800} This is to certify that the dissertation entitled Constitutive Modeling of the Thermal Response of Rubber-like Materials presented by Yuhui Wang has been accepted towards fulfillment of the requirements for the PhD degree in Mechanical Engineering MSU is an Affirmative Action/Equal Opportunity Institution LIBRARY Michigan State University - a--.----o-o-c-u-u-o-v-o-----—-v PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 2/05 p:/ClRC/DateDue.indd-p.1 CONSTITUTIVE MODELING OF THE THERMAL RESPONSE OF RUBBER-LIKE MATERIALS By Yuhui Wang A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mechanical Engineering 2006 ABSTRACT CONSTITUTIVE MODELING OF THE THERMAL RESPONSE OF RUBBER-LIKE MATERIALS By Yuhui Wang The thermal conductivity and diffusivity of elastomers can change due to finite deformation. Some studies have proposed the constitutive relationship between thermal conduction and finite deformation for uniaxial loading of specimens. These correlations often omit the values of the components of the diffusivity and ignore the results of continuum mechanics in developing the constitutive relation. The present study focuses on two means of correlating the diffusivity of elastomers with respect to finite deformation. First, measured diffusivities fi'om other studies are correlated with deformation, through the principal stretches or the Cauchy- Green strains, in the current (measurement) frame and afier transforming the diffusivity to the reference (undeformed) frame. The change in each component of the diffusivity tensor appears to be determined by the deformation in the direction of the component only. The correlation in the current frame suggests separate fitting parameters for each direction of each material. In the reference frame, the two directions seem to follow a similar trend. In the second part, advantage is taken of the constraints introduced by the isotropy that elastomeric samples often exhibit in their undefonned state. This material symmetry gives rise to mathematical restrictions on the constitutive equations. By satisfying the principle of material frame indifference and the second law of thermodynamics, and putting the constitutive function in an undeforrned reference configuration, a detailed form of the heat flux vector function is derived as a fimction of the pure stretching tensor. Parameters are determined for several elastomers that have wide use and ofien undergo large deformation. TO my family iv ACKNOWLEDGMENTS I would like to thank my advisor, Dr. Neil T. Wright, for his support and help throughout the study. It was my pleasure to conduct this interesting and challenging research under his guidance. I am also very grateful to my academic committee members, Dr. Andre Benard, Dr. Dahsin Liu, and Dr. Baisheng Yan, for their valuable advice and inspiring comments. Many of my colleagues have contributions to my dissertation. I would like to thank Xianglan Bai, Xing Huang, and Yuwei Chi for their insightful comments and sugges- tions. Finally, I am especially indebted to my husband, Hongbo Zhou, who has always been supportive. I am very lucky to have him in my life. I also would like to thank my lovely daughter, Julina Zhou, for bringing me so much pleasure since she was born. LIST OF TABLES LIST OF FIGURES 1 Introduction TABLE OF CONTENTS 1.1 Background .................................. 1.2 Helmholtz free energy ............................ 1.2.1 The effects of deformation on 11) ...................... 1.2.2 The effects Of temperature on 1,0 ...................... 1.3 Heat flux vector q .............................. 1.3.1 Thermal diffusivity ............................. 1.3.2 The effect of deformation on K or a ................... 1.3.3 The effects of temperature on K or a .................. 2 Related work 2.1 Experimental work .............................. 2.1.1 Uniaxial stretching ............................. 2.1.2 Multiaxial stretching ............................ 2.2 Existing models for k or a .......................... 2.2.1 Temperature dependence .......................... 2.2.2 Deformation dependence .......................... 3 Theoretical Framework 3.1 Choice of reference configuration ...................... 3.2 Relationship of q between configurations .................. 3.3 Material flame Indifference ......................... 3.4 Initially isotropic material .......................... 4 A Correlation of Diffusivity 4.1 Uniaxial Loading ............................... 4.2 Biaxial Loading ................................ 4.3 Discussion ................................... 5 A Constitutive Relation for q 5.1 Methods .................................... 5.1.1 Material frame indifference ........................ 5.1.2 Three linearly independent variables ................... 5.1.3 Application of second law of thermodynamics .............. 5.1.4 Material Symmetry ............................. 5.2 Analysis .................................... 5.2.1 Heat flux vector in terms of U ....................... 5.2.2 A special property of 2,111, «in, and 1123 ................... vi 56 57 57 57 58 5.3 Invariants ................................... 61 5.4 Data Analysis ................................. 62 5.4.1 Analysis of the uniaxial stretching of elastomers ............. 63 5.4.2 Analysis of biaxially stretched elastomers ................. 66 6 Discussion 69 APPENDICES 75 vii 4.1 4.2 4.3 5.1 LIST OF TABLES Polymers subject to uniaxial loading correlated with any,- / (134 = 0.03. . . 46 RTV polyurethane subject to biaxial loading duh/am = an- ...... 48 Correlating coefficients for equation 4.8 for neoprene rubber that is anisotropic in the unloaded condition and then is subject to biaxial loading . . . 49 Polymers subject to biaxial loading with form a/aeq = —(z,/)1I +¢2U"1+ «mu-2) .................................. 68 viii 1.1 1.2 2.1 2.2 3.1 3.2 4.1 4.2 5.1 5.2 LIST OF FIGURES A solid undergoing a thermomechanical process from reference configura- tion Bo to current configuration B ................... Typical force-extension curve for vulcanized rubber ............ The relationship between diffusivity and stretch ratio in the direction of stretching of silicon rubber subject to uniaxial loading [2]. The error bars denote i7.5% uncertainty in the stretch ratio ........... The normalized components of thermal diffusivity of silicone rubber as functions of their respective stretch ratios. Note the differing Slope indicating that a single parameter is insufficient ............. Observer transformation diagram describing Material Frame Indifference Two reference configurations related by a rigid body motion ....... Variation of aQfi/aeq with respect to C“- for silicone rubber subject to uniaxial loading. ............................. Variation (aofi-i/aeq) as a function of 0;,- for polyurethane subject to bi- axial loading. Specimens A1, A2, and A3 are shown. ......... Normalized thermal conductivity tensor as a function of the pure stretch- ing tensor for silicone .......................... Normalized thermal conductivity tensor as a function of pure stretching tensor for polyurethane RTV ....................... ix [\D 16 23 30 36 44 47 65 67 Chapter 1 Introduction Elastomers are rubber-like materials that are widely used in the automotive, aerospace, and medical industries in conditions of severe thermal and mechanical stress that often involve large temperature gradients and deformation. Natural rubber, for ex- ample, has excellent resistance to cutting, gouging, and abrasion. Another elastomer, synthetic rubber or styrene-butadience copolymer (SBR), is used predominantly in automobile tires because of its excellent abrasion resistance. Silicone rubber, another elastomer, possesses a high degree of flexibility at temperatures as low as —90°C and yet is stable to temperatures as high as 250°C, and thus can be used for high or low temperature applications (insulation, seals, diaphragms, or tubing). Elastomers are expected to provide trouble-free performance during their service life. Designs in— corporating elastomers and other engineered products rely increasingly on computer models of the thermomechanical performance, which leads to the need of constitutive relationships for their response to varies boundary conditions. 1. 1 Background A general theory of characterizing the thermomechanical response Of solid materials has long been available [5]. The deformation of a solid may be described as in Figure 1.1. Ci 33 Figure 1.1: A solid undergoing a thermomechanical process from reference configura- tion Bo to current configuration B Based on the conservation laws and the second law of thermodynamics, the general theory reveals that five constitutive functions are needed to describe the response of a solid material. These functions may include the internal energy function (5, specific internal entropy function 7“), Cauchy stress function T, heat flux vector function 6;, and internal state vector function ,6. These constitutive functions depend on the material point X, deformation gradient tensor F = 8x/0X, where x is the spatial position vector that the material point X occupies in the current configuration, temperature T, temperature gradient VOT = BT/BX, and internal state vector ,6 = (61,62, mflN), where m, 62"", H N are internal state variables. Since we are considering homogeneous 2 materials, the material point X can be omitted for convenience. The constitutive functions are written as e = é(F,T, VT, 3) (1.1) n = 17(F, T, VT, 5) (1.2) T = T(F, T, VT, 3) (1.3) q = 61(15‘, T, VT, [3) (1.4) B = 3(F,T, VT, 3) (1.5) The superposed hat serves to distinguish these functions from their values. The internal state vector 3 describes aspects of the internal structure of the material associated with irreversible or dissipative effects. The evolution of internal variables accounts for, indirectly, the effects of the history of the deformation. Because rubber- like materials demonstrate reversible processes during deformation, 3 is unneeded as a constitutive function for rubber-like materials with some exceptions; the other four constitutive functions 53, 1?, T, 5; will be considered independent of B for such a material response. Furthermore, the constitutive functions for rubber-like materials exhibiting hyperelastic behavior can be further simplified and only two independent constitutive functions, u and q, where 7,!) = 6 — T17 is defined as Helmholtz free energy, are needed. As shown in the Appendix, these have the following forms w = in: T) (1.6) q = (1(P, T, VT) (1.7) The other two constitutive functions can be expressed in terms of if and q as n = filET) = —ziT(F,T> (1.8) T = T(F, T) = §Ftl3g~‘(F, T) (1.9) where J is the determinant Of the deformation tensor J = detF, the subscript denotes the partial differentiation, and the superscript T denotes the transpose. Equations 1.6 and 1.7 demonstrate that the modeling of a thermomechanical response of elas- tomers requires constitutive models that account for both the mechanical and thermal dependence of the thermomechanical properties of these materials. 1.2 Helmholtz free energy 1.2.1 The effects Of deformation on w Experiments have revealed that macroscopic deformations imposed on elastomeric materials lead to molecular reorientation [30], which can induce changes in mechan- ical and thermophysical properties [2]. Measurements have demonstrated that there can be significant changes in the thermal conductivity or diffusivity tensors of elas- tomers subject to finite deformation [2, 22]. Additionally, the constitutive equation 1.6 shows that deformation determines the value of 1]) through F. Many studies have considered, experimentally and theoretically, the relationship between F and ii. A widely accepted form of 21) is still a matter of active study, although a number of forms have been proposed (see reviews by, for example, 'D'eloar [30] and Holzapfel [19]). The models of ‘t/} have been developed based on either a statistical or a phenomenological approach. The phenomenological method describes materials as continua. This methods fits mathematical equations to experimental data, which have been developed second law of thermodynamics, and material frame indifference. In studying of constitutive function if, the phenomenological method has led to, for example, the neo—Hookean, Mooney-Rivlin, Rivlin-Saunders, and Ogden models [19]. These models describe the mechanical response of elastomers for various boundary conditions or levels of defor- mation. Each of these models fits well experimental data under certain conditions. Most studies have focused on isothermal deformation, with less attention being de- voted to the effects of temperature on 1,0. Some exceptions include Ogden [25], who proposed a method for finding w as a function of biaxial stretches and temperature, and Humphrey and Rajagopal [20], who showed that in-plane biaxial tests allow measurement of thermoelastic response fimctions similar to the isothermal results of Rivlin and Saunders [27]. Although the phenomenological method is an effective way to develop models, it fails to capture the relationship between the mechanism of deformation and the underlying molecular structure of the material. In contrast, the statistical method has been recently used to offer interpretation of constitutive models in terms of mole- cular changes. E‘ied [13], for example, gave an elementary molecular explanation for each of the Mooney and Rivlin-Saunders theories of rubber deformation. Moreover, Wineman developed a continuum mechanics model that considered a development of a secondary molecular network, which results from microstructural change due to elevated temperature and deformation [37, 38]. In contrast to the phenomenological method, the molecular-statistical method may be used to develop constitutive models based on the molecular structural of materials. A comprehensive review of the several studies done using this method was given by Treloar [30]. Most studies consider the unique molecular structure of rubber-like materials, and by analyzing the features of the conformational change of the molecular network of the material under deformation, Gaussian or non-Gaussian models are developed. The Gaussian model correctly describes the behavior of real rubber to a first approximation only, within a very limited range of deformation, for stretch ratio up to 5.5 during uniaxial stretch. The Gaussian model is equivalent to the neO-Hookean model, which has been developed using the phenomenological approach. Tensile force (N mm") M l o lllllllllllll ; 0 2m 400 an WW) Figure 1.2: Typical force-extension curve for vulcanized rubber By extension of the statistical approach, three-chain and four-chain network models have been developed based on non-Gaussian theory [30]. The non-Gaussian theory gives an improved representation of the general features of the experimental response as compared to the Gaussian theory; the later entirely fails to represent the point of inflection, where the concave curve changes, and subsequent rapidly increasing slope at high strains, a typical force-deformation curve for a rubber is demonstrated as in figure 1.2 [30]. Arruda and Boyce developed an eight-chain molecular model [1]. This statistical-molecular method used insight Of the chemical structure of elastomers to model the mechanical response. 1.2.2 The effects of temperature on U) Compared to the study of F on w, few studies have examined the effects of T on 1/2, focusing instead on the isothermal response. A notable exception is the study of mate- rial behavior at elevated temperature by Wineman and Min [38]. Their study is based on the experimental work of Tobolsky [29] whose experimental results demonstrate that the molecular network changed when the temperature becomes high enough. Wineman and Min demonstrate the importance of temperature on the mechanical re- sponse of elastomers in studies of the development of a secondary molecular network resulting from microstructural changes due to deformation at elevated temperature. This secondary networks form due to consists of scission and subsequent cross-linking of the macromolecular network bonds. At high temperature, say greater than the temperature of onset of the ’chemorheological range’, the affected molecules will re- coil and cross-link to form a new network. This microstructural change results in a change in the mechanical response and leads to permanent set. Wineman and Min present a constitutive model based on this combination of two, or possibly more, microstructural networks. 1.3 Heat flux vector q A phenomenological law, based on experimental Observations, that relates the heat flux vector q to the spatial temperature gradient VT is asserted as Duhamel’s law of heat conduction [19] q = —K(F,T)VT (1.10) where K(F, T) is the spatial thermal conductivity tensor written in the current con- figuration. If the material is thermally isotropic, which means there is no preferred direction for the heat conduction, the thermal conductivity tensor becomes K = H, where k is a scalar, and equation 1.10 simplifies to q = —IcVT (1.11) which is called Fourier’s law of heat conduction, and k denotes the coefficient of thermal conductivity. Substituting equation 1.10 into the first law of thermodynamics for an incompress- ible solid without internal motion or energy generation yields mpg—31:: = V - (KVT) (1.12) and for isotropic conduction papal: = V - (kVT) (1.13) where p is the material density at current configuration and 0,, the specific heat capacity at constant temperature. 1 .3.1 Thermal diffusivity Thermal diffusivity is another material parameter that arises for transient heat-flow calculations. Some experiments measure the thermal diffusivity instead of the ther- mal conductivity of elastomers because the thermal diffusivity can be measured via transient experiments. Here, we demonstrate the relationship between the diffusivity and conductivity so that experimental results may be compared. We denote a to be the Spatial thermal diffusivity tensor, which is related to K by Here we will assume that elastomers have no change of volume upon deformation, which has been well-justified on the basis of experiments that show volume changes on deformation to be very small, i.e. on the order 10"4 or less [30]. This assumption is a common idealization invoked in continuum and computational mechanics, and leads to the result that detF = 1. Therefore, p = p(), Since p = podetF, where p0 is the material density described in referential configuration. In addition, Broerman et a1. [2] deduce that op is independent Of stretch ratio in their uniaxial loading experiment based on the assumption that the tensile force is a linear function of temperature as proposed by Treloar [30]. Therefore, Similar relations hold between the deformation and a and between deformation and K. 10 1.3.2 The effect of deformation on K or a Heat conduction in elastomers changes substantially due to deformation. When amor- phous polymers or rubber-like materials are stretched uniaxially, these original ther— mally isotropic materials become anisotropic with increased conduction in the direc- tion of stretching, which has been measured in several studies by Broerman et al. [2], Washo [36], Tautz [28], Hennig [18], and Hands and Horsfall [16]. This property change results from the unique molecular structure of elastomers. Elastomers consist of long-chain molecules that are randomly oriented in their native state. When the material is subject to a uniaxial deformation, the molecular chains are oriented pref- erentially along the stretching direction. The thermal conductivity of the oriented material will be greater along the back bone of the chain with its strong covalent bonds, and weaker valences parallel to the chain axis, and smaller perpendicular to the stretching due to the fewer van der Vaals bonds. Therefore, the orientation causes a preference for heat conduction in direction of stretching and the material demonstrates anisotropic thermal conduction. In order to understand better the changes in thermophysical properties of elas- tomers in response to deformation, some experiments have been conducted to measure the changes in heat conduction due to deformation [36, 18, 16]. Broerman et a1. [2] measured the thermal diffusivity of silicone rubber subject to uniaxial loading using a thermo-optical technique. They reported that for initially isotropic silicone rub- ber, the component of the thermal diffusivity tensor in the direction of stretching increases by 10% for a stretch ratio of two, and the component of the diffusivity or- thogonal to the direction of stretch decreases by 5%. This agrees qualitatively with 11 previous studies of LeGall. LeGall [23] studied multiaxial loading cases, where they measured the three orthogonal components of diffusivity during biaxial stretching of polyurethane, natural gum, and neoprene rubbers. The deformation was homo- geneous in the central region of the specimen, where the diffusivity was measured. The results showed an increased diffusivity in the direction of stretching and a de- creased value in the orthogonal direction, in the out of plane direction. LeGall and Wright did not provide a correlation of these data. Wang and Wright [35] studied both the uniaxial loading and biaxial loading experimental data for initially isotropic elastomers. By transforming those data to the respective reference configurations, where the elastomers demonstrate isotropic thermal property, they found that the thermal diffusivity or thermal conductivity is proportional to the inverse of Right Cauchy deformation tensor C = FTF. Within the expectation, the coefficients are different for different materials. Although a few experiments have been conducted related to this area and some observations have been made, no complete and rigorous mathematic theory has been advanced. 1.3.3 The effects of temperature on K or a Most properties of elastomers depend heavily level on temperature. For example, the mechanical response changes greatly at the glass transition temperature [6]. At a tem- peratures below the glass transition temperature, the molecular structure is random and amorphous, but the mechanical behavior is unlike a rubber and these materials become hard and rigid like a glass. Above the glass transition temperature, the mate- 12 rial are more compliant. The thermal conductivity also shows significant dependence on temperature when elastomers are either above or below the glass transition tem- perature [16, 6, 9, 10, 11, 12, 7]. These studies have assumed an isotropic thermal response of elastomers. In addition to the experimental studies, some theoretical stud- ies have offered quantitative relations between the thermal conductivity or thermal diffusivity and temperature, and two different theories are proposed [16, 6, 7]. 13 Chapter 2 Related work 2.1 Experimental work 2.1 . 1 Uniaxial stretching Measurements have demonstrated that there can be significant changes in the thermal conductivity or diffusivity tensors of elastomers subject to finite deformation. In order to measure the thermal conductivity or diffusivity accurately, Broerman et a1. [2] used Forced Rayleigh Scattering (PBS) to measure the thermal diffusivity of silicone rubber subject to uniaxial stretching. In FRS, a transient optical grating is formed in the material by a diffraction pattern that results from absorption of intersecting laser beams, which have been split from a single source. The rate of decay of the efficiency of diffraction is a function of one of the components of the thermal diffusivity, depending on the relative orientation of the lasers to the specimen. Multiple experiments are required in order to measure multiple components of the diffusivity. 14 Broerman et al. [2] measured the thermal diffusivity properties Of a crosslinked silicone elastomers that were uniaxially stretched up to a stretch ratio of 2.1. The stretch ratio A,- is defined as the current length 1,- divided by a undeformed length L, the direction of stretching is 2' = 1. Motivated by the stress-optic rule [21], Broerman et al. [2] described their results using a stress-thermal rule, which was first suggested by van den Brule [3]. The stress-optic rule for simple elongations is A7113 = 000’ (2.1) where a = tn — t33, with t11 and t33 the components of the Cauchy stress, and An13 = n11 — n33, with ml and 1133 the components of the refractive index of refraction tensor. The stress-thermal rule is written analogously as (011 - a33)/aeq = 00’ (2-2) where all and 0133 are the components of thermal diffusivity tensor in the direc- tion parallel and perpendicular to the direction of stretching, aeq is the equilibrium thermal diffusivity, which is defined as the thermal diffusivity of the unstretched elastomers, and c is a material parameter. For the stretched specimen, the measure- ments of A7213 were made by using a low-power HeNe laser and measurements of the diffusivity 0111 and 0133 were made by using a low-power HeNe laser and PBS. By plotting the normalized thermal diffusivity 011 / aeq and 0133/0139 against A, Broerman et al. reported two linear relationships with different coefficients. It was shown that for initially isotropic silicone rubber, the component of the thermal dif- 15 fusivity tensor in the direction of a uniaxial stretching normalized by aeq, that is 011 /aeq increases by 10% with A1 = 2.1. The component of the diffusivity orthogo- nal to the stretching normalized by aeq, that is 033/0164 decreases by 5% at A1 = 2.1. Figure 2.1 shows these variations in 011 and 033 for stretching in the 1-direction. 1.15 r l. [ ' all/aeq 1.10 ~ ‘ “331“” . H—l t—I—I l—I—l i—I—i “ l—I—I J 105 t-I-t W) 5’ 1 H2“ :3- l. w. 100 I i=5?” 5 . . 1:64—1 l—H H—i l—H H... H—l i—H . H—i ”=1 l—H 0.95 r 0.90 1 1 A 1 1 1 1 1 1 1.0 1.5 2.0 2.5 M Figure 2.1: The relationship between diffusivity and stretch ratio in the direction of stretching of silicon rubber subject to uniaxial loading [2]. The error bars denote i7.5% uncertainty in the stretch ratio Broerman et al. cite Tautz [28] as also finding a linear relationship between con- ductivity or diffusivity and stretch. Tautz [28] measured the thermal conductivity of natural rubber only along the stretch direction for a stretch ratio up to 3.5. A nearly 16 linear relationship existed between the thermal conductivity and stretch ratio, but, the data may not fit the linear relationship well with stretch ratio greater than 2.5 since all the data lie above the trend line. Similar responses are measured in non-elastomeric polymers. Wash and Hansen [36] also characterized the relationship between thermal conductivity of amorphous polymers and the stretch ratio. They measured the thermal conductivity of polystyrene along the direction parallel and orthogonal to the stretch direction with stretch ratios of up to 3.2 and Of polymethylmethacrylate (PMMA) with stretch ratios of up to 6.0. The thermal conductivities were normalized by the respective value of the equilib— rium conductivity. These data showed a linear relationship with respect to stretch ratio. Again, the thermal conductivities were greater than the equilibrium value and those in the perpendicular direction are less than the equilibrium value. They also examined the relationship between the thermal conductivity and molecular weights by measuring the conductivity of PMMA along the stretch direction and perpendicu- lar to the direction with stretch ratios of up to 3.5, B—10834 with stretch ratios of up to 3.9, and B—10836 with stretch ratios of up to 3.0. When these data were plotted against the stretch ratio, the similar linear results were obtained. With an increasing molecular weight, the anisotropy of the thermal conductivity become greater. They explain these results using the theory of Hansen-Ho [17]. Equation 2.2 correlates the difference between these orthogonal components of the diffusivity tensor with the difference in the corresponding Cauchy stress components, where Cauchy stress is defined as a force exerted on a unit area of an object in the current configuration. The magnitude of each component is not uniquely described, 17 however. The change in each component of the diffusivity appearing linear with respect to the stretch ratio in the stretching direction suggests that the stress is also linearly correlated with stretch, which typically occurs over only limited ranges of stretches in most elastomers. 2.1.2 Multiaxial stretching Several studies have examined the results of uniaxial loading, which provides part of the data needed for constitutive modeling of thermal conductivity of elastomers. In order to develop constitutive models suitable for general deformation, measurements for elastomers subject to more complicated loading are needed. LeGall and Wright [23] measured the change in thermal diffusivity of several elastomers that were subject to biaxial stretching. In order to make these multiaxial measurements of thermal diffusivity of elastomers, a computer-controlled Optical-thermomechanical system was designed and developed [26]. LeGall and Wright [22] measured the three orthogonal components of the dif- fusivity tensor during biaxial stretching of polyurethane, natural gum rubber, and neoprene rubber. Rectangular specimans, 4.5cm x 4.5cm x 3mm, were stretched. The polyurethane was initially isotropic and subject to equibiaxial deformation, where both in-plane directions are stretched equally. The diffusivity measurement was made in the central region of the specimen where the deformation was homogeneous. They used an extention of the flash method [8] to measure the orthogonal components of the thermal diffusivity tensor 0: simultaneously. In order to get repeatable thermoelastic results, three specimens were subject to three different mechanical preconditioning 18 protocals and measured with stretch ratios of up to 1.83, 1.98, and 2.0. These results showed an increased diffusivity in the direction of stretch and a decreased value in the orthogonal direction, similar to the uniaxial results. In addition to measuring the initially isotropic elastomers, they also measured the principal components of ther- mal diffusivity of premanufactured natural gum rubber (N GR) and neoprene sheets, which were anisotropic in the undeformed states due to manufacture processes. Their findings also illustrated the Mullin’s effect and the differences due to proportional stretch [22]. LeGall and Wright did not provide a correlation of their data. 2.2 Existing models for k or a The literature reviewed in section 1.3 demonstrate that the thermal conductivity and thermal diffusivity are functions of deformation and temperature level. Experimental and theoretical studies have examined the dependency on temperature [6], [9], [10], [12] [7], and on deformation [28],[17], [32]. 2.2.1 Temperature dependence Dashora [6] examined the temperature variation of the thermal conductivity of di- electric linear elastomers by considering their molecular structure. At temperatures above the glass transition temperature Tg, but below the melting temperature Tm, the temperature dependence of k is controlled by the variation of the mean free path of phonon that depends on the structural features of the system, in which structural 19 scattering is considered to dominate. The contribution of structural scattering to the thermal resistance can be written as W1 = A (2.3) where W1 is the resistance related to k as k = 1/W1, A is a constant for a given elastomer. In the rubbery plateau between T9 and Tm, some vacant sites are occurred with similar effect as the point defects in crystalline solids. Therefore, a rise in temperature leads to an increase of defect density, and consequently, an increased thermal resistance that is linearly proportional to temperature, that is W2 = ClT (2.4) where Cl is a constant for a given polymer. Therefore, these assuming resistances are in series, the relationship between the total resistance W = W1 + W2 and k = 1 / W results in l/k=A+ClT (2.5) The predicted decrease in k with T, above T9, is consistent with the experiment results of Eiermann [9, 10] and Hands [16] and with the analysis of Hands et al. [15]. 2.2.2 Deformation dependence Hansen and Ho [17] proposed that the thermal conductivity of linear high polymers increases in the direction of deformation as the molecular orientation increases in the 20 direction of deformation and decreases in the orthogonal directions. This model is based on a molecular analysis and assumes that a single segment in an amorphous linear polymer has two chemically bonded neighbors on the same molecule, and is surrounded by segments of other molecules. The segment is assumed to interact with its nearest neighboring segments at frequency V1 and with neighboring molecules at a frequency 112, (V1 > V2) and that the energy transferred by each interaction is proportional to the temperature difference between the interacting segments. The energy of each segment is assumed to be a function of the local temperature gradient and position. Based on an energy balance on each segment, and given N segments involved into the energy transfer, Hansen and Ho propose k = (csv1p1/2v2/3N) Z «1. (2.6) where cs is the heat capacity per segment, p1 a proportionality factor, I) the volume occupied by the segment, and q, the macroscopic average energy through the segment, which is assumed to be a linear function of the position. Although the theory of Hansen and Ho [17] fits the experimental data well, there are difficulties in applying this theory because V1, V2, p1, and q,- are not known, nor is it apparent how to measure these parameters. Washo and Hansen [36] developed equations, based on the analysis of Hansen and Ho [17], relating the components of thermal conductivity of a uniaxially stretched elastomer to its stretch ratio. Comparison of the calculated relative thermal con- ductivity was linearly related to the average segment position, which was assumed to behave as an affine deformation, allowing the thermal conductivity parallel to a 21 uniaxial stretching to be calculated to vary as kII/keq — 1 = C (1, — 1) (2.7) where kn is the thermal conductivity along the stretching direction and keg the conductivity in the undeformed state. Then, assuming a uniaxial stretch and an incompressible solid, the thermal conductivity perpendicular to the stretching would vary as kgg/keq — 1 = C (if/2 — 1) = C(Ag — 1) (2.8) where 11:22 is the thermal conductivity perpendicular to the stretching direction and [€33 = 1:22 and A3 = A2, because the material is assumed to be transversely isotropic. Assuming that the principal directions of the diffusivity correspond to those of the stretching, then we may rewrite equations 2.7 and 2.8 as —k——I=C(F—I) (2.9) keg Plotting the experimental results obtained by Broerman et al. in terms Of kg,- versus /\,-, as illustrated in Figure 2.2, in which R2 is the R—squared error when using least square method. We see that a single constant, as suggested in equations 2.7 and 2.8, is insufficient to describe the variation of diffusivity with stretching. 22 - a,da,,,vs A, 1 (122/0.qu k; 1.05 ~ a“ = 0.092411] + 0.9005 5 R’=o.9a41 22" _1.00 ~ 8' A ‘ 022:0.1193241'03791 ”5 R2=0.555 a.” 11x111111m1111111141411111m1111111111144414LJ 0.5 0.7 0.9 1.1 133-1 1.5 1.7 1.9 2.1 2.3 . M Figure 2.2: The normalized components of thermal diffusivity of silicone rubber as functions of their respective stretch ratios. Note the differing slope indicating that a single parameter is insufficient. Washo and Hansen contrast equations 2.7 and 2.8 with the results of Hennig [18], who assumed a network of parallel resistances to write %=(E)+(E) and Hashin and Shtrikrnan (1963) who assumed a series network of resistances to write 23 Washo and Hansen note that these four equations 2.7, 2.8, 2.10, and 2.11 give sim- ilar results for small deformation. Hennig also related deformation to anisotropy by writing 1/k11-1/keq _ 1 2 1 1/k22 —1/keq " 5N ’\1 A1 (2'12) where N is a material parameter, and the right hand side is based on the neo-Hookean description of rubber elasticity. Broerman et al. [31] suggest using a stress-thermal rule to write keg where they define a as the difference in stress between the parallel and perpendicular directions. Broerman et al. leave this in terms of the stress, but for uniaxial stretching, with traction free lateral surfaces, and assuming a neo-Hookean material response, this may be rewritten as k - k 1 __..11 22 = 02 (A? _ Xi) (2.14) where C2 is a new material parameter. If the material exhibits a Mooney-Rivlin response with uniaxial stretching, equation 2.13 may be rewritten as k11- k22 ( 2 1 ) 1 —-—— = C )1 - —- - C — - A 2.15 24 Interestingly, if equation 2.7 is subtracted from equation 2.8, then kn - 1922 _ 1/2 keg _ 0 (A1 — 1/,\1 ) (2.16) The term in parentheses in equation 2.16 is a root of the parenthetical term in equation 2.14. These descriptions have been made in the context of the stress-thermal rule. This has a limitation that only the differences in the components for a material with a transversely isotropic response subject to uniaxial deformation are correlated. By considering the stress-thermal rule for neo-Hookean and Mooney-Rivlin responses, we see differences in these variations between these models. There is a clear need to find a new description, preferably one that describes the variation of the individual components. 25 Chapter 3 Theoretical Framework Several correlations relating the change in diffusivity of solids with respect to finite deformation have been proposed. None of these have been taken advantage of the available results of continuum mechanics and finite elasticity. Here, the utility of choosing the unloaded, isotropic state as the reference configuration is demonstrated. This allows the thermal conductivity and diffusitivity to be formulated using principal stretchs, Cauchy-Green strains, or the invariants of Cauchy-Green strains. Using these measures of deformation should allow the consitiutive description of thermal transport to be more easily generalized to multiple dimensions and complex deformations. 3.1 Choice of reference configuration The thermomechanical response of a material must be characterized with respect to a configuration or coordinate frame, which describes a region in a space that the material occupies at a given instant. Each constitutive model is developed with respect to a specific configuration. Models of the same material response might be 26 different when described in different configurations. One consequence is that the choice of a configuration that describes symmetries for a material in that configuration will introduce constraints on the form of the constitutive law that may be formulated. That is, such symmetries give rise to mathematical restrictions on the constitutive equations. Here, we will describe the thermomechanical response of elastomers in a reference and a current configuration, which are convenient for the constitutive and experimen- tal descriptions, respectively. The reference configuration is here defined by the region of space that the material occupies at reference time to, which is arbitrary, but is here taken as the time at which an elastomer is undeformed and in thermal equilibrium with its surroundings. In contrast, the current configuration is defined as the region Of space that the material occupies at some time t > to, when the elastomer may have undergone significant deformation or changes in temperature. For materials subject to infinitesimal deformations, constitutive models of the thermal response may be developed with negligible error by assuming that the refer- ence and current configurations are indistinguishable. Elastomers, however, are Often subject to large multiaxial deformation, and thus the variables must be described separately in reference and current configurations. Elastomers are often thermally isotropic when undeformed. In order to take advantage of this symmetry when form- ing the constitutive model, we will describe the heat conduction of elastomers in an undeformed reference configuration. In following work, the notation will follow the conventions of finite elasticity (c. f. Gurtin [14]). 27 3.2 Relationship Of q between configurations Conservation of energy dictates the total heat transfer across the surface of a material boundary is unchanged whether the material boundary is described in the current or reference configurations. Thus, / <1 0. Treloar [30] states that the maximum stretch ratio of silicone rubber generally falls below 10 suggesting that streches over 12.7 are unlikely, which ensures that the K11 > 0 since keg > 0. These values would be expected to vary with specific material formulation. Considering that the physical meaning of A1 makes A1 2 1.0 and A2 = 1 / JA— , we have A2 S 1.0. Therefore, f2(A2) has the minimum value when A2 = 1, which is that f2(A2) = —0.0158 + 0.132 + 0.879 = 0.995 (6.10) therefore, Kgg/keq > 0.995 > 0. Since keg > 0, then K22 > 0. This guarantees that K is a positive definite tensor. In deriving a general constitutive relation q as in Chapter 5, three linearly in- dependent vectors, VOT, U’lVoT, and U’zVoT, were carefully chosen based on the analysis of the existing models in Section 2.2.2 of Chapter 2 and the correla- tion from Chapter 4. When another set of three linearly independent vectors (i. e. VoT, UVOT, and U2V0T) was chosen as the basis of heat flux vector function q, one component of thermal conductivity tensor was a parabolic function of U, which yields inaccurate representation of practical situation. Special cases of fluid flow of polymers may have this situation [33, 34]. Wang derived a generalized Fourier law 71 to extend the classical Fourier law to the processes of heat transfer in fluids that involves macroscopic relative motion between two sides exchanging heat, in which q is a function of velocity gradient tensor L. By using material frame indifference and the second law of thermodynamics, an explicit form of K is obtained as K = —(¢OI+¢1D+¢2D2) (6.11) where D is a symmetric part of L. A special form of D such that c(t) 0 0 D= 0 0 0 (6.12) 0 0 0 . .l where c(t) is a scalar function of time t, describes the Situation for a convective velocity changing only in one direction. The thermal conductivity can be found by referring to equation 6.11. The nonzero components of K are the diagonal components, and kn is related to c(t) as kn = -(¢0+¢16(t) +¢26(t)2) (613) Therefore, (1:11 is a parabolic function of c(t). As c(t) varies, kn reaches its extreme value when c(t) = -q>1/(2 :1: (02). In the current study, by referring the theoretical and data analysis in chapter 4, one set of linearly independent vectors, VT, U-lVT, and U'2VT, is chosen carefully 72 to express q. We developed an explicit form of K as a function of U for solids in a systematic, rigorous way by using the principle of material frame indifference, second law of thermodynamics, and properties of material symmetry in a reference configuration. By examining the form of equation 5.37, which is derived for both uniaxial and biaxial situations, the form is seen to fulfill the initial condition of those specimens, and fits the experiment data with small error. Chapter 4 develops a correlation by using phenomenological arguments and least squares fitting of experimental data. How a or K is related to the deformation ten- sor is determined by the curve fitting of the results from uniaxial stretching [2] and biaxial stretching [23]. This method is taken in their respective reference configu- rations (undeformed configuration) to Simplify the constitutive function since those Specimens possess material symmetry in their reference configuration. After exam- ining those data in the reference configuration, figures 4.1 and 4.2 show that each component of a takes the same functional form of its respective stretch ratio, which verifies the material symmetry in the reference configuration and leads to a simple constitutive relation for which (10 is approximately inversely proportional C. These results demonstrate that for an initially isotropic material, if the constitutive relation is developed in the reference frame, one functional form can be sufficient. It is, there- fore, unnecessary to separate the diffusivity in the direction of the situation from the component of the diffusivity in the direction orthogonal to the stretching so that two different linear relations are used for those two directions. In the correlation of (10 with respect to 0'1, two parameters are introduced as in equation 4.8 and demonstrated in Table 4.1 and 4.2, where a is 1 for the uniaxial data 73 but deviates from 1 by as much as 10% for the biaxial data and b is approximately —1 with differences for the different materials, as expected. By considering the special case of equation 4.8 with C = I, meaning that a specimen is undeformed, and a is essentially 1. The deviation may come from the scatter in the measurements. On the other hand, transformation to the reference configuration may cause hide experimental scatter, which needs to be carefully examined. Other possible factors that influence the thermal diffusivity or thermal conductivity are the molecular weight and cross-linking density of the material, which have been examined experimentally in the studies of Washo and Hansen [36] and Rivlin and Saunders [27]. These factors may contribute to the change of b, which warrants further study. Based on this phenomenological study, it becomes possible for the work of chapter 5, in which one general functional form is suggested and developed in a systematic, theoretical framework. An explicit form of K is derived in a mathematical way. The coefficients 1/21, 1,02, and 1,03 are decided by those invariants as described in equation 5.36. Although these coefficients can be decided by curve-fitting the experimental data, the physical meaning and an explicit form of Jk(V()T, U) still need further examination. 74 APPENDICES 75 Constitutive functions for elastic materials In Chapter 1, we cited the thermodynamic analysis of Coleman and Gurtin [5] to note that the thermomechanical response of an elastic material is generally described by four constitutive functions. We also noted that the four equations can be reduced to a smaller set by using the principle of material frame indifference and restrictions due to the second law of thermodynamics. This analysis of these simplifications is now considered routine enough that a specific reference need not be cited, but it is included here for completeness. Recall equations 1.1, 1.2, 1.3, 4 from Chapter 1, but now written in terms of a reference configuration as e = é(F, T, VOT) (1) n = 6(F. T. VOT) (2) T = ‘ (F,T,V0T) (3) q = 609‘. T. VOT) (4) These four constitutive functions must satisfy the principle of material frame indif- ference, meaning that the functional form of the constitutive relations for different Observers should be the same. A description of Observer transformation is demon- strated as Figure 3.1 in Chapter 3. 76 To demonstrate how the principle of material frame indifference may be used to develop functional forms, consider the example of the Cauchy stress function T. The Cauchy stress is introduced by Cauchy’s stress theorem, which states that t = Tn (5) where t is the traction force and n is the unit direction associated with the traction force. For a second Observer, whose current frame is rotated by Q with respect to that of the first observer, we have t“ = Qt and n“ = Qn. Therefore, with t“ = T’n" and t* = Qt we can write T*n* = QTn (6) and due to 11* = Qn T'Qn = QTn (7) Therefore, 1* = QTQT (8) ' If the two observers have the same reference frame, the parameteric variables are related by T" = T, VEST = VoT, and F“ = QF. Material flame indifference requires the constitutive functional forms to be the same, which leads to T* = T(F*, T*, V5T) TlQF.T.voT) (9) 77 Substituting flom equations 9 and 3 into equation 8 leads to 'i‘(QF.T. mm = 0T(F.T. voT)0T (10) Similar substitutions for the other constitutive functions yield the following relations é(QF, T, VoT) = 5(F, T, VoT) (11) fi(QF. T. VOT) = 1“7(1".T.'\70T) (12) Q(QF. T. VoT) = QQ(F.T. VOT) (13) In addition to satisfying the principle of material flame indifference, the constitu- tive functions must satisfy the second law of thermodynamics. The second law may be written as the Clausius-Duhem inequality [19] expressed as F-la - VOT 110(5‘ — T13) — JFTT -FF'1 + JF T S 0 (14) where p0 is the mass density described in the reference configuration and J = det F. Introducing the specific Helmholtz free energy, ii) = e — T7), the Clausius-Duhem inequality may be rewritten as q - F-TVOT p(15+T'fi) —TT.FF‘1 + T S 0 (15) 78 where p = p0 / J is the mass density in the current configuration. Assume that at an initial time to, a fixed point X0 in the reference configuration has following state: T(X0,t0) = a, T(X0,t()) = fl, VT(X0, to) = a, VT(X0,t()) = b, F(X0,t()) = A, F(X0,t0) = B, where a is a positive number and A is a positive tensor. Otherwise, a, [3, a, b, A, B are arbitrary. Fields of x and T(X, t) that could satisfy the initial state are x(X,t) = X0+F(t)(X—X0) (16) with F(t0) = A and F(t0) = B and T(X, t) = aexp [EU — to) + (a + (t — t0)(b — a?” - (X — X0)/a] (17) The initial state of the point must also satisfy the second law of thermodynamics. Substituting flom the initial conditions above into equation 15 results in fi(x01t0)l1/3T(ai a) A)fl + 113V0T(aa 3) A) ' b + 523F011 31 A) ° B +9002. a, A)] — TT(o, a, A) . BA-1 + $6191. a, A) . A-Ta g 0 (18) Equation 18 must hold for all a, B, a, b, A, B. First fix a, B, a, A, B. Then dividing through by p(X0, to), equation 18 may be rewritten as ¢V0T(0181A) + f0(a.fl.a.A.B.P0(X0)) S 0 (19) 79 which must hold for all b. Equation 19 can be violated unless ¢V0T(ar 8) A) ' b = O (20) which shows that TVOTWJ. A) = 0 (21) which is true for all a,a, and A. It follows that the constitutive response function A 1M0, a, A) is independent of the temperature gradient VOT and we may write 8 = T(F. T) (22) Returning to the inequality in relation 18 and making used of equation 20, we may write 8x6. to) [1131(0. Am + (318. A) - B + 968. a. A)] -TT(o,a, A) . 13A-1 + $618. a, A) . A-T a g 0 (23) which must hold for all a, B, a, A, and B. By assigning fixed values to a, a, A, and B, equation 23 requires (3(X0. to) [$18.19) + 6(a.a. A)] 9 + go(a.a. A. B) s o (24) 80 for all values of 6. Equation 24 can be violated unless iffla. A) + (761.8124) = 0 (25) This says that that constitutive response function 1'7(T, VOT, F) is independent Of VOT. Moreover, A fi(T1V0T) F) = _wT(Tv F) = fi(T) F) (26) Again returning to the equation 23, and employing this new result, we have p(X(), t0)rLF(a, A) - B — TT(a,a, A) - BA"1 + imam, A) - A—Ta _<_ 0 (27) which must hold for all a, a, A, and B. Now assigning fixed values to a, a, and A requires, in order not to violate the inequality by the choice of B, that [p(xo. to)(i1.~(a. A) — TT(a.a. A)A'T] - B = 0 (28) and that $901. a, A) - A-Ta s 0 (29) Equation 28 requires the constitutive response function T(T, VOT, F) to be indepen- dent of VOT and given by T(T, VOT, F) = pFr/3F(T, F) = T(T, F) (30) 81 A more general result flom inequality 29 may be developed by assuming that 61(0, a, A) is a continuous function of VOT and writing where 0(a) means that lim|al_,0 O(a)/|a| is bounded. Then, from inequality 29 we have [A‘19(a. 0. A) + 0(a)] - a s o (32) which must hold for all a, a, and A. If we divide through by la] and take lim |a| —> 0, we then get 41-16(81). A) - e = 0 (33) for all unit vectors e. Thus, A’lffia, 0, A) = 0, which implies that (“1(0, 0, A) = 0. This shows that the constitutive response function q(T, VoT, F) must have the property that it vanishes when the temperature gradient vanishes. Moreover, the final residual inequality 6(T, VoT, F) . F‘TVOT g 0 (34) must hold for all T, VOT, and F. This represents the classical Fourier inequality. 82 Bibliography [1] Arruda, E. M. and Boyce, M. C., 1993, “A ThreeDimensional Constitutive Model for the Large Stretch Behavior of Rubber Elastic Materials,” J. M ech. Phys. Solids, 41(2), pp. 389-412. [2] Broerman, A. W., Venerus, D. C. and Schieber, J. D., 1999, “Evidence for the stress-thermal rule in an elastomer subjected to simple elongation,” Journal of Chemical Physics, 111(15), pp. 6965-6969. [3] Brule, B. H. A. A. van den, 1989, “A network theory for the thermal conductivity of an amorphous polymeric material,” Rheologica Acta, 28 pp. 257-266. [4] Choy, C. L., Chen, F. C. and Luk, W. H., 1980, “Thermal Conductivity of Oriented Crystalline Polymers,” Journal of Polymer Science: Polymer Physics Edition, 18, pp. 1187-1207. [5] Coleman, B. D. and Gurtin, M., 1967, “Thermodynamics with Internal State Variables,” Journal of Chemical Physics, 47(2), pp. 597-613. [6] Dashora, P., 1993, “A Study of Variation of Thermal Conductivity Of Elastomers with Temperature,” Physica Scripta, 49, pp. 611-614. [7] Dashora, P. and Gupta, G., 1996, “On the temperature dependence of the ther- mal conductivity of linear amorphous polymers,” Polymer, 37(2), pp. 231-234. [8] Doss, D. and Wright, N. T., 2000, “Simultaneous Measurement of the Orthogonal Components of Thermal Diffusivity in PVC Sheet,” Journal of Heat Transfer, 122(1), pp. 27-32 [9] Eiermann, K., 1966, “Model Interpretation of Thermal Conductivity in High Polymers,” Rubber Chemistry and Technology, 39, pp. 841-857. [10] Eiermann, K. and Hellewege, K. H., 1962, “Thermal Conductivity of High Poly- mers flom —180°C to 90°C,” Journal of Polymer Science, 57, pp. 99-106. [11] Freeman, J. J. and Greig, D., 1984, “The Low Temperature Thermal Conduc- tivity of Rubber,” Advances in Cryogenic Engineering, 30, pp. 105—110. [12] Freeman, J. J. and Greig, D., 1986, “The Low Temperature Thermal Conduc- tivity of Rubber,” Nonmetallic Materials and Composites at Low Temperature, 3, pp. 9-13. 83 [13] Ffied, E., 2002, “An elementary molecular-statistical basis for the Mooney and Rivlin-Saunders theories of rubber elasticity,” Journal of the Mechanics and Physics of Solids, 50(3), pp. 571-582. [14] Gurtin, M. E., 1982, An Introduction to Continuum Mechanics, Academic Press, New York. [15] Hands, D., Lane, K., and Sheldon, R. P., 1973, “Thermal Conductivities of Amor- phous Polymers,” Journal of Polymer Science, Part C-Polymer Symposium, 42, pp. 717-726. [16] Hands, D. and Horsfall, F., 1977, “The thermal diffusivity and conductivity of natural rubber compounds,” Rubber Chemistry and Technology, 50(2), pp. 253- 265. [17] Hansen, D. and Ho, C. C., 1962, “Thermal Conductivity of High Polymers,” Journal of Polymer Science: Part A, 3, pp. 659-670. [18] Hennig, H., 1967, “Anisotropy and Structure in Uniaxially Stretched Amorphous High Polymers, ” Journal of Polymer Science, Part C, 16, pp. 2751-2760. [19] Holzapfel, G. A., 2000, Nonlinear Solid Mechanics, J. Wiley & Sons, New York. [20] Humphrey, J. D. and Rajagopal,K. R. , 1998, “Finite Thermoelasticity of Con- strained Elastomers Subject to Biaxial Loading,” Journal of Elasticity 49(3), pp. 189-200. [21] Janeschitz-Kriegl, H., 1983, Polymer Melt Rheology and Flow Birefringence, Springer-Verlag, Berlin. [22] LeGall, E. L., 2001, Thermal Difl‘usiuity of Elastomers Subject to Finite Biaxial Deformation, Ph.D. Dissertation, University of Maryland, Baltimore County. [23] LeGall,E and Wright, N .T., 2000, “Changes in Thermal Diffusivity of Rubber Due to Mechanical Preconditioning,” Proceedings of the Fourteenth Symposium on Thermophysical Properties, Boulder, Colorado, June, 2000, pp. 1-22. [24] Ogden, R. W., 1984, Non-linear Elastic Deformations, Dover Publications, Mi- neola, New York. [25] Ogden, R. W., 1992, “On the thermoelastic modeling of rubber-like solids,” Journal of Thermal Stresses, 15, pp. 533-557. [26] Ortt, E. M., Doss, D. J., LeGall, E., Wright, N. T. and Humphrey, J. D., 2000, “A device for evaluating the multiaxial finite strain thermomechanical behavior of elastomers and soft tissues,” Journal of Applied Mechanics,67, pp. 465-471. [27] Rivlin, R. S., and Saunders, D. W., 1951, “Large elastic deformations of isotropic materials. VII. Experiments on the deformation of rubber,” Philosophical Trans- actions of the Royal Society of London, A243(2), pp. 251-288. 84 [28] Tautz, H. E., 1959, “Bextimmung der Warmeleitfahigkeit von Kautschukvulka— nisaten in Abhangigkeit von der Dehnung,” Experimentelle Technik Der Physik, 7, pp. 1-13. [29] Tobolsky, A.V.,1960, Properties and Structures of Polymers, J. Wiley 85 Sons, New York. [30] Tmloar, L. R. G., 1975, The Physics of Rubber Elasticity, 3rd Edition, Oxford University Press. [31] Venerus, D., Schieber, J ., Iddir, H., Guzman, J. and Broerman, A. W., 1999,“Re- laxation of Anisotropic Thermal Diffusivity in a Polymer Melt Following Step Shear Strain,” Physical Review Letters, 82(2), pp. 366-369. [32] Venerus, D. C., Schieber J. D., Iddir, H., Guzman, J. and Broerman, A., 2001, “Anisotropic Thermal Diffusivity Measurements in Deforming Polymers and the Stress-Thermal Rule,” International Journal of Thermophysics, 22(4), pp. 1215- 1225. [33] Wang, L., 1994, “Generalized Fourier Law,” International Journal of Heat and Mass Transfer, 37(17), pp. 2627-2634. [34] Wang, L., 2001, “Vector-field Theory of Heat Flux in Convective Heat Transfer,” Nonlinear Analysis A, 47, pp. 5009-5020. [35] Wang, Y. and Wright, N. T., 2005, “A Relationship Between Thermal Diffusivity and Finite Deformation in Polymers” , International Journal of Thermophysics, 26(6), pp. 1849-1859. [36] Washo, B. D. and Hansen, D., 1969, “Heat Conduction in Linear Amorphous High Polymers: Orientation Anisotropy,” Journal of Applied Physics, 40(6), pp. 2423-2427. [37] Wineman, A., 1999, “Changes Due to Scission in an Elastomeric Cylinder Un- dergoing Circular Shear and Heat Conduction,” Finite Thermoelasticity, 236, ASME, pp. 15-19. [38] Wineman, A. and Min, J ., 2003, “Time-dependent scission and cross-linking in an elastomeric cylinder undergoing circular shear and heat conduction,” Inter- national Journal of Non-Linear Mechanics, 38, pp. 969-983. 85 l[illl]l[l][lllll]]l[l]l[lll]l][[[ll