure-Mechanical Approach to Surface Tension of Solids
,”
Surf. Sci.
,
600
(
19
), pp.
4527
4536
.10.1016/j.susc.2006.07.018
29.
Kramer
,
D.
, and
Weissmüller
,
J.
,
2007
, “
A Note on Surface Stress and Surface Tension and Their Interrelation via Shuttleworth's Equation and the Lippmann Equation
,”
Surf. Sci.
,
601
(
14
), pp.
3042
3051
.10.1016/j.susc.2007.05.005
30.
Maugin
,
G. A.
,
2010
,
Configurational Forces: Thermomechanics, Physics, Mathematics, and Numerics
,
CRC Press
, Boca Raton, FL.
31.
Steinmann
,
P.
,
McBride
,
A.
,
Bargmann
,
S.
, and
Javili
,
A.
,
2012
, “
A Deformational and Configurational Framework for Geometrically Non-Linear Continuum Thermomechanics Coupled to Diffusion
,”
Int. J. Nonlinear Mech.
,
47
(
2
), pp.
215
227
.10.1016/j.ijnonlinmec.2011.05.009
32.
Gurtin
,
M. E.
, and
Murdoch
,
A. I.
,
1975
, “
A Continuum Theory of Elastic Material Surfaces
,”
Arch. Ration. Mech. An.
,
57
(
4
), pp.
291
323
.10.1007/BF00261375
33.
Gurtin
,
M. E.
, and
Murdoch
,
A. I.
,
1978
, “
Surface Stress in Solids
,”
Int. J. Solid. Struct.
,
14
(
6
), pp.
43
440
.10.1016/0020-7683(78)90008-2
34.
Dingreville
,
R.
, and
Qu
,
J.
,
2005
, “
Surface Free Energy and Its Effect on the Elastic Behavior of Nano-Sized Particles, Wires and Films
,”
J. Mech. Phys. Solid.
,
53
(
8
), pp.
1827
1854
.10.1016/j.jmps.2005.02.012
35.
He
,
J.
, and
Lilley
,
C. M.
,
2008
, “
Surface Effect on the Elastic Behavior of Static Bending Nanowires
,”
Nano Lett.
,
8
(
7
), pp.
1798
1802
.10.1021/nl0733233
36.
Duan
,
H. L.
,
Wang
,
J.
, and
Karihaloo
,
B. L.
,
2009
, “
Theory of Elasticity at the Nanoscale
,”
Adv. Appl. Mech.
,
42
, pp.
1
68
.10.1016/S0065-2156(08)00001-X
37.
Murdoch
,
A. I.
,
1976
, “
A Thermodynamical Theory of Elastic Material Interfaces
,”
Q. J. Mech. Appl. Math.
,
29
(
3
), pp.
245
275
.10.1093/qjmam/29.3.245
38.
Gurtin
,
M. E.
, and
Struthers
,
A.
,
1990
, “
Multiphase Thermomechanics With Interfacial Structure 3. Evolving Phase Boundaries in the Presence of Bulk Deformation
,”
Arch. Ration. Mech. An.
,
112
(
2
), pp.
97
160
.10.1007/BF00375667
39.
Gurtin
,
M. E.
,
Weissmüller
,
J.
, and
Larché
,
F.
,
1998
, “
A General Theory of Curved Deformable Interfaces in Solids at Equilibrium
,”
Philos. Mag. A
,
78
(
5
), pp.
1093
1109
.10.1080/01418619808239977
40.
Steigmann
,
D. J.
, and
Ogden
,
R. W.
,
1999
, “
Elastic Surface-Substrate Interactions
,”
Proc. R. Soc. A
,
455
(
1982
), pp.
437
474
.10.1098/rspa.1999.0320
41.
Fried
,
E.
, and
Todres
,
R.
,
2005
, “
Mind the Gap: The Shape of the Free Surface of a Rubber-Like Material in Proximity to a Rigid Contactor
,”
J. Elasticity
,
80
, pp.
97
151
.10.1007/s10659-005-9019-z
42.
Chhapadia
,
P.
,
Mohammadi
,
P.
, and
Sharma
,
P.
,
2011
, “
Curvature-Dependent Surface Energy and Implications for Nanostructures
,”
J. Mech. Phys. Solid.
,
59
(
10
), pp.
2103
2115
.10.1016/j.jmps.2011.06.007
43.
Moeckel
,
G. P.
,
1975
, “
Thermodynamics of an Interface
,”
Arch. Ration. Mech. An.
,
57
, pp.
255
280
.10.1007/BF00280158
44.
dell'Isola
,
F.
, and
Romano
,
A.
,
1987
, “
On the Derivation of Thermomechanical Balance Equations for Continuous Systems With a Nonmaterial Interface
,”
Int. J. Eng. Sci.
,
25
(
11–12
), pp.
1459
1468
.10.1016/0020-7225(87)90023-1
45.
Müller
,
I.
,
1971
, “
Entropy, Absolute Temperature, and Coldness in Thermodynamics
,”
Courses and Lectures—International Centre for Mechanical Sciences
,
Springer
,
New York
.
46.
Gogosov
,
V. V.
,
Naletova
,
V. A.
,
Bin
,
C. Z.
, and
Shaposhnikova
,
G. A.
,
1983
, “
Conservation Laws for the Mass, Momentum, and Energy on a Phase Interface for True and Excess Surface Parameters
,”
Fluid Dyn.
,
18
, pp.
923
930
.10.1007/BF01090749
47.
Daher
,
N.
, and
Maugin
,
G. A.
,
1986
, “
The Method of Virtual Power in Continuum Mechanics Application to Media Presenting Singular Surfaces and Interfaces
,”
Acta Mech.
,
60
(
3–4
), pp.
217
240
.10.1007/BF01176354
48.
Germain
,
P.
,
1973
, “
The Method of Virtual Power in Continuum Mechanics—Part 2: Microstructure
,”
SIAM J. Appl. Math.
,
25
, pp.
556
575
.10.1137/0125053
49.
Maugin
,
G. A.
,
1980
, “
The Method of Virtual Power in Continuum Mechanics: Application to Coupled Fields
,”
Acta Mech.
,
35
, pp.
1
70
.10.1007/BF01190057
50.
Daher
,
N.
, and
Maugin
,
G. A.
,
1987
, “
Deformable Semiconductors With Interfaces: Basic Continuum Equations
,”
Int. J. Eng. Sci.
,
25
(
9
), pp.
1093
1129
.10.1016/0020-7225(87)90076-0
51.
Daher
,
N.
, and
Maugin
,
G. A.
,
1986
, “
Virtual Power and Thermodynamics for Electromagnetic Continua With Interfaces
,”
J. Math. Phys.
,
27
, p.
3022
.10.1063/1.527231
52.
Murdoch
,
A. I.
,
2005
, “
Some Fundamental Aspects of Surface Modelling
,”
J. Elasticity
,
80
(
1
), pp.
33
52
.10.1007/s10659-005-9024-2
53.
Šilhavý
,
M.
,
2011
, “
Equilibrium of Phases With Interfacial Energy: A Variational Approach
,”
J. Elasticity
,
105
, pp.
271
303
.10.1007/s10659-011-9341-6
54.
Park
,
H. S.
,
Klein
,
P. A.
, and
Wagner
,
G. J.
,
2006
, “
A Surface Cauchy–Born Model for Nanoscale Materials
,”
Int. J. Num. Meth. Eng.
,
68
(
10
), pp.
1072
1095
.10.1002/nme.1754
55.
Park
,
H. S.
, and
Klein
,
P. A.
,
2008
, “
A Surface Cauchy–Born Model for Silicon Nanostructures
,”
Comput. Meth. Appl. Mech. Eng.
,
197
(
41–42
), pp.
3249
3260
.10.1016/j.cma.2007.12.004
56.
Park
,
H. S.
, and
Klein
,
P. A.
,
2007
, “
Surface Cauchy–Born Analysis of Surface Stress Effects on Metallic Nanowires
,”
Phys. Rev. B
,
75
(
8
), p.
085408
.10.1103/PhysRevB.75.085408
57.
Park
,
H. S.
,
Cai
,
W.
, and
Espinosa
,
H. D.
,
2009
, “
Mechanics of Crystalline Nanowires
,”
MRS Bulletin
,
34
(
3
), pp.
178
183
.10.1557/mrs2009.49
58.
Sharma
,
P.
,
Ganti
,
S.
, and
Bhate
,
N.
,
2003
, “
Effect of Surfaces on the Size-Dependent Elastic State of Nano-Inhomogeneities
,”
Appl. Phys. Lett.
,
82
(
4
), pp.
535
537
.10.1063/1.1539929
59.
Sharma
,
P.
, and
Ganti
,
S.
,
2004
, “
Size-Dependent Eshelby's Tensor for Embedded Nano-Inclusions Incorporating Surface/Interface Energies
,”
ASME J. Appl. Mech.
,
71
(5)
, pp.
663
671
.10.1115/1.1781177
60.
Sharma
,
P.
, and
Wheeler
,
L. T.
,
2007
, “
Size-Dependent Elastic State of Ellipsoidal Nano-Inclusions Incorporating Surface/Interface Tension
,”
ASME J. Appl. Mech.
,
74
(
3
), pp.
447
454
.10.1115/1.2338052
61.
Eshelby
,
J. D.
,
1951
, “
The Force on an Elastic Singularity
,”
Philos. Trans. R. Soc. A
,
244
(
877
), pp.
87
112
.10.1098/rsta.1951.0016
62.
Eshelby
,
J. D.
,
1957
, “
The Determination of the Elastic Field of an Ellipsoidal Inclusion, and Related Problems
,”
Proc. R. Soc. A
,
241
(
1226
), pp.
376
396
.10.1098/rspa.1957.0133
63.
Duan
,
H. L.
,
Wang
,
J.
,
Huang
,
Z. P.
, and
Karihaloo
,
B. L.
,
2005
, “
Eshelby Formalism for Nano-Inhomogeneities
,”
Proc. R. Soc. A
,
461
(
2062
), pp.
3335
3353
.10.1098/rspa.2005.1520
64.
Duan
,
H. L.
,
Wang
,
J.
,
Huang
,
Z. P.
, and
Karihaloo
,
B. L.
,
2005
, “
Size-Dependent Effective Elastic Constants of Solids Containing Nano-Inhomogeneities With Interface Stress
,”
J. Mech. Phys. Solid.
,
53
(
7
), pp.
1574
1596
.10.1016/j.jmps.2005.02.009
65.
Duan
,
H. L.
, and
Karihaloo
,
B. L.
,
2007
, “
Effective Thermal Conductivities of Heterogeneous Media Containing Multiple Imperfectly Bonded Inclusions
,”
Phys. Rev. B
,
75
, p.
064206
.10.1103/PhysRevB.75.064206
66.
Benveniste
,
Y.
, and
Miloh
,
T.
,
2001
, “
Imperfect Soft and Stiff Interfaces in Two-Dimensional Elasticity
,”
Mech. Mater.
,
33
(
6
), pp.
309
323
.10.1016/S0167-6636(01)00055-2
67.
Huang
,
Z.
, and
Sun
,
L.
,
2007
, “
Size-Dependent Effective Properties of a Heterogeneous Material With Interface Energy Effect: From Finite Deformation Theory to Infinitesimal Strain Analysis
,”
Acta Mech.
,
190
, pp.
151
163
.10.1007/s00707-006-0381-0
68.
Fischer
,
F. D.
, and
Svoboda
,
J.
,
2010
, “
Stresses in Hollow Nanoparticles
,”
Int. J. Solid. Struct.
,
47
(
20
), pp.
2799
2805
.10.1016/j.ijsolstr.2010.06.008
69.
Mogilevskaya
,
S. G.
,
Crouch
,
S. L.
, and
Stolarski
,
H. K.
,
2008
, “
Multiple Interacting Circular Nano-Inhomogeneities With Surface/Interface Effects
,”
J. Mech. Phys. Solid.
,
56
(
6
), pp.
2298
2327
.10.1016/j.jmps.2008.01.001
70.
Lim
,
C. W.
,
Li
,
Z. R.
, and
He
,
L. H.
,
2006
, “
Size Dependent, Non-Uniform Elastic Field Inside a Nano-Scale Spherical Inclusion Due to Interface Stress
,”
Int. J. Solid. Struct.
,
43
(
17
), pp.
5055
5065
.10.1016/j.ijsolstr.2005.08.007
71.
He
,
L. H.
, and
Li
,
Z. R.
,
2006
, “
Impact of Surface Stress on Stress Concentration
,”
Int. J. Solid. Struct.
,
43
(
20
), pp.
6208
6219
.10.1016/j.ijsolstr.2005.05.041
72.
Mi
,
C.
, and
Kouris
,
D. A.
,
2006
, “
Nanoparticles Under the Influence of Surface/Interface Elasticity
,”
J. Mech. Mater. Struct.
,
1
(
4
), pp.
763
791
.10.2140/jomms.2006.1.763
73.
Zöllner
,
A. M.
,
Buganza Tepole
,
A.
, and
Kuhl
,
E.
,
2012
, “
On the Biomechanics and Mechanobiology of Growing Skin
,”
J. Theoret. Biol.
,
297
, pp.
166
175
.10.1016/j.jtbi.2011.12.022
74.
Sun
,
C. Q.
,
2009
, “
Thermo-Mechanical Behavior of Low-Dimensional Systems: The Local Bond Average Approach
,”
Prog. Mater. Sci.
,
54
(
2
), pp.
179
307
.10.1016/j.pmatsci.2008.08.001
75.
Johnson
,
W. C.
,
2000
, “
Superficial Stress and Strain at Coherent Interfaces
,”
Acta Mater.
,
48
(
2
), pp.
433
444
.10.1016/S1359-6454(99)00359-6
76.
Yvonnet
,
J.
,
Mitrushchenkov
,
A.
,
Chambaud
,
G.
, and
He
,
Q.-C.
,
2011
, “
Finite Element Model of Ionic Nanowires With Size-Dependent Mechanical Properties Determined by Ab Initio Calculations
,”
Comput. Meth. Appl. Mech. Eng.
,
200
(5-8)
, pp.
614
625
.10.1016/j.cma.2010.09.007
77.
McBride
,
A.
,
Javili
,
A.
,
Steinmann
,
P.
, and
Bargmann
,
S.
,
2011
, “
Geometrically Nonlinear Continuum Thermomechanics With Surface Energies Coupled to Diffusion
,”
J. Mech. Phys. Solid.
,
59
(
10
), pp.
2116
2133
.10.1016/j.jmps.2011.06.002
78.
Özdemir
,
I.
,
Brekelmans
,
W. A. M.
, and
Geers
,
M. G. D.
,
2010
, “
A Thermo-Mechanical Cohesive Zone Model
,”
Computat. Mech.
,
46
(
5
), pp.
735
745
.10.1007/s00466-010-0507-z
79.
Benveniste
,
Y.
,
2006
, “
A General Interface Model for a Three-Dimensional Curved Thin Anisotropic Interphase Between Two Anisotropic Media
,”
J. Mech. Phys. Solid.
,
54
(
4
), pp.
708
734
.10.1016/j.jmps.2005.10.009
80.
Le-Quang
,
H.
,
Bonnet
,
G.
, and
He
,
Q.-C.
,
2010
, “
Size-Dependent Eshelby Tensor Fields and Effective Conductivity of Composites Made of Anisotropic Phases With Highly Conducting Imperfect Interfaces
,”
Phys. Rev. B
,
81
(
6
), p.
064203
.10.1103/PhysRevB.81.064203
81.
Gu
,
S. T.
, and
He
,
Q.-C.
,
2011
, “
Interfacial Discontinuity Relations for Coupled Multifield Phenomena and Their Application to the Modeling of Thin Interphases as Imperfect Interfaces
,”
J. Mech. Phys. Solid.
,
59
(
7
), pp.
1413
1426
.10.1016/j.jmps.2011.04.004
82.
Kapitza
,
P. L.
,
1941
, “
The Study of Heat Transfer in Helium II
,”
J. Phys. (USSR)
,
4
, pp.
181
210
.
83.
Yvonnet
,
J.
,
He
,
Q.-C.
,
Zhu
,
Q.-Z.
, and
Shao
,
J. F.
,
2011
, “
A General and Efficient Computational Procedure for Modelling the Kapitza Thermal Resistance Based on XFEM
,”
Comput. Mater. Sci.
,
50
(
4
), pp.
1220
1224
.10.1016/j.commatsci.2010.02.040
84.
Yvonnet
,
J.
,
He
,
Q.-C.
, and
Toulemonde
,
C.
,
2008
, “
Numerical Modelling of the Effective Conductivities of Composites With Arbitrarily Shaped Inclusions and Highly Conducting Interface
,”
Composit. Sci. Tech.
,
68
(
13
), pp.
2818
2825
.10.1016/j.compscitech.2008.06.008
85.
Miloh
,
T.
, and
Benveniste
,
Y.
,
1999
, “
On the Effective Conductivity of Composites With Ellipsoidal Inhomogeneities and Highly Conducting Interfaces
,”
Proc. R. Soc. A
,
455
(
1987
), pp.
2687
2706
.10.1098/rspa.1999.0422
86.
Javili
,
A.
,
McBride
,
A.
, and
Steinmann
,
P.
,
2013
, “
Numerical Modelling of Thermomechanical Solids With Highly Conductive Energetic Interfaces
,”
Int. J. Numer. Meth. Eng.
,
93
(5)
, pp.
551
574
.10.1002/nme.4402
87.
Mosler
,
J.
, and
Scheider
,
I.
,
2011
, “
A Thermodynamically and Variationally Consistent Class of Damage-Type Cohesive Models
,” Technical Report.
88.
Steinmann
,
P.
, and
Häsner
,
O.
,
2005
, “
On Material Interfaces in Thermomechanical Solids
,”
Arch. Appl. Mech.
,
75
(
1
), pp.
31
41
.10.1007/s00419-005-0383-8
89.
Steinmann
,
P.
,
2008
, “
On Boundary Potential Energies in Deformational and Configurational Mechanics
,”
J. Mech. Phys. Solid.
,
56
(
3
), pp.
772
800
.10.1016/j.jmps.2007.07.001
90.
Javili
,
A.
, and
Steinmann
,
P.
,
2009
, “
A Finite Element Framework for Continua With Boundary Energies—Part I: The Two-Dimensional Case
,”
Comput. Meth. Appl. Mech. Eng.
,
198
(
27–29
), pp.
2198
2208
.10.1016/j.cma.2009.02.008
91.
Javili
,
A.
, and
Steinmann
,
P.
,
2010
, “
A Finite Element Framework for Continua With Boundary Energies—Part II: The Three-Dimensional Case
,”
Comput. Meth. Appl. Mech. Eng.
,
199
(
9–12
), pp.
755
765
.10.1016/j.cma.2009.11.003
92.
Javili
,
A.
, and
Steinmann
,
P.
,
2010
, “
On Thermomechanical Solids With Boundary Structures
,”
Int. J. Solid. Struct.
,
47
(
24
), pp.
3245
3253
.10.1016/j.ijsolstr.2010.08.009
93.
Javili
,
A.
, and
Steinmann
,
P.
,
2011
, “
A Finite Element Framework for Continua With Boundary Energies—Part III: The Thermomechanical Case
,”
Comput. Meth. Appl. Mech. Eng.
,
200
(
21–22
), pp.
1963
1977
.10.1016/j.cma.2010.12.013
94.
Gough
,
J.
,
1805
, “
A Description of a Property of Caoutchouc on Indian Rubber; With Some Reflections on the Case of the Elasticity of This Substance
,”
Memoirs of the Literary and Philosophical Society of Manchester
, Vol. 1, pp.
288
295
.
95.
Joule
,
J. P.
,
1859
, “
On Some Thermo-Dynamic Properties of Solids
,”
Philos. Trans. R. Soc. A
,
149
, pp.
91
131
.10.1098/rstl.1859.0005
96.
Marsden
,
J. E.
, and
Hughes
,
T. J. R.
,
1994
,
Mathematical Foundations of Elasticity
,
Dover Publications
,
New York
.
97.
Šilhavý
,
M.
,
1997
,
The Mechanics and Thermodynamics of Continuous Media
,
Springer
,
Berlin, Germany
.
98.
Truesdell
,
C.
, and
Noll
,
W.
,
2004
,
The Non-Linear Field Theories of Mechanics
,
Springer
,
Berlin, Germany
.
99.
Gurtin
,
M. E.
,
Fried
,
E.
, and
Anand
,
L.
,
2009
,
The Mechanics and Thermodynamics of Continua
,
Cambridge University Press
,
New York
.
100.
Holzapfel
,
G. A.
,
2000
,
Nonlinear Solid Mechanics: A Continuum Approach for Engineering
,
John Wiley & Sons
,
Chichester, UK
.
101.
Besson
,
J.
,
Cailletaud
,
G.
,
Chaboche
,
J. L.
, and
Forest
,
S.
,
2010
,
Non-Linear Mechanics of Materials
,
Springer
,
Heidelberg, Germany
.
102.
Bowen
,
R. M.
, and
Wang
,
C. C.
,
1976
,
Introduction to Vectors and Tensors: Linear and Multilinear Algebra
,
Plenum Press, New York
.
103.
Kreyszig
,
E.
,
1991
,
Differential Geometry
,
Dover Publications
,
New York
.
104.
Ciarlet
,
P. G.
,
2005
,
An Introduction to Differential Geometry With Applications to Elasticity
,
Springer
,
The Netherlands
.
105.
Gurtin
,
M. E.
,
2000
,
Configurational Forces as Basic Concepts of Continuum Physics
,
Springer
, New York.
106.
Eringen
,
A. C.
,
1967
,
Mechanics of Continua
,
Wiley
,
New York
.
107.
Hutter
,
K.
,
1977
, “
The Foundations of Thermodynamics, Its Basic Postulates and Implications: A Review of Modern Thermodynamics
,”
Acta Mech.
,
27
, pp.
1
54
.10.1007/BF01180075
108.
Müller
,
I.
,
1967
, “
On the Entropy Inequality
,”
Arch. Ration. Mech. An.
,
26
, pp.
118
141
.10.1007/BF00285677
109.
Simha
,
N. K.
, and
Bhattacharya
,
K.
,
2000
, “
Kinetics of Phase Boundaries With Edges and Junctions in a Three-Dimensional Multi-Phase Body
,”
J. Mech. Phys. Solid.
,
48
(
12
), pp.
2619
2641
.10.1016/S0022-5096(00)00008-9
110.
Miehe
,
C.
,
1995
, “
Entropic Thermoelasticity at Finite Strains. Aspects of the Formulation and Numerical Implementation
,”
Comput. Meth. Appl. Mech. Eng.
,
120
(
3–4
), pp.
243
269
.10.1016/0045-7825(94)00057-T
111.
Javili
,
A.
,
McBride
,
A.
, and
Steinmann
,
P.
,
2012
, “
Numerical Modelling of Thermomechanical Solids With Mechanically Energetic (Generalised) Kapitza Interfaces
,”
Computat. Mater. Sci.
,
65
, pp.
542
551
.10.1016/j.commatsci.2012.06.006
112.
Ye
,
J. C.
,
Lu
,
J.
,
Liu
,
C. T.
,
Wang
,
Q.
, and
Yang
,
Y.
,
2010
, “
Atomistic Free-Volume Zones and Inelastic Deformation of Metallic Glasses
,”
Nature Mater.
,
9
(
8
), pp.
619
623
.10.1038/nmat2802
113.
Javili
,
A.
,
McBride
,
A.
,
Steinmann
,
P.
, and
Reddy
,
B. D.
,
2012
, “
Relationships Between the Admissible Range of Surface Material Parameters and Stability of Linearly Elastic Bodies
,”
Philosoph. Mag.
,
92
(
28–30
), pp.
3540
3563
.10.1080/14786435.2012.682175
114.
Ciarlet
,
P. G.
,
1988
,
Mathematical Elasticity
, Vol.
1
,
Elsevier, Amsterdam, The Netherlands
.
115.
Ogden
,
R. W.
,
1997
,
Non-Linear Elastic Deformations
,
Dover Publications
, New York.
116.
Gurtin
,
M. E.
, and
Murdoch
,
A. I.
,
1975
, “
Addenda to Our Paper a Continuum Theory of Elastic Material Surfaces
,”
Arch. Ration. Mech. An.
,
59
(
4
), pp.
1
2
.10.1007/BF00281513
117.
Ru
,
C. Q.
,
2010
, “
Simple Geometrical Explanation of Gurtin–Murdoch Model of Surface Elasticity With Clarification of Its Related Versions
,”
Sci. China, Phys. Mech. Astron.
,
53
(3)
, pp.
536
544
.10.1007/s11433-010-0144-8
118.
Haiss
,
W.
,
2001
, “
Surface Stress of Clean and Adsorbate-Covered Solids
,”
Rep. Progress Phys.
,
64
(
5
), pp.
591
648
.10.1088/0034-4885/64/5/201
119.
Dingreville
,
R.
, and
Qu
,
J.
,
2007
, “
A Semi-Analytical Method to Compute Surface Elastic Properties
,”
Acta Mater.
,
55
(
1
), pp.
141
147
.10.1016/j.actamat.2006.08.007
120.
Malvern
,
L. E.
,
1969
,
Introduction to the Mechanics of a Continuous Medium
,
Prentice-Hall
,
Englewood Cliffs, NJ
.
121.
Huang
,
Z. P.
, and
Wang
,
J.
,
2006
, “
A Theory of Hyperelasticity of Multi-Phase Media With Surface/Interface Energy Effect
,”
Acta Mech.
,
182
(
3-4
), pp.
195
210
.10.1007/s00707-005-0286-3
122.
Miller
,
R. E.
, and
Shenoy
,
V. B.
,
2000
, “
Size-Dependent Elastic Properties of Nanosized Structural Elements
,”
Nanotechnology
,
11
(
3
), p.
139
.10.1088/0957-4484/11/3/301
123.
Wei
,
G. W.
,
Shouwen
,
Y. U.
, and
Ganyun
,
H.
,
2006
, “
Finite Element Characterization of the Size-Dependent Mechanical Behaviour in Nanosystems
,”
Nanotechnology
,
17
(
4
), pp.
1118
1122
.10.1088/0957-4484/17/4/045
124.
Hill
,
R.
,
1958
, “
A General Theory of Uniqueness and Stability in Elastic-Plastic Solids
,”
J. Mech. Phys. Solid.
,
6
(
3
), pp.
236
249
.10.1016/0022-5096(58)90029-2
125.
Simpson
,
H. C.
, and
Spector
,
S. J.
,
1985
, “
On Failure of the Complementing Condition and Nonuniqueness in Linear Elastostatics
,”
J. Elasticity
,
15
, pp.
229
231
.10.1007/BF00041996
126.
Benallal
,
A.
,
Billardon
,
R.
, and
Geymonat
,
G.
,
1993
, “
Bifurcation and Localization in Rate-Independent Materials. Some General Considerations
,”
Bifurcation and Stability of Dissipative Systems
(CISM Courses and Lectures, Vol.
327
),
Springer
,
Berlin, Germany
, pp.
1
44
.
127.
Reddy
,
B. D.
,
1982
, “
Surface Instabilities on an Equibiaxially Stretched Elastic Half-Space
,”
Math. Proc. Cambridge Phil. Soc.
,
91
, pp.
491
501
.10.1017/S0305004100059569
128.
Reddy
,
B. D.
,
1983
, “
The Occurrence of Surface Instabilities and Shear Bands in Plane Strain Deformation of an Elastic Half-Space
,”
Q. J. Mech. Appl. Math.
,
36
, pp.
337
350
.10.1093/qjmam/36.3.337
129.
Bigoni
,
D.
,
Ortiz
,
M.
, and
Needleman
,
A.
,
1997
, “
Effect of Interfacial Compliance on Bifurcation of a Layer Bonded to a Substrate
,”
Int. J. Solid. Struct.
,
34
(
97
), pp.
4305
4326
.10.1016/S0020-7683(97)00025-5
130.
Han
,
W.
, and
Reddy
,
B. D.
,
1999
,
Plasticity: Mathematical Theory and Numerical Analysis
, Vol.
9
,
Springer
,
New York
.
131.
Bigoni
,
D.
, and
Gei
,
M.
,
2001
, “
Bifurcations of a Coated, Elastic Cylinder
,”
Int. J. Solid. Struct.
,
38
(
30–31
), pp.
5117
5148
.10.1016/S0020-7683(00)00322-X
132.
Steigmann
,
D. J.
,
2010
, “
Elastic Waves Interacting With a Thin, Prestressed, Fiber-Reinforced Surface Film
,”
Int. J. Eng. Sci.
,
48
(
11
), pp.
1604
1609
.10.1016/j.ijengsci.2010.06.032
133.
Steigmann
,
D. J.
,
2012
, “
Refined Theory for Linearly Elastic Plates: Laminae and Laminates
,”
Math. Mech. Solid
,
17
(4)
, pp.
351
363
.10.1177/1081286511419971
134.
Shenoy
,
V. B.
,
2005
, “
Atomistic Calculations of Elastic Properties of Metallic FCC Crystal Surfaces
,”
Phys. Rev. B
,
71
(
9
), p.
094104
.10.1103/PhysRevB.71.094104
135.
Forest
,
S.
,
Cordero
,
N. M.
, and
Busso
,
E. P.
,
2011
, “
First vs. Second Gradient of Strain Theory for Capillarity Effects in an Elastic Fluid at Small Length Scales
,”
Computat. Mater. Sci.
,
50
(
4
), pp.
1299
1304
.10.1016/j.commatsci.2010.03.048
136.
Gurtin
,
M. E.
, and
Jabbour
,
M. E.
,
2002
, “
Interface Evolution in Three Dimensions With Curvature-Dependent Energy and Surface Diffusion: Interface-Controlled Evolution, Phase Transitions, Epitaxial Growth of Elastic Films
,”
Arch. Ration. Mech. An.
,
163
, pp.
171
208
.10.1007/s002050200193