.z\‘:-12Rit1:i:':.:2 . amine-#5:?“ <‘ ' " ' RATIONAL AND PRACTICAL NONEQUILIBRIUM THERMODYNAMICS Thesis for the Degree of Ph. D. MICHIGAN STATE UNIVERSlTY SARA ERB INGLE 1971 """"""""" Erik-5"“ LIBRARY Michiga .1 Sub Univeriy This is to certify that the thesis entitled RATIONAL AND PRACTICAL NONEQUILIBRIUM THERMODYNAMICS presented by Sara Erb Ingle has been accepted towards fulfillment of the requirements for Ph. D. degree in Chemistry Major professor Date_Decmh2r_3.,_19.1L 0-7639 ? amoI’NG av " HOME 8. SflNS' BDUK BlN'JERY 1N6. m LIBRARY BINDERS SKIIIHILIICIIW ABSTRACT RATIONAL AND PRACTICAL NONEQUILIBRIUM THERMODYNAMICS “By Sara ErbIngle A fundamental theory of nonequilibrium thermody- namics based upon rational mechanics is presented, is compared with traditional nonequilibrium thermodynamics, and is applied to two practical problems: heat conduc- tion of pure fluids and the Dufour effect in fluid mix- tures. The rational mechanical theory of nonequilibrium thermodynamics is developed logically and systematically from the conservation laws of mechanics and the Second Law of thermodynamics. Full account is taken of the kinetic energy of diffusion, and a variety of sets of independent variables is investigated. For the case of linear reSponse functions, a full, illuminating comparison is made between the rational approach and the traditional approach. It is found that the two approaches give simi- lar, but not identical, results. Equations directly applicable to experimental determination of thermal conductivity are obtained from Sara Erb Ingle macroscopic transport equations by means of a well—defined, self-consistent perturbation scheme. The barycentric ve- locity and variability of all properties are included ex- plicitly. By introduction of a second perturbation scheme, the transport equations for the Dufour effect, which is the deve10pment of a temperature gradient due to diffusion, have been solved for geometrically well-defined cells which have either all adiabatic walls or adiabatic lateral walls and diathermic ends. The heat of mixing and varia- tion of other physical properties are fully included. For typical nonelectrolytes, the temperature difference produced by the Dufour effect could be as large as 0.2 degrees for a diffusing mixture with initial mass frac- tion difference of 0.8. These results can be used (1) to design experiments to test the Onsager heat-matter reciprocal relation and (2) to avoid undesired tempera- ture variations in diffusion experiments. RATIONAL AND PRACTICAL NONEQUI LI BRI UM THE RMODYNAMI CS BY SARA ERB INGLE A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1971 To Jim ii ACKNOWLEDGMENTS I wish to thank the National Science Foundation for the Predoctoral Fellowships which I held throughout my graduate career. I would also like to thank Michigan State University for a First-Year Fellowship. I am deeply grateful to my research director, Dr. Frederick Horne, for his consultation and direction, for his provoking research ideas, and for keeping my weight down by his being so hard to find. I appreciate the helpful suggestions and criti- cisms of Dr. James Dye, who served as my second reader. My thanks are due to Mr. James Olson and Mr. Ping Lee for their helpful discussions of research. I also thank Mr. William Waller and Mr. Richard Vandlen for their help with my computer problems. Finally, I would thank the entire department for putting up with "Grumpy Old Sally," and particularly my husband Jim who has borne the brunt of my frustrations. iii TABLE OF CONTENTS Page LISTOFTABLES......"............ vi LIST OF FIGURES O O I I O O O O O O O I O O O O O O Vii LIST OF SYMBOLS FOR CHAPTER II . . . . . . . . . . X Chapter I 0 INTRODUCTION 0 O O O O O O I O O O O O O O I 1 II 0 RATIONAL MECHANICS O O O O O O O O O O O O O 1. Introduction . . . . . . . . . . . . . . 2. Thermodynamic Process and the Equations of Balance . . . . . . . . 12 3. Constitutive Assumptions . . . . . . . . 18 4. Ordinary Binary Fluid Mixtures . 23 5. Transport Equations in the Independent Variables T, p, W1 . . . . . . . . . . . 32 6. A New Rational Mechanical Derivation . . 38 7. Comparison of Several Theories for Binary Mixtures of Fluids. . . . . . 57 8. Expressions for EntrOpy Production and the Entropy Balance Equation . . . . . . 62 9. Conclusions and Discussion . . . . . . . 71 III. THERMAL CONDUCTIVITY. . . . . . . . . . . . 77 1. Introduction . . . . . . . . . . . . . . 77 2. Transport Equations. . . . . . . . . . . 81 3. Approximation Methods. . . . . . . . . . 83 4. Practical Results. . . . . . . . . . . . 91 IV. THE DUFOUR EFFECT . . . . . . . . . . . . . 102 1. Introduction . . . . . . . . . . . . . . 102 2. Transport Equations. . . . . . . . . . . 106 3. Perturbation Scheme. . . . . . . . . . . 109 4. Zeroth-Order Solutions for an Adiabatic System. . . . . . . . . 112 5. Higher Order Solutions for an Adiabatic System. . . . . . . . . 116 6. Solutions for a Cell with Diathermic Ends. . . . . . . . . . . . . 122 7. Calculated Temperature Variations. . . . 123 8. Discussion . . . . . . . . . . . . . . . 145 iv Page BIBLIOGRAPHY O O O O O O O O O O O O O O O O O O O 158 APPENDIX A. Some Equations of Thermostatics . . . 161 APPENDIX B. A Rational Mechanical Development Of the TIP Theory I O C O O O O O O O 162 APPENDIX C. The Temperature Solution T1 for Thermal Conductivity in a Pure Fluid I O O O O C O O O O O O O 166 APPENDIX D. The Temperature Solutions T55 for the Dufour Effect in Adiabatic Cells and Cells with Diathermic Ends. . . . 168 APPENDIX E. The Temperature Equations and Solution for T555 for the Dufour Effect in Adiabatic Cells . . . . . . 174 APPENDIX F. The Relationship of Heat of MiXing to 3(Hl - H2)/3W1. o o o o o o 178 LIST OF TABLES Table Page 1. Approximate physical properties of CCl4. . . 95 2. Typical values of physical properties for a binary mixture of organic liquids. . . . . 124 vi LIST OF FIGURES Figure Page 1. Plot of To versus (z/a) for CCl4 at 30 s, 60 s, 90 s, 120 s, and 150 s . . . . . . . . 97 2. Plot of T1 versus (z/a) for CCl4 at 30 s, 60 s, 90 s, 120 s, and 150 s . . . . . . . . 100 3. Temperature variations in an adiabatic cell for Afifi = 0.0, pDQi = 3 x 10‘2J -13-1, K = 2 x lO-lJ m-ls—lK-l, and mi = 0.8 at (t/e) = 0.01, 0.05, 0.09, 0.13, and 0.17 . . . . . . . . . . . . . . . 129 m 4. Temperature variations in an adiabatic cell for 085 = 0.0, pDQi = 3 x 10‘20 ‘1s-1, K = l x lO-lJ m-ls-lK-l, and mi = 0.8 at (t/e) = 0.01, 0.05, 0.09, 0.13, and 0.17 . . . . . . . . . . . . . . . 132 m 5. Temperature variations in an adiabatic cell for Afih = - 6.5 x 1030 kg‘l, 9991 = 3 x lO-ZJ m—ls-l, K = 2 x lO-lJ m-ls K'l, and mi = 0.8 at (t/e) = 0.01, 0.05, 0.09, 0.13, and 0.17 . . . . . . . . . . . . 134 -1 6. Temperature variations in an adiabatic cell for (t/e) = 0.1, 600*l = 3 x lo'zJ m-ls-l, K = 2 x lO-lJ m-ls_1K-l, and w: = 0.8 with Aim = - 19.5 x 103, - 6.5 x 103, 0.0, 6.5 x 1033 kg"1 . . . . . . . . . . . . 137 vii Figure 7. 10. 11. 12. Temperature variations in an adiabatic cell for (t/0) = 0.1, Afih = - 6.5 x 1030 kg-l, K = 2 x lo-lJ m-ls_lK-l, and mi = 0.8 with “”01 = 6 x 10'2, 3 x 10‘2, 3 x 10‘3, and -3 XlO-ZJ mnls"l . . . . . . . . . . . Temperature variations in an adiabatic cell for (t/G) = 0.1, AHm = 0.0, K = 2 x 10-1J m 1s T1 1, andm pDQi = 3 x lO-ZJ m‘ls"1 with w; = 0.8, 0.5, 0.2, 0.05, and 0.01 . . . . . . . . . . . . . . . . . . . Temperature variations in a cell with diathermic ends for A3 = 0.0, pDQ* = -2 -1 -1 m -1 .. -1 -l 3 x 10 J m s , K = 2 x 10 J m s K , and mi = 0.8 at (t/e) = 0.05, 0.09, 0.13, and 0.17 C O O O O O O O O C O I O O O O 0 Temperature variations in a cell with diathermic ends for AHm = 0.0, pDQ* = 3 x 10 2J m 1s 1, K = l x lo—lJ m‘ s-lK-l, and mi = 0.8 at (t/e) 0.05, 0.09, 0.13, and 0.17 . . . . . . . . . . . . . . . . . Temperature variations in a cell with diathermic ends for Afifi = - 6. 5 x 103J kg-l, 000i = 3 x 10 2J m 1s 1, K = 2 x 10 1J m , and wl = 0.8 at (t/O) = 0.05, 0.09, 0.13, and 0.17 . . . . -ls- -1K -1 Temperature variations in a cell with diathermic ends for (t/G) = 0.1, Afih = - 6.5 x 1030 kg‘l, K = 2 x 10‘10 m s‘lx'l, and w? = 0.8 with AER = - 19.5 x 103, - 6.5 x 103, 0.0, and 6.5 x 103 . . -1 viii Page 139 142 144 147 149 151 Figure Page 13. Temperature variations in a cell with diathermic ends for (t/G) = 0.1, Afih = - 6.5 X 103J kg-l, K = 2 x lO-IJ -1 -l -l o . m s K , and ml = 0.8 With pDQi = 6 x 10'2, 3 x 10'2, 3 x 10'3, and — 3 x lO-ZJ m-ls-l . O O C O C C C C C C O O 153 14. Temperature variations in a cell with diathermic ends for (t/O) = 0.1, Afih = — 6.5 x 1030 kg‘l, K = 2 x 10‘10 m-ls-lK-l, and pDQi = 3 x 10-2J m-ls-l, with 03 = 0.8, 0.5, 0.2, 0.05, and 0.01. . . 155 ix LIST OF SYMBOLS FOR CHAPTER II Roman miniscules b: -- external specific forces acting on component a (2.5)* cu -— number density of component a (2.19) d:. -- symmetric part of the gradient of the component 3 velocity (3.3) fi -- total entropy flux (3.2) j: -- diffusion flux of component a relative to the barycentric velocity (5.2) ki -— entrapy flux not due to heat flow (3.2) mi -- specific interaction force exerted on component 1 by component 2 (6.7) m: -- Specific interaction forces acting on component a due to interaction with other components (2.5) p -- a constitutive coefficient (6.47) qk -- a heat flux (Bartelt's) (2.8) qfi —- heat flux for total energy balance equation (2.10) r -- Specific radiative heat supply (2.8) ui -— diffusion velocity of component 1 relative to component 2 (6.2) u? -- diffusion velocity of component a relative to the barycentric velocity (2.1) vk -- barycentric velocity (2.2) v: -- velocity of component a (2.2) *These numbers refer to the equations in Chapter II where the symbols first appear. X w -- mass fraction of component a (5.2) Xk -- space coordinate (2.1) Roman majiscules AI -- specific internal Helmholtz free energy (3.11) Cb -- specific heat capacity at constant pressure (5.21) E§ -- specific heat capacity at constant volume (5.20) ny -- constitutive coefficient (4.4), (4.6) ny -- constitutive coefficient (6.44), (6.45) Dy -- combinations of constitutive coefficients (5.12) E -- total energy (2.9) EB -- internal energy (total energy less bulk kinetic energy) (6.8) EI -- specific internal energy (total energy less kinetic energy) (2.8) E -- partial specific internal energy of component 0 °‘ (5.21) * Fél) -- interaction forces of Bearman and Kirkwood °‘ (2.18) fig -- partial specific enthalpy of component a (5.22) K -- a constitutive coefficient (4.9) Ma -- molecular weight of component (2.17) P -- pressure (4.22) P' —- (= P) a constitutive coefficient (6.46) Qi -- a generalized heat flux (7.2) S —- specific internal entropy (3.9) Sa -- partial specific density of component (5.14) T -- temperature Tij -— total stress tensor (2.10) xi Tij -- sum of the partial stress tensors (6.7) V -- specific volume (4.38) V -- specific scalar potentials acting on component 0 °‘ (2.14) Va -- partial specific volume of component a (5.5) XA —- a variable which must be zero at equilibrium (4.17) X? -- external forces of Bearman and Kirkwood (2.17) Zi -- any force in the entropy production which in- cludes the gradient of the chemical potential difference (7.1) Greek miniscules ax —- viscosity coefficients (6.47) 8 -- thermal expansivity (5.5) 8' -- isothermal compressibility (5.5) 0 -- Kronecker delta (3.15) ij n -- shear viscosity coefficient (6.46) ”GB -- viscosity coefficients (4.8) u -- viscosity coefficient (4.7) u' -- a viscosity coefficient (6.48) “a -- specific chemical potential of component a (4.38) u' -- chemical potential of component a plus the exter- nal potentials (5.12) ”11 -- derivative of the chemical potential with re- spect to wl (5.17) v -- number of components Egj -- symmetric part of the partial stress tensor (3.1) Na -- constitutive coefficients (4.8) xii p -- density of component a (2.1) a p -- density a Gji -- partial stress tensor of component a (2.5) BKoqi -- partial stress tensors of Bearman and Kirkwood 3 (2.16) Tij -- a difference of partial stress tensors (6.10) ng -- antisymmetric part of the partial stress tensor of component (2.6) 0 -- the coefficient of bulk viscosity (6.46) ¢aB -- viscosity coefficients (4.8) Xx -- viscosity coefficients (6.46) mi. -- linear combination of the antisymmetric parts 3 of the gradients of component velocity (3.3) Greek majiscules Ei -- the force of rational mechanics and of Bearman and Kirkwood which corresponds to the diffusion flux (7.4) nij -— viscous pressure tensor (5.5) 0 -- rate of specific entropy production (3.9) Qxy -- phenomenological coefficients (7.1) xiii CHAPTER I INTRODUCTION Few chemists have an academic knowledge of the field of nonequilibrium thermodynamics, yet we encounter it in almost every chemical phenomenon we study. The treatment of processes as equilibrium phenomena is often reasonable, but the validity of such assumptions can be proven only by exploring the phenomena from nonequilibrium vieWpoints. Moreover, the vast majority of physical phe- nomena are not at equilibrium. In order ultimately to understand such important phenomena as metabolism, at one extreme, and spectrosc0py, at the other, we must treat them as dynamic nonequilibrium phenomena. To these ends we use rational mechanics to examine the theory of nonequilibrium thermodynamics, and we com- pare this theory to the traditional macroscopic nonequi- librium theory which has been used in practical applica- tions for many years. In the development either of an equilibrium or of a nonequilibrium theory, assumptions are made in order that the theory be viable; however, it is important that one knows the limitations of his theory as determined by these assumptions. Moreover, we make actual application of the full theory to two problems of considerable practical importance: namely, we deve10p complete phenomenological theories for thermal conduction experiments and for the Dufour effect. The rational mechanics approach to nonequilibrium thermodynamics consists in the logical mathematical deriva- tion of the consequences of the laws of physics and the Second Law of Thermodynamics when these laws are applied to some well-defined physical system. We restrict our work to fluids and, in general, define a rather simplistic (to the mathematical mind) system. Yet what we are con- cerned with as chemists is what we can measure; regrettably, the translation of a mathematical truth into an expression for an experimental observable is no easy process. If we had a handy entrepy meter, nonequilibrium thermodynamics might not be so formidable, but even this quantity, funda- mental to the definition of equilibrium and nonequilibrium, must be translated into observable quantities. We cannot deny that our mathematical statements are true; what we seek is to know what they can tell us about the world we observe. Nonequilibrium thermodynamics in the form of a theory to fit empirical observations has been used since the Nineteenth Century, but its recent growth is due to the pioneer work of Onsager in 1931. A coherent, practical theory has been developed by Prigogine (1947, 1955), Kirkwood and Crawford (1952), de Groot (1945, 1955), Meix- ner and Reik (1959), Fitts (1962), and de Groot and Mazur (1962). This theory has been the foundation of modern applications in nonequilibrium thermodynamics. More re- cent still is the theory of thermodynamics and rational mechanics which has been develOped by mathematicians of the first rank. Much work is still being done in the field of continuum mechanics, but among its greatest con- tributions we can name the works of Truesdell (1957), Truesdell and Toupin (1960), and Coleman and Noll (1963). Our second chapter deals with rational mechanical theories of binary fluid mixtures. Though there exists a great body of works on such fluids, the significance of the theory to the exPerimentalist has not been explored and an explicit comparison of the rational mechanical theories to the practical theory has not been made. Such extensions and comparisons are the subject of Chapter II. The third and fourth chapters rest heavily on the practical theory. Both chapters are eXperimentally ori- ented in that they concern the solution of nonequilibrium prdblems under particular initial and boundary conditions and, therefore, experiments could be designed using the results here. In Chapter III we present the theory of thermal conductivity for a pure fluid in a flat plate or thermal diffusion cell to which a temperature gradient is applied. A self-consistent, well-ordered perturbation scheme enables us to account for variations of thermodynamic quantities in Space and time and to predict accurately tem- perature variations in the cell. Hopefully, these results will allow experimentalists to improve their accuracy. Chapter IV is a theoretical investigation of the Dufour effect in liquids. This phenomenon is the produc- tion of a temperature gradient in response to a concentra— tion gradient. The effect is quite small in liquids and is complicated by heat of mixing, and, thus, it has been ignored for many years. Our theory provides a quantita- tive prediction of the temperature variations caused by the Dufour effect and serves as a guide to the experi— mentalist in measuring the effect. Philosophically, the reason for measuring it is that such an experiment would test the previously untested heat-matter Onsager reciprocal relation. The most important practical usefulness of a full theory of the Dufour effect is that it gives guide- lines to the diffusion experimentalist who wishes to avoid temperature effects. CHAPTER II RATIONAL NONEQUILIBRIUM THERMODYNAMICS 1. Introduction In recent years many investigations of the thermo- dynamics of binary mixtures of nonequilibrium fluids have been made based on the works of Coleman and Noll (1963), Coleman and Mizell (1963, 1964), Truesdell (1957), and Truesdell and Toupin (1960). Typical of these are the works of Muller (1967, 1968), Bartelt (1968), Doria (1969), Bartelt and Horne (1970), and Truesdell (1969). The logi- cal superiority of these methods, often referred to as rational mechanics, over the traditional heuristic ap- proach called "thermodynamics of irreversible processes" (TIP) cannot be questioned, but it is also true that the TIP theory as exemplified by Fitts (1962), de Groot and Mazur (1962), Gyarmati (1970), and Haase (1969) has been most useful and has enjoyed wide, successful application. A full comparison of TIP results with those derived from rational mechanics is clearly desirable, although most of the eXpounders of rational mechanics disdain to make such a comparison. We show that the two approaches lead to very similar results and that the practitioners of TIP, with their practicality and intuition, have arrived at almost the same results as the rational mechanists, with their mathematical rigor and integrity. In the course of making the comparison, we also illuminate the meaning and significance of various parts of both theories. The essential difference between the rational mechanics approach and the TIP approach is implied in the word "rational." The rational mechanics approach is based on known balance (or conservation) equations and the re- sults of these laws are logically and systematically de- rived. There is no point at which the rational mechanist can interject his intuition as to how the physical system being described ought to behave; this is, in fact, why rational mechanists abhor the practical theory while, at the same time, the users of TIP have difficulty in grasp- ing the physical significance of rational mechanics. The assumption essential to rational mechanics is the Second Law of Thermodynamics. Further assumptions, called constitutive assumptions, are made after a set of variables has been chosen for the physical system. These constitutive assumptions are equations which mathematically describe how dependent quantities respond to the selected independent variables. Initially, all constitutive assump- tions obey the principle of equipresence; i;g;, every con- stitutive equation depends on the same independent variables unless this dependence contradicts the symmetry of the material, the entropy inequality, or fundamental mathe- matics. Through such contradictions the response functions are limited to dependence only on certain independent variables or combinations thereof. The theory of TIP likewise depends on balance equations and on the Second Law, but the phenomenological equations, which correspond to the constitutive equations, are not rigorously developed. The entropy production ob— tained is said to be a sum of products of "fluxes" and "forces." Just what is a flux and what is a force is de- cided by intuition. Moreover, the fluxes are always written only as linear functions of the forces in the phenomenological equations. All thermodynamic functions are taken to be the same functions of a set of state variables as they are in thermostatics. The work of Muller (1967, 1968) was the first suc- cessful formulation of a theory for binary fluids through a rational mechanical development. It was from this theory that both the theory of Doria (1969) and that of Bartelt (1968) and Bartelt and Horne (1970) were developed. The principles layed down by Coleman and Noll and extended to mixtures by Mfiller are used by both. The work of Bartelt and Horne is particularly valuable because through care- ful selection of variables it attained a bilinear form for the entropy production and because it gives more practical equations (i.e., transport equations) than does the more general theory of Doria. Bartelt and Horne pre- sented a linear theory, in that the constitutive equations are assumed to be linear functions of those independent variables which are not scalar state variables, whereas the theory of Doria is not restricted to the linear case. Bartelt did consider general non-linear constitutive equations, but the consequences of the Clausius-Duhem inequality were not fully taken into account in these general equations. The choice of independent variables greatly af- fects the appearance of any rational mechanical theory; the choices even lead to some contradictory results in the limit of the linear theory. No matter how rigorously the linear approximation is made with respect to one set of variables, it is still an approximation, and an equally rigorous linearization with respect to another set of variables is not the same approximation. There is an Obvious similarity between, on the one hand, the multitude of sets of conjugate fluxes and forces which TIP boasts and, on the other hand, the multitude of sets of indepen— dent variables with respect to which the rational mecha- nist can linearize his equations. The variables in each theory are selected by trial and error in order to attain the result in its best possible form for some con- text. The choices of TIP have traditionally been experi- mentally motivated; those of rational mechanics are mathematically motivated. The mere choice of variables even before linearization can cause varying results from the Clausius-Duhem inequality (Truesdell, 1969) if an insufficient number of variables is used. The variables of Bartelt and Horne give a concise bilinear form for the entropy production, but, unfortunately, they do not lead to linear constitutive equations which are fully comparable to those found in TIP. The chief diffi- culty appears in attempting to relate the nine viscosity coefficients of the rational mechanics approach to the two more familiar coefficients of bulk and shear viscosity. Doria's variables allow him to give rigorous mathe- matical interpretations of the Clausius-Duhem inequality together with proofs that some constitutive relations are independent of certain independent variables. When Doria's constitutive equations are linearized, however, one does not obtain a bilinear form for the entropy production. Many restrictions can, nevertheless, be obtained from the general entropy inequality. We present here a new rational mechanical develop- ment of the thermodynamics of nonequilibrium binary fluid mixtures. It is based on a set of variables which is a mixture of those used by Bartelt and by Doria. Unlike the work of Bartelt, the independence of the internal Helmholtz free energy of the component density gradients is proven in general rather than only in the linear case. 10 Our development is superior to others in that more of our phenomenological coefficients can be given physical meaning, and the nature of their source can be better understood. Two of the nine viscosity coefficients are the recognized shear and bulk viscosity coefficients. The entropy pro- duction, though not bilinear in form, is readily comparable to the entrOpy production of TIP. In addition, this work serves as clarification and extension of previous work by other authors. The implica- tions of the transport theory of Bartelt and Horne are examined, the great similarity of Bearman and Kirkwood's (1958) microscoPic theory to the theories of rational mechanics is demonstrated, the assumptions which must necessarily be made in a rational mechanical development in order to arrive at the results of TIP are illustrated, the limitations of the postulate of local state1 are ex- plored, and the similarities as well as the irreconcili- able disparities among several approaches to nonequilib- rium fluid mixture theory are discussed. 1"For a system in which irreversible processes are taking place, all thermodynamic functions of state exist for each element of the system. These thermodynamic quan- tities for the nonequilibrium system are the same functions of the local state variables as the corresponding equilib- rium thermodynamic quantities."--Fitts (1962). This is usually called the postulate of local equilibrium, which is a misnomer. Kestin (1966) introduced the more apt ter- minology, "postulate of local state." 11 Particular attention is given to the kinetic energy of diffusion2 and its consequences. Rational mechanics has generally included the kinetic energy of diffusion, but has also failed to give transport equations which are as prac- tical as those of TIP. The enigma of the contribution of kinetic energy of diffusion has been considered only by Gyarmati whose work is extremely lucid, but unviable. To avoid the problems created by inclusion or exclusion of the kinetic energy of diffusion he defines two total energies, one from which one subtracts the bulk kinetic energy, the other from which he subtracts the bulk and diffusion ki- netic energies, thus arriving at the same internal energy in both cases. Both total energies are subject to the conservation of energy equation although they differ by the kinetic energy of diffusion. This implies that the kinetic energy of diffusion is conservative, which is clearly an incorrect assumption. We feel that the logical way to approach the subject is to define one total energy and two internal energies, one of which includes the ki- netic energy of diffusion. 2 v to be 2 (l/2)mdy:, where ma is the mass of component a, a=l Ya is its velocity, and v is the total number of components. Others, particularly practitioners of TIP, take the kinetic energy of the volume element to be (l/2)my , where m is the mass of the volume element and y is the velocity of the cen- We take the kinetic energy of the volume element v ter of mass, mv = Z maya. The difference between the two . V “:12 2 v 2 . energies, E (1/2)ma(ya — y ) = 2 (1/2)ma(ya - y) , is the a=1 a=l kinetic energy of diffusion. 12 It is the definition of two internal energies which presents a dilemma in the postulate of local state. To which internal energy does a Gibbsian equation apply? Cer- tainly the energy internal to the macroscopic volume ele- ment under discussion includes the kinetic energy of dif- fusion. However, we know that the internal energy of the volume element changes due to work done on or by the volume element and to heat transferred to or from the body. This, too, would imply that internal energy includes the kinetic energy of diffusion. The internal energy is a function of three state variables (ng;, temperature, pressure, and composition) in a binary fluid mixture; if kinetic energy is part of this energy, then it, too, is a function of three state variables, which seems unlikely. The postulate of local state obviously deserves the further investigation which we present here. 2. Thermodynamic Process and the Equations of Balance In order to describe a thermodynamic process in any system, it is necessary to know how the several func- tions which are characteristic of the system change spa- tially and temporally and to know the balance equations which the system obeys. Bartelt (1968), whose notation is slightly different from ours, gave as the generaliza- tion for mixtures of the Coleman and Noll (1963) 13 definition: A process is a thermodynamic process for a v—component mixture if it can be described by the follow- ing set of 50 + 6 functionsB: i. the component densities, pa 0 0 I O a 11. the component veloc1t1es, Vi iii. the specific interaction forces, mg, acting on component a due to interaction with other com- ponents iv. the specific external forces, bi, acting on component a v. the partial stress tensors, ogj vi. the specific internal energy, EI vii. the temperature, T viii. the specific radiative heat supply, r ix. the heat flux, qi x. the specific entrOpy, S xi. the entropy flux, fi' These functions must satisfy the following 20 + 2 balance equations for the system: 1. the balance of mass for each component4, 3Cartesian tensor notation is employed throughout this chapter. Vectors are denoted by a single subscript, v-, while tensors are doubly subscripted, tij' The summa- t1on convention on indices is also used, i;g,, 3 v.T.. = v.T... 1 1] i=1 1 13 4The local time derivative (a/at) is related to the substantial time derivative (d/dt) by: (d/dt) = (B/at) + vi(3/3xi). 14 a - dpa/dt + apauk/axk + pa(3vk/3xk) - 0, (2.1) where u: is the diffusion velocity of component a defined by c _ a _ ui — V1 V1' (2.2) where vi is the barycentric velocitys, __ a ovi - Zpavi. (2.3) o = 200‘. loan: = 0. (2.4) 2. the balance of each vector component of linear momentum for each component, a a _ . a a Boavi/at + 8(9 v - oji)/8xj - pmii-pabi, (2.5) c a a j 1 .v. 3. the balance of angular momentum for the mixture, 0 — ZTij — 0, (2.6) where sz is the antisymmetric part of the partial stress a 13 a a a Tij — (1/2)(Oij - Oji)’ (2.7) 5 v 2 = 2 , where v is the total number of components c=1 in the system. 15 4. the balance of internal energy, p(dEI/dt) + aqk/Bxk = pr — pimfiu: + Zo:j(3vg/3xj). (2.8) Equation (2.8) results from the equation of conservation of total energy after the balance equations for potential energy and kinetic energy have been subtracted. The "total energy" of Truesdell or Bartelt and that considered here includes only the internal and kinetic energies: 01 _ a E — EI + Zpavivi/Zp. (2.9) The balance equation of this "total energy" is _ a p(dE/dt) + 8(q3 - viTij)/8xj — pr + Zpavzbi, (2.10) where q; is a heat flux and Tij is the total stress tensor defined by T.. _ a _ a a 13 — 2(oij pauiuj)° (2.11) .This total stress tensor is identical to the total stress tensor of the microscopic theory of Bearman and Kirkwood (1958). The balance of kinetic energy is derived from (2.4) with the help of (2.3), with the result (1 a. co: 0). 9d()oavgvi/29)/dt = 3[Z(oji - paujuiwi + (1/2)zpaugu:u:]/3xj - §c§i(avg/axj) co as + pimiui + {pubivi (2.12) 16 Subtraction of (2.11) from (2.10) gives (2.8), wherein the two heat fluxes are related by: = * a _ a a a a a a qj qj + 2(Oji pcujui)ui + (l/2)Xpaujui i‘ (2.13) Only conservative external forces are considered here; thus, in terms of a specific scalar potential, Va: b“ = - av /8x (2 14) i a i' ' The definition of internal energy is such that the internal energy does not include any external force terms. Inclusion of the conservative external forces is achieved through addition of terms to the chemical potential of each component. If terms for kinetic energy or external fields are included explicitly or implicitly in definitions of partial specific energies, it must be realized that the sum of these energies will not give the internal energy alone, but rather the internal energy plus all or some portion of the kinetic energy and of the potential energy. The sum of the momentum balance equations (2.5) gives the total momentum balance equation which does not depend on the component interaction forces, mg. This means that the interaction forces are not all independent: {m9 = 0. (2.15) The validity of these balance equations is generally accepted. One may work either with internal energy including 17 or not including the kinetic energy of diffusion so long as he uses the balance equations appropriate to his choice. The greatest confusion arises in reviewing TIP, where the kinetic energy of diffusion is generally included in the internal energy and, therefore, is implicit in all partial quantities. Likewise, if one adopts the Gibbsian equations for enthalpy and the free energies, he must specify to which "internal" thermodynamic function he is applying them. To specify a thermodynamic process only the 40 + 5 a a 0 functions, 0 , v., m.,o:., E 0i 1 1 13 I' T, 91! Sr and f1, must be specified while the v + 1 functions, b: and r, can be ob- tained in terms of the other functions by use of the balance equations (2.5) and (2.8). We now make several identifications between the present macrosc0pic theory and the microsc0pic theory of Bearman and Kirkwood. The partial stress tensors of Bear- BK a man and Kirkwood, o ., are related to the macroscopic in partial stress tensors by 013' = Oij ‘ pcui BK “ a a 9. (2.16) J The microsc0pic theory excludes radiative energy source terms which are included in the macrosc0pic theory. The . . a conservative forces of Bearman and K1rkwood,Xi, are re- lated to those of Bartelt by (1"- = xi - 0(3Va/3xi) M b. a 1’ (2.17) 18 where Ma is the molecular weight. Likewise, for interaction forces, the relationship is (1)* _ a Fic - (p/pa)Mami. (2.18) Thus, the equation of Kirkwood and Bearman, apavi/at — 8i cji pa(vjvi Vjvi vjviH/Bxj -(1)* a + Ca ia + caxi, (2.19) where ca is the number density of component a, is identical to (2.5) when the identities, (2.16), (2.17), and (2.18) are used. 3. Constitutive Assumptions A material is defined by a constitutive assumption which is a restriction on the processes admissible in a body of that material. A set of constitutive equations completely describes the response of the system to varia- tions of the independent variables. A mixture of visco- elastic materials susceptible to diffusion and heat con- duction is completely determined by the values of 3v + 4 . a a a . functions. AI, S, qi' ki’ gij, Tij, and mi, where AI is the specific internal Helmholtz free energy, and gij is the symmetric part of the stress tensOr, 09., 13 a _ a a gij - l/2(oij + oji). (3.1) 19 The flux, k1, is the difference between the total entropy flux and the entrOpy flux due to heat flow, ki = fi - qi/T. (3.2) The 30 + 4 functions selected here could be re- placed by any equivalent set of 3v + 4 functions. The choice made in this section is motivated by those terms which appear in the balance equations of Section 2 and in the entrOpy balance equation. According to Bartelt's definition, a binary mixture of fluids is one whose constitutive functionals are func- tions of the independent variables 91.02.T,(301/3xi).(apz/Bxi),u1.dij.dij.wij. (3.3) where dgj and wij arise from the separation of avg/3x into its symmetric and antisymmetric components. The i symmetric tensor is: a _ a a dij - (1/2)(8vj/8xi + avi/ij) (3.4) and the antisymmetric tensor is given by mi]. = [9/2(p - pmnuavg/axi - avg/3x3.) - (avj/Bxi - aviaxj) \) B _ B + (l/p)8£1[(apB/3xi)uj (BOB/ij)ui]} (3.5) 20 or by “ij = (1/2)[3(pu%/02)/3Xi - 3(oui/02)/3xj]. (3.6) The terms added to the antisymmetric components of avg/3xi change neither its symmetry nor its objectivity. This very complicated definition leads to a very simple relationship between wij and w?., which are not independent since 13 w.. = mi. = - w?., (3.7) 13 1] 13 and hence, l _ l _ l wij mi] — (1/2)(8vj/8xi BVi/axj) - (1/2)(3v§/axi - avg/axj) = Zwij.(3.8) Thus, in the difference between the antisymmetric tensors (which appears frequently in subsequent sections) the terms which were added to the antisymmetric component of avg/3xi have cancelled out. The variables of (3.3) satisfy the principle of objectivity, which means that they are inde- pendent of any reference frame. Extreme caution-must be used in the manipulation of equations which contain these variables, for, while avg/axj is objective, v: is not be- cause an observed velocity depends on the observer's ref- erence frame. We want to keep our equations in terms of objective expressions. The entrOpy and entropy flux are related by the balance equation 21 p(dS/dt) = - afi/Bxi + p0 + pr/T, (3.9) where 0 is the rate of specific entropy production not due to external radiation sources. The function, 0, can be obtained for any specified thermodynamic process from the constitutive equations for S and fi and from the energy balance equation (2.8) used in conjunction with other constitutive relations. The second law of thermodynamics here becomes the mathematical statement 9 3 0, (3.10) which is known as the Clausius-Duhem inequality or the entrOpy inequality. This inequality must hold for every admissible thermodynamic process. The internal Helmholtz free energy, A , is defined I by A = E - TS. (3.11) Substitution of this expression into (3.9) and use of (3.2) and (2.8) gives pT¢ = - pdAI/dt - pS(dT/dt) + T(8kj/3xj) 0(0) 0.0. (qj/T)(8T/8xj) pijuj + {dijgij 2w..T... (3.12) 1] 31 22 Equations (2.7), (3.1), (3.4), and (3.7) have been used to obtain the last two terms. Both of the derivatives,dAI/dt and aki/Bxi, can be expanded in terms of the independent variables. After this procedure all terms which do not satisfy the second law postulate are dropped, i;g;, we drop each term in which the independence of some particular variable allows the term to be either positive or negative when all other independent variables are set equal to zero. These re- strictions reduce the constitutive equation for A to a I function of five variables: AI = AI (plrpzIapl/axilapz/axilT)0 (3.13) In particular, AI is not a function of the diffusion velocity, ui. Another requirement of the Clausius-Duhem inequality is S = - aAI/aT (3.14) as is found in thermostatics. (Some equations of thermo— statics are given in Appendix A.) Chain-rule expansion of dAI/dt and Ski/8x1, intro- duction of the conservation of mass (2.1) in the form _ _ (I _ 0: expansion of Bug/8xi in the manner 23 V 8 9 a . = - 8. + 6?. - 1 d9. “3/ x1 (0 pa)wlj/0 13 ( /p)8£198 13 V _ 8 (l/p)8£l(3pB/3xi)uj, (3.16) and use of equations (2.4) and (2.6), and of the restric- tions for the Clausius-Duhem inequality permit (3.12) to be written as pT¢ = - 9 TL (3A1 Sol - _£_3AI 8oz p 1 D1 D1 5xi p2 3p2 5xi 1 1 AI 82p 8k 829a 2" T’pa ataxl + ET[ 89a ]'8‘_x 3x 8 x. a 5x. 1 J + T 3ki _ Si 3T + T 3ki 396 T T xi pa 5x1 2 3k. 3k - E—-m1ul + TEE—w.-L w.. + T ——J--dl Q2 1 1 p Sui 31 aui 13 3A p 8k. I a _g 1 a + a a + 29 3o padijdl] ZTgl anl dij 28Iijdij 1 -23k iapl-i—apzpul+2'rlw p Bui p1 5xj 92 5x3 .1 i 1] )1 3kg ad“ 8kz 801 + T a 8x + T EETT'_§§1‘ (3'17) adij 13 1 4. Ordinary Binary Fluid Mixtures An ordinary binary fluid mixture is defined as one whose constitutive assumptions are linear in the indepen- dent variables 24 1 1 2 (Sol/8x1),(8p2/3xi),ui,dij,dij,wij,3T/axi. (4.1) This definition includes all the variables given in (3.3) except those which are scalars. The constitutiye func- tionals are restricted from being dependent on tensor or vector products of the independent variables. A tensor of one rank can only be a linear function of tensors of the same rank. Thus, vectors will be linear in the in- dependent variables which are vectors and will have scalar coefficients which are functions of the independent scalar variables. Similarly, tensors of one symmetry will be functions of the variables which are tensors of the same symmetry and will have scalar coefficients. Scalars can only be functions of scalars. Applying the principle of equipresence and elimi- nating immediately from each constitutive relation all variables which are not of the same tensorial rank or symmetry, we write the constitutive equations AI = AI(pl,p2,T) (4.2) S = S(plppsz) (4.3) qi = - CqT(3T/3xi) - qu(8pl/8xi) - Cq2(302/8xi) - C u; (4.4) 25 H mi = - CmT(8T/3xi) - le(apl/8xi) - Cm2(8p2/8xi) l - Cmuui (4.5) k. = CkT(8T/3xi) + Ckl(3p1/8xi) + Ck2(8pa/8xi) l 1 + Ckuui (4.6) 1—— Tij - uwij (4.7) 59...-“ +§¢ Ba +§2 {d8} (48) ij 0 ij 8:1 aBdkk ij 8:1 ”as ij ' ° where dfik is the trace of dgj and {dgj} is the traceless part of dgj' Separation of dij into these two parts is significant physically in that the trace of dgj is re- lated to dilatation and the traceless part of dgj is re- lated to shear. Inclusion of the first term on the right hand side of (4.8) is required because 6ij is the ever- present unit tensor and, since it is symmetric, the sym- metric tensor Efj depends on it; — "a is a scalar co- efficient. The Clausius-Duhem inequality when applied to (3.17) with the linear constitutive equations for ki and AI requires that ki not be a function of the density gradients or temperature gradient; therefore, (4.6) be- come 8 _ 1 _ k- — Kuir Cku —VK. (4.9) Insertion of the constitutive equations (4.4), (4.5), (4.7), 26 entropy production equation 0T9 where U l ’11 ll :13 ll _ l 1 - D + Y + Fp1(8pl/axi)ui + Gpl(3p2/3xi)ui C l(aT/axiHapl/Bxi) + qu + Hd}. + 18?. ii 11 (CqT/T)(3T/3Xi)(3T/3xi) + [(T/ol)(3K/3T) 2 1 + Cqu/plT + p CmT/plpzlplui(8T/3xi) + (02/02)cm unilu i 2 2 21¢aBd££dnn 2 0:18 = + Zuwijw ji 2 Z 20 a=l B=1 a B (D/pl)(3AI/Bpl) + (T/pl)(3K/301) 2 + 9 le/plo2 - TK/oo1 - (9/92)(8AI/392) + (T/pl)(3K/392) 2 + o C mz/oloz + TK/oo2 TKoz/o + 991(3AI/391) Tr1 (4.8), and (4.9) into (3.17) leads to the (8T/8xi)(302/3xi) (4.10) (4.11) (4.12) (4.13) (4.14) (4.15) 27 I = - TKpZ/p + ppz(8AI/8p2)--1r2 (4.16) The entropy production (4.10) is a minimum at equilibrium. From equations (4.10), (4.11), and (4.12) we see that it can be written as a function of twenty-one variables, XA' _ 1 2 1 1 XA - (aT/axi'dij'dij'wij'ui)’ (4.17) and the condition for equilibrium is that every XA be equal to zero. As Bartelt and Horne (1970) have noted, vanishing of density gradients and the barycentric velocity is not required for equilibrium. The minimum entropy requirement can be eXpressed by the mathematical conditions (30/8XA)0 = 0 (4.18) 2 (3 ¢/aanxB)o > 0, (4.19) where the subscript 0 indicates that the derivative func— tion is to be evaluated at equilibrium. Applications of (4.18) and (4.19) to (4.10) impose certain restraints on the coefficients of the bilinear form of the entropy pro- duction. Two of these restrictions are H = 0 or HI = 001(3AI/3p1) - Tsz/o (4.20) and 28 I = 0 or n2 = 092(8AI/8p2) + TKpZ/p. (4.21) From (4.20) and (4.21) one can identify the sum of VI and n2 as the pressure, P, W1 + “2 = 991(3AI/391) + 002(3AI/302) = P. (4.22) where the thermostatic definition of pressure has been used. The coefficients ml and 82 thus have characteristics of par- tial pressures. Other results include cql = cqz = o (4.23) CqT : 0 (4.24) n 3 0 (4.25) 911 1 0 (4.26) 022 1 0 (4.27) 411422 — (1/4) (412 + 621% _>_ 0 (4.28) 011 3 0 (4.29) 022 1 0 (4.30) 011022 - (1/4)(012+n21)2 3, o (4.31) Cmu 1 0 (4.32) 2 F = 0 or p le/plpz = TK/Di - (T/ol)(3K/3Pl) 2 — fil/pl (4.33) 29 G = 0 or pzcmz/plo2 = — (T/pl)(aK/apz) + ”2/63 (4.34) pZCqTCmu/pzT - (1/4)[T(8K/8T) + Cqu/T + pzcmT/pzlz 3 0. (4.35) These restrictions reduce the entropy production (4.1) to the form pT0 = D + Y. (4.36) Moreover, (4.23) shortens the constitutive equation (4.4) to _ _ l _ qi — cquui ch(eT/axi). (4.37) Equations (4.33) and (4.34) make it possible to eliminate the coefficients le and sz from (4.5). The complete set of constitutive relations is then (4.2), (4.3), (4.7), (4.8), (4.9), (4.37), and (4.5) with (4.33) and (4.34). The ap- parent lack of symmetry with respect to the two components in equations such as (4.13), (4.14), (4.33), and (4.34) is due to the non-independence of the variables u: and u? and 1 can be eliminated by retaining both variables. From the Gibbsian equation of thermostatics for two components dAI = PdV - SdT + (pl - u2)dwl, (4.38) where “a is the Specific chemical potential of component a and V is the Specific volume, V = l/p, (4.39) 30 the Euler equation Ewan“ = AI + P/p (4.40) can be derived. Equations (4.38) and (4.40) yield a thermostatic identity, pa = AI + p(8AI/8pa) DB’T (4.41) We assume that this is the definition of chemical potential in nonequilibrium processes. Using (4.20), (4.21), and (4.41), we find p1 — p2 = fll/pl — “2/02 - TK/Ol, (4.42) which will be most useful later. The constitutive equation for mi is particularly interesting. By (2.5) and (2.15) 1 _ 1 _ 1 _ 2 pmi _ dolui/dt [(02/9)(30ji/3xj) (91/9)(30ji/3xj)] l 2 + Ji - (0102/0)(bi - bi), (4.43) where Ji consists entirely of terms of the order of velocity squared, 1 l = , 8 8 , . 3 . 3 . Ji plul( Vj/ x3) + plu3( vl/ x3) 2 l l - ((01 - 02)/0102](301ujui/3xj) 1 1 (olozuiuj/O)[(1/ol)(391/3xj) 2 3 - (01/02)(302/3xj)]. (4.44) 31 Equation (4.43) does not define mi, as we have no con- stitutive equation for the external forces, but in the event that there were no external forces it would be a defining relationship. The difficulty in carrying out an explicit analysis under the restriction that there be no external forces lies in the fact that while the constitu- tive equations are required to be objective, the balance equations are not. The balance equations obey the princi- ple of the invariance of work under changes of frame. The external forces which appear in the momentum equations are, in fact, only apparent external forces. True forces are defined in an inertial frame, and the momentum balance equations hold true in those inertial frames where real forces are exerted. A detailed discussion of inertial frames, forces, and apparent forces is given by Truesdell and Toupin (1960). Because we have written a general formulation which holds in all frames, it is impossible to require that the external forces be objective or, in fact, simultaneously zero in all reference frames. The right hand Side of (4.43) appears to be non- linear according to the definitions of this section. Con— stitutive relations could be introduced for the stress tensors, and this would result in inclusion of second derivatives of the component velocities. The terms con- taining the barycentric velocity are even more unusual since they lead to terms containing three or more variables 32 when the gradient of the barycentric velocity is expressed in terms of the objective independent variables chosen in Section 3. Comparing (4.43) with the constitutive equation for mi, (4.5), we see that the linearity of mi there is in distinct contrast to the right hand Side of (4.44). The first term in (4.43) is called an inertial term and all the terms in Ji are called viscous terms or non-linear terms. Clearly, no one-to-one correlation can be made be- tween (4.43) and (4.5), but (4.43) places restrictions on m; which will be utilized later. 5. Transport Equations in the Independent Variables T, 0, W1 The transport equations given in this section follow algebraically from the equations of the previous sections. The independent thermodynamic variables considered in this section are the more commonly used p, WI, and T rather than pl, 92, and T. These transport equations are equivalent to the 20 + 2 balance equations of Section 2, where in the binary case we have 0 equal to 2. The balance of mass equations (2.1) are replaced by and .l p(dW1/dt) = " BJi/BXi: (5.2) where j: is the diffusion flux, 33 .1 3i = pluil. (5.3) The balance equation for linear momentum of the system as a whole is found by summing equations (2.5) over a to obtain p(dvi/dt) - 82(ogi - paugu§)/8xj = {pab:. (5.4) The form 0(dvi/dt) = - (B/B')(8T/3xi) - (l/DB')(80/3Xi) + Zpabg + BHji/axj (5.5) is obtained by simultaneous addition of aP/axi and sub- traction of its equivalent form, aP/axi (B/B')(8T/3xi) + (l/OB')(30/3Xi) + [0(\7l - V2)/B'](3wl/8xi). (5.6) The viscous pressure tensor, Hij' is defined as _ a _ o a = H.. — {(c.. p uiuj + "65ij) Tij + Pdij. (5.7) Here 8 represents the thermal expansivity, B = - (l/p)(39/3T)P'Wl' (5.8) 8' is the isothermal compressibility, 34 ' = B (1/0)(89/3P)T’w1. (5.9) and Va is the partial specific volume of component c. There exists another momentum balance equation corresponding to the second of the two balance equations introduced initially in (2.5). It can be termed the balance equation for diffusion momentum, dji/dt. One form of this equation is found by rearrangement of (4.43). d.1 31 = 33 3 01 -‘:l a 02 - J + m at p 5xj ji p 5x3. ji 1 p 9 p + 1:2 (b: - bi). (5.10) In terms of the variables of this section, Ji [see (4.44)] has the form . 1 J = _ (pl - 02) Bjjji + 8v jl + 31 3vi l 0102 8xj xj 1 j ij + j 31 (92 " 91) a i x. J 09192 3 8w 2 2 1 + 2 (92 ‘ 0291 + 91) SET ' (5°11) 9192 3 If the constitutive relation for mi, (4.5), is introduced in (5.10) with the use of (4.33) and (4.34), the balance equation is 35 1 dj. 9 1 = _ .1 _ _g$ _ ._ . plpz at Du3i DT axi 8T(“1 “2)/3xi 1 +.;L 3(0ii + nldij) pl ij 2 a .. + .. - 41 (031 "2 11) — —Jl— J . (5.12) 02 3"j 0192 1 The coefficients Du and D in (5.12) are functions of T, T p, and W1' only: _ 2 2 Du - (p /0102)Cmu (5.13) DT = (02/0102)CmT + K/pl + (T/pl)(8K/3T) - (§1"§2)° (5.14) The chemical potential, ué, includes external potentials, which have been introduced by means of (2.14), 0' = u + V . (5.15) The gradient of the chemical potential difference, which appeared in the form of (4.42), has been replaced by _ I ___. _ u 3(ui uZ)/8xi 8T(ui “2)/axi — (Sl - Sz)(3T/3xi), (5.16) where Sa is the partial specific entropy of component a and where I Bxi 8 3x1 — ._ 2 1 B 02 8x (V’ - V ) l 2 8p + 08‘ Bxi (5.17) with “11 = (8u1/3w1)T,P. (5.18) The conservation of moment of momentum or angular momentum is Simply the requirement that the total stress tensor be symmetric, a a a _ a a o {(6ij + pauiuj) — 2(6ji + paujui). (5.19) The balance of energy can be replaced by a tem- perature equation by means of identities from thermostatics. Either of the following temperature equations containing heat capacities can be used: — dT _ dS 8T g3 pCv—E-pT—Tt-+——TOB t _ 8(V-V) dw - pT (s — § ) - l. 2 1, 1 2 8 St (5.20) where C; is the specific heat capacity at constant volume, or dB 91%“ (0'6P - PB) 3% - (TB - PB') 3% dw where C? is the specific heat capacity at constant pressure and Ed is the Specific internal energy, Ea = Ha - PVa == pa + TSa - PVa (5.22) The balance of entropy, (3.9), can be used in (5.20) with any choice for the entropy production. The internal energy balance equation (2.8) can be used in (5.21). Comparison of the resulting two final equations is easily effected with the help of the thermostatic relationship ___ 2 ' CP - Cv + B T/oB . (5.23) The results of this comparison could be taken to be a measurement of the accuracy of representing the quantities which appear in the entropy production equation by con~ stitutive equations, Since (5.21) is exact while (5.20) contains the entropy production which is inevitably ap- proximated to some extent by the linearity of the con- stitutive relations. There is also the possibility, however, that the thermostatic equations (5.20) and (5.21) are not applicable in the nonequilibrium case. Equation (5.21) is indeed an equation for internal energy but in thermostatics this im- plies the internal equilibrium energy of some volume 38 element. In the nonequilibrium case there are three ways in which we could interpret this. Firstly, the internal energy of (5.21) may be only the equilibrium part of the total internal energy which in the nonequilibrium case contains both equilibrium and nonequilibrium parts. Secondly, the internal energy of (5.21) might be the nonequilibrium internal energy plus the kinetic energy of diffusion, Since the sum of these two energies is the total energy internal to the volume element being de- scribed. Finally, the energy of (5.21) could be the non- equilibrium internal energy, which would be consistent with our other usage of the symbol EI in this chapter. If the first interpretation were true, an arbitrary sepa- ration of thermodynamic quantities into equilibrium and nonequilibrium parts would have to be made. In the second interpretation the partial specific thermodynamic quanti- ties would be defined as previously with additional terms for the component kinetic energy of diffusion. Ramifica- tions of these possibilities will be discussed in Section 8 after a survey of the possible forms of the entropy production which can be used in (5.20). 6. A New Rational Mechanical Derivation The theory of mixtures presented in Sections 2 through 5 is a typical example of a rational mechanical develOpment. It is only one of many theories which can 39 be derived. The particular advantage of that theory is the very concise bilinear form of its entrOpy production. In this section we derive another theory in a similar manner, but in this instance we obtain more familiar vis- cosity coefficients while the entrOpy production equation becomes much more complicated. The Sv + 6 functions which define a thermodynamic process in mixtures need not be those of Section 2, but can be any set of equivalent functions; and, likewise, the balance equations may be written in many forms. In this section we choose to let our independent variables be 891 8p2 8T avi Bu. sym (6 l) D ID ITIu I ”1"“! I_I I“): or o l 2 i axi axi 5xi axj xi 1] where _ 1 _ 2 _ l _ _ 2 One can readily verify that asym _ These variables are objective and more nearly symmetric with respect to the two components than the variables of Sections 2 through 4. We have adOpted here the diffusion velocity and tensors of Doria while retaining the component densities 40 and density gradients of Bartelt (3.3) and Mfiller. Doria used p and W1 and their derivatives in place of the com- ponent densities and their derivatives in (6.1). Mfiller's variables differ from Bartelt's in that he used the dif- fusion velocity, ui, and an antisymmetric strain tensor, Zwij. The balance equations, written in terms of these variables when possible, are: 2 2 dpl — _ 3% u 8pl _ p1 u 8p2 - plp2 8u. - 9 8V] at - 8 _2 x. x. l x. 9 3 x3 9 j a J p a J a J (6.4) d92 93 391 9% 392 + 9192 Buj Vj at = _2 u. x. + _2 u. x. x. - p2 x. 9 3 l 9 3 l p a J a J (6.5) dv 1 _ 8 l 2 p182 dul 0192 3V1 91 2 ani T-at+Tuj63:.-+ 2‘92‘91’“j6‘i‘ J 9 j 2 a 2 a _ 9192, pl + 9192 u u 92 3 1 8x 3 1 x 9 3 J 9 3 a J 31-- 9 39 9 89 = ._l£'- l'T' _1£.+ .3 T' __l ~ x. _2 i x. 2 i' x. 3 J 9 j J 9 3 3 D192 2 ____ 1 + 9mi + 9 (bi bi) (6.7) 41 dB 8v. 8q$ p p B _ 1 _ l 2 l _ 2 p at _ 8xj Tij xj + prl+ p ui(bi bi)° (6'8) The total stress tensor, Tij’ is that defined in Section 2, and Tij is defined by I Q + Q ' = = aa- Tij Tij + (p102/9)uiuj Tij + {pauiuj (6.9) The second stress tensor, Tij, arises in the rearrangement of the momentum balance equations and has the definition Tij = (92/0)9ij - (pl/9)0ij. (6.10) This stress is a measure of the difference between the component stresses per unit mass. By writing 02 1 ' 1 I " (pl/p)( ij ij 13 ij ), (6.11) we see that Tij represents the difference between com- ponent 1's contribution to the stress and the ideal con— tribution of component 1 to the total stress if the stress of components 1 and 2 were the same per unit mass. The stress tensor is, of course, also just the negative of the difference between component 2's contribution and the ideal contribution to the stress. The specific inter- action force m1 is the same as m: of Sections 2 through 5. The energy which appears in (6.8) is the total energy de- fined in (2.9), less the bulk kinetic energy, so 42 _ 2 EB - EI + (plpz/Zp )uiui. (6.12) In a manner completely analogous to that of Section 3, the entropy production can be obtained in terms of the set of variables (6.1). It has the form pT0 = p(dAI/dt) - pS(dT/dt) + T(8kj/8xj) (qj/T)(8T/8xj) - pm.u. + T + sym - (pz/pz) Tijuj(8pl/8xi) + (pl/p2)Tijuj(8p2/8xi). (6.13) The heat flux, qj, is still that of Bartelt, but it can now be written as — 2 — qj - q; + uiTij + (0102/20 )(pl pz)ukukuj. (6.14) After examining the balance equations and the entropy production, we choose to write constitutive equa- tions for AI, S, Tij'3Tij’ qj, kj' and mj. The derivatives dAI/dt and akj/axj in (6.13) must be expanded in terms of the independent variables, so that we then have dAI 3A1 dpa 8A1 d 3904 ' p t ‘ ‘ 92 Spa dt ’ Z 8pc )8? 8x. 8 5x1 _ aAI dT _ aAI d 8T _ ”‘1 cm1 "‘TT‘8E “’3 8T 8? 83;“ p 86; ‘8? 8x1 8vi . r) 8AI dwii _ 8AI x. awij dt 9 vi 8t " 8 5—— x. 3 8A 8u _ I ‘41 i 8 8x. 3 and 2 T8ki = 2T‘8ki)8pa + 2T 8ki\ 8 pa + T 8ki 8T 8x. 8p 8x. 48p )8x.8x. 8T 8x. 1 c 1 a 1 j 1 8x. 3 + T 8k; a2T + T 8ki 8u. sym + T 8ki w 8T 8x.8x. 8n. x. 8n. ij %8x. 1 3 3 1 J 3 8k 82u. 8k 82v 8k 8w.. + T l) 1 + T x \ i + T g 11 8ui 8xj8x2 8vij8x£8xj Bwij 8x2 ° 8xj *8xj (6.16) The coefficient of any derivative of an independent variable in (6.13) must be zero since that derivative could be either positive or negative. However, several of the ‘ 44 variable derivatives in (6.15) and (6.16) are not inde- pendent. The time derivatives of component densities and of component density gradients are given by the balance of mass equations. For dpl/dt and dpz/dt we use (6.4) and (6.5), while for the time derivatives of the density gra- dients, we use d(8pa/8xj)/dt = 8(dpa/dt)/8xj - (8vi/8xj)(8pa/8xi) (6.17) as well as (6.4) and (6.5) to write 30 Eu oz an 02 $0 29 9 Bo Bo d 1_ i _g 1 + _1 2 _ 1 2 1 2 a? 8x. 5x. 2 8x. 2 8x. 5 L‘li x. x j 3 p 1 p 1 p 1 3 2 2 2 2 _ 29192u 392 391 o2u 3 p1 o1u 3 92 p3 u15xl 8x3 pZu i8 x. 8xj p2 ui8xi8xj 2 p2 ap1 ap1 2‘)1 3p2 3‘)2 301 3V1 + 2—3 18x. 8x + §Ji8x 8x _ 28x. 8x p 1 p i j 1 2 2 2 - 8 vi _ p1 8p2 8ui - p2 8pl 8ui p15x.8x. _2 8x. x. .2'8x. 8x. 3 1 O 3 1 D J 1 2 0192 3 u +, p 3x.“ (6.18) 1 d 3‘)2 _ aui fig 301 + p1 3p2)+ 20192 ap2 301 ‘3? 8xj 8xj p2 8xi 02 8x1 3 18xi 8x 45 2p192 8p1 3p2 Di 82‘)2 + u. + —u. 3 15x 5x. 2 15x.5x. o i J o 1 3 9% 32.1 29E 902 391 + —2ui5x.5x. - 3ui5x. 8x. 0 l J O l J 2 2 292, 8pl 8p2 28p2 8vi 8 vi 34i8x. 8x. - 8x. 8x. - p28x.5x. 3 1 J 1 o 1 J 2 2 2 + p1 ap2 aui + fig 801 aui _ 0192 a “i -2'8x 8x 5x. 8x. p 8x.8x. ’ D 3 i D J 1 1 3 (6.19) The strain tensor derivatives are not all independent either since one can readily show that = sym _ sym Bwij/sz 8(8ui/8x2) /8xj 8(3u2/8xj) /3xi. (6.20) Using this identity we write sym 8kg 8 8ni + 8k£ 8w.i 8u. sym 8x 8x. 8m.. x 3 1 2 3 13 2 8x. 3 31:1 8k. 8ki 3 ani sym = T111 sym + Flea“ + 801?]. 3x1 8x3. (6'21) 8 5;; Recognizing that these substitutions must be made in (6.13), we invoke the Clausius-Duhem inequality in order to set equal to zero all those coefficients of derivatives of independent variables which can be positive or negative 46 while all other variables can be set to zero. Thus, the coefficients of dT/dt, d(8T/8xi)/dt, dui/dt, d(8vi/8xj)/dt, d(8ui/8xj)8ym/dt, dwij/dt, 82ui/8x18xj, 82T/8xi8xj, and 82vi/8xj8x£ must be set equal to zero. An immediate consequence of this is that AI’ as before, cannot be a function of 8T/8xi, ui, 8vi/8xj, (8ui/8xj)sym, or wij' The normal thermostatic relation- ship between entropy and the Helmholtz free energy is found as before, These results reduce the entropy production to the form %3 8pm dt 8xi 89a 8x1 8x1 8k. 2 sym 1 8 pa 8ki 8T 8k£ 8ui + 2T 8p 8x 8x + T 8T 8x + T8u 8x a i j i 1 2 8x. 3 _ sym + 8k£ T 8ki + 8k 8ki 8 8ui 8u.w£1 u. 8w . 8w . 8 8x. 1 43 1 21 £3 2 3 L 8x. 3 1r 8xj 9 3 3 2 lj i8x 2 ij 18x 8vi 8ui sym 47 If we consider the coefficient of 82vi/8xj8x , we obtain the following restriction: p1 6ij + 929—27334” 8x fiw) 8x 8x: (6. 24) This can be integrated at once to give 8AI 8AI 8v. _ _ .______ 1 k2 - N1 p[p133:l + p23 3:: J 8x , (6.25) where N1 is independent of 8vi/8xj. Now, if this result is substituted into the coefficients of 82p1/8xi8xl and 82p2/8x18x1 which themselves must be equal to zero, two further restrictions ensue, 2 2 2 2 838A1u+p£ 811111-12- 89.1‘1111318111In p a(501,1 o a 501 j o a 392 o a 592 3' x3 x1 8xj 8xi 8Ni 3N. + T—s—p-f- + T—Tai- + T[291‘:3;—:[3:1 8 5;; 8 5;; 82% 8215.1:L 8vk 301? + p92 301 302 8xk = 0, (6.26) %( 8 §—— 8 -—- xj 8xi and 48 2 D2 2 31$. ___.____ -3111, D 8 8p1 1 93(8913 p 8p2 1 8x. x. J 8 J 2 2 _£1 8AIu + 8N].- 3N. 8 p 392 3 392 392 0013 301 3 302 8x1 8xj 8x1 8x1 axj 82AI 82AI 8vk + 2 = . 6.27 pp1a 8p1 8p2 pp2a 8p2 892 8xk ( ) 8xj 8xi 8xi 8xj The coefficient of 8vk/8xk must be equal to zero in each equation so that from (6.26) we can write 891) pp1 8p1 pp2 (892) 891 ppl 891 8 -—— 8 ——- 8 -—- -——- ——— 8xj 8xi 8xi 8xi axj 8AI + pp2 ”1.302 (6.28) a . 8x. 3 An equation of this form has the general solution 8AI 8AI .8p1 ; 8 -——) 8 -—- 8x1 8x1 The function Qij is antisymmetric and neither Qi' nor fi can be a function of 891/8xj. They must, therefore, be functions of 8p2/8xj only. By the objectivity of the 49 scalar, AI' Qij must be zero. From the coefficient of 8vk/8xk in (6.27) we obtain a similar expression in terms of 8p2/8xj, and the two results imply that 8AI 0275—: " 01W. . (6.30) 8 8x. 3 Returning to (6.26), we see that all those terms but the last one, the one in 8vk/8xk, must be zero in- dependently. We rearrange this equation using (6.30) to a form similar to that of (6.28) and then using the general solution we find 8p2 'I'Ni - AIplui — Qij'a'}; + F1 (6.31) where Qij and Fi are independent of 8p2/8xj and Qij is antisymmetric. Considering the terms in (6.26) which are independent of 8vk/8xk and employing the results (6.30) and (6.31), we write . + F. 8(391 AUI i Qikaxk+ i 39 a — 8Xi We multiply both sides of the equation by (8p2/8xi)(8p2/8xj) and take advantage of the antisymmetry of Qij to arrive at 50 a A u 3p2 302 + F1 3:2 392 8p1 I 8x. 8x. fi8xj 8 8x. 3 8p 8p 8p 8p ' a: AIuj 8x? EEE Fj 8x2 8x2 (6'33) 8 l j i J i 8xi Because both sides of (6.33) are scalars, this statement is equivalent to 3 3p2 Bp2 8p2 ap2 —-5—p—1—- AIuj axj 3X1 + Fj 5-;- Fit—i- - 0. (6.34) 8 ——— 3 8x. 1 Therefore, 8AI 8F. 0 = - 6.35 —5-—(u 91 3 —1-3 301 . < ) axi 8x1 and AI cannot be a function of (8p1/8xj)(8p2/8xj). Thus, AI can be an objective function only of (8p1/8xi)2 and (8p2/8xi)2. But we know that p2 8AI 8F. (8p2 E: .53;._ - _—3Bi—.¢ f 5;; , (6.36) is; W:- so AI can be at most only first order in 8p2/8xj and this is impossible. The only remaining possibility is that AI is independent of both 8p1/8xi and 8p2/8xi, so we have shown that 51 AI = A1(pl'°2'T)' (6.37) This result is very desirable in that we would hope that our thermodynamic variables are functions only of the state variables, p1, p2,fm in the nonequilibrium case as they are in the equilibrium situation. This we have shown to be true of the Helmholtz free energy. The result of equation (6.37) greatly simplifies the entropy production equation so that we now have the form 8ki 8p1 8ki 3p2 8k _ __i_ .312 - 31 ET 0T4) — T871 87:1 + T872 8—xi + T 8T 3‘1 T 8?]. 8k. - pmiui + T “1 + Tij wji a a 8k. 8u. sym l 2 8pl 8p2 13 8n. 1) 8x. 8A 8A 8v I I . 1 + p(918p1 p28p2) 613 Tij8xj 2 + 3; 3A1 _ 3A1 5 p2.1.. “391 o 591 592 13 2 13 15xj 2 pl ‘8AI 3AI pl 302 + —— - 6.. + ——T!. u. > 0. 6.38 p 53:' 83; 1] D2 13 i8§§ — ( ) Among the restrictions which are obtained in arriving at this form are: 52 8ki/8(8pa/8xj) + 8kj/8(8pa/8xi) = 0 . (6.39) 8k2/8(8Vi/8xj) = 0 (6.40) T[8k£/8(8uj/8xi)4-8kj/8wzi + 8ki/8wkj] = 0. (6.41) Although further restrictions on the general forms of the response functions could be derived, their physical significance becomes less and less clear. For this reason we now restrict ourselves to a linear theory. The omission of terms which are of second or higher order in the inde- pendent variables will naturally introduce approximation into a theory which until this point has had none. (Re- call that this section is not merely a continuation of the previous one.) The error due to these omissions is minimized by retaining explicit second order dependence of the total stress tensor and of the energy, E, and by using linear functions to represent only those parts of E and Tij which do not have explicit second order depen- dence. As in Section 4, we can immediately limit the con- stitutive equations to functions of tensors of the same rank. We also have the restriction (6.39) and the inde- pendence of A from density gradients which say that ki I cannot be a function of 8pl/8xj, 8p2/8xj, or ui, so the constitutive equations take the forms AI = AI(ol.pz,T) (6.42) 53 k“ = K’”' 6 43 J J ( ) 39 8p = - ' _ I 3T _ | l ' 2 qj Cquu j CqT8xj qu8xj qzaxj (6.44) 39 8p = _ I _ | 3T - ' 1 _ ' 2 “‘3' Cmuu j CmTax. Cm18x.cm2—3x. (6.45) 3 j 3 I = _ | Tij = - p513“ + “1577513' + “253?" + “3W5” n j n 8ui + “fa—“x (6.47) 3' asym = , Tij “ wij (6.48) The primes are to indicate that the coefficients may not necessarily have the same dependence on pl, p2, and T as similar coefficients of Section 4. These explicit func- tions can be placed in the entropy production equation which again becomes a function of twenty-one variables, all of which must be zero at equilibrium. The equations (4.18) and (4.19) still apply for the variables sym uj,8T/8xj,8vi/8xj,(8ui/8xj) ’wji' The final form of the entropy production is then _ . 8K' qu . 8T , 8T 8T pT¢ Cm mu “in + T‘Sfi + T + CmT ujgig + qT8§3 5;; 8v 8v 8vu 8v. k 2 1 + ¢___.___ + 2n——— ——1 + L {5-49) 8xk 8x2 8xj 8xi ' 54 where 8n 8v 8ui 8v. 8 8u _ k 2 “k 9. 1" (X1+°‘1’:—x;53:+ (X2+°‘2"5x—j xi+°‘3'5§;'a§; + SS8uk + u w - E£¢3Vku 891 0L45); 5xi wij ji p2 8xk 15xi + 91¢ 8v vku 892 -‘E% 2n8viu 8p1+plzn8viu 8oz 02 MR: 1‘53}: 92 ”5323’ 183?? ’7 83?? 1332’. __ ‘32 a“): ap1 + 01 auku 3p2 3p1 ple‘r‘“ 2‘53?— pleaxk 2sz "7x2;— “1‘5“ 9 an. S 89 + —ix 1 u 2 (6 50) 2 2 5x. 18x ' ' o J j The positive-definiteness of entropy production, (4.18) and (4.19), has also generated the requirements qu = qu = 0 (6.51) “T393A13?92 8K' +—7P p'q-T g—— + ple = 0 (6.52) D2 0 AI AI 1 8K' T‘s—pl - r02 - :2- P' +T r- + ‘3sz - 0 (6.53) P' - 3A1 + 8A1 (6 54) ’ p pl‘a'a- 9283‘ ' 3A1 p192 _ I _. I _ p-TK +01 ZW- 1:72 -TK + (111 112). (6. 55) 55 Again we have identified the thermostatic pressure, P', in (6.46) and we can use (4.41) for the difference in chemical potential. Equation (6.55) gives us some physical feeling for the coefficient K'. It is dependent on the difference between the specific partial pressures (equilibrium parts of the specific stress tensors) and on the chemical poten- tial difference. Using (6.55), (6.52), and (6.53) we can 0 . | I solve for the coefficients le and sz, 0C1;1 = - (oz/02)P - Sp/apl + (plpz/p)8(ul - u2)/Bol (6.56) ocfiz = (pl/02)P - 3P/302 + (ploz/p)3(u1 - u2)/392. (6.57) Many other restrictions similar to those imposed on the coefficients of Section 4 can also be found. At this point we have only fourteen unidentified coefficients in the constitutive equations. Nine of these are viscosity coefficients of which we readily recognize o and n as the coefficients of bulk and shear viscosity, respectively. We have some feeling for p, but Cfiu, CfiT, C' , and C' remain unidentified. qu qT By using (6.56) and (6.57) in the constitutive equation for internal force and by making transformations analogous to those in Section 3, we can write 56 p192 aT(”1 ‘ “2) p192 — — = — V + - ' _ pmi p 8x1 [ pCmT + p (51 82) + 9192 a(“1'“2) a'r . p2 P+ 3p 3‘)1 8F 8):. + ’2' 8p 8x. 1 p l 1 01 3p ap2 + - — P + - DC. “to (6058) p2 8p2 8xi mu 1 When this equation is put in the balance of diffusion momentum equation, the result is again identical to that of Bearman and Kirkwood except that the coefficients are more clearly identifiable. One form of this equation is 3(u -u ) . .1 _ _ p192 — _ — D192 1 2 pcmuji - [ 9931 + p (51 82) + p 8T 3p 31 D2 3 1 + 571,-] 3x + —5- T'xjwji + 111613) pl 3 2 0102 3(u1 pi) - 77'8x3(°ji + Tr2631 p 8x1 1 + dji + D2 D2 3 (.ljl) at p p2 8xj 3j i .1.1 p ap1 p1 3p2 13V: - Ji3j( 2 8x. - 2 8x + ji x. 091 3 092 3 J 18vi + Jjgig o (6.59) This theory does not vary radically from that of Sections 2 through 5, but it is decidedly different from it. The entropy production (6.49) contains five terms 57 which are predicted by TIP and also those terms in L, (6.50), which would typically be called viscous terms. The constitutive equations are very similar to those of Section 4 but not interchangeable with them. 7. Comparison of Several Theories for Binary Mixtures of (Fluids There have been many approaches to the thermo- dynamics of irreversible processes in binary fluid mix- tures. Both of the rational mechanical theories presented here compare readily with the microscopically grounded theory of Bearman and Kirkwood and with the theories of Pitts, de Groot and Mazur, etc., for nonequilibrium thermo- dynamics. All of these are so-called "linear theories" with functions being represented in a linear manner such as that presented in Section 4. However, some also con- tain "linearization" of other forms including dropping of terms in any equation in which such terms appear to be non-linear. The theory of Bearman and Kirkwood is based upon the same balance equations as those of Section 2, includ- ing partial stress tensors which are defined in a slightly different manner as discussed in Section 2. The TIP theories do not treat partial stress tensors and only rarely include any terms arising from kinetic energy of diffusion. The exact limitations of the balance equations used in TIP and their differences from those used in 58 rational mechanics can best be seen by studying the balance equations of Section 6 which are equivalent to those of Section 2. The theories of TIP neglect entirely balance of diffusion momentum (6.7) and all terms which are introduced in this equation and not in any other balance equations, iifiiv interaction forces and partial stress tensors. The energy, EB, of the balance equation (6.8) is then taken as the thermodynamic internal energy since there is no longer a substitution to be made for the external forces of (6.8) which would introduce kinetic energy of diffusion. It is, thus, obvious that TIP and rational mechanics will not yield the same entropy pro- duction equation. It is interesting to note that the re- sults of a rational mechanical approach to the problem as defined by the TIP balance equations leads precisely to the usual TIP result. This development is given in Appen- dix B. It is also possible to show that rational mechanics and TIP are the same except for kinetic energy of diffusion terms. We do this by rearranging the equations of the rational mechanical theory of Sections 2 through 5. In both microscopic and macroscopic TIP theory the diffusion flux, j;, as well as the heat flux, 01' is written as a 1. linear function of two forces: 2. (7.1) 82.nT/8xi + 911 1 _ Q 10 I U. P l i (20089.nT/8xi + 9012i (7.2) I 0 ll 59 where the flag are Onsager or phenomenological coefficients. In microscopic theory the force, 21, generally takes the form zi = 8T(u1 - u2)/8xi. (7.3) (De Groot and Mazur, by inclusion of kinetic energy of diffusion in their energy, E, admit the possibility of an inertial term, but not of viscous terms, in Zi but do not generally include it in their work.) Equation (5.12) can be solved for j: to yield - j: = (DT/Du)T + (p2/9192)lepluiu%. (3.3) flhe expression for D can be modified by replacing the con- stitutive equation for qi by qi itself to yield 63 D = - (qi/T)8T/8xi + [(T/pl)8K/8T + (92/9192)CmT]plu:(8T/8x£) + (92 /9192)lepl% where .= 2+ q: q: The entropy balance pT(dS/dt) then has the form T(dS/dt) 1u 1 (8.4) a _ a a a 2(Oji paujui1ui + (l/2)Epaug uiui. (8.5) equation = - T(8fk/8xk) + pT¢ + pr (8.6) + + T(8fk/8xk) - qi(8£nT/axi) + pr [(T/ol)3K/3T + (oz/ploz)CmT]plui(3T/ax£) 2 1 l (p /0192)Cmuplu£u2 + Y' b) A second form of the entropy production is ob- tained as in Section 7 by introduction of a new heat flux into the linear equations. Again we have used the con- stitutive equation for mi. Thus, pT¢ = F + Y, where F = - qi(8£nT/8xi) - j (8.8) (8.9) (1] 64 — — .1 Tfi - T(S1 - 82)]i .Q P. - I) q; + 2(agj - p uguq a a a a l J)uj + (1/2)Zpauiuju 0!. j — — I 1 + Tki - T(sl - $2)ji. (8.10) The entropy production is then pT(dS/dt) = - T(8fk/8xk) — jisi - qi(8£nT/axi) + Y + pr. (8.11) c) A third form of the entropy production is ob- tained by returning to the expression for entropy produc- tion before the introduction of the constitutive equations. Equation (3.17) for pT¢ is rewritten using the restrictions from the constitutive equations and (4.41), (4.15), and (4.16): _ _ _ .1 pT¢ - fj(3T/3Xj) [3(U1 H2)/8lejj a _ a X(pmi and/axi)ui + 2(ogj + naaij)(8vg/8xj). (8.12) For ZmEug we multiply (4.43) by pUi/pz and rearrange this eXpression to find a a _ a a _ a a _ a a pZmiui - pd(Zpauiui/2p)dt 8{2ui(oij pauiujH/axj — 8(2paugugug/2)8Xj - )(ogj - paugug)(avi/axj) + 2°:j(3V:/3xj) - {paugbg. (8.13) 65 This uses the second of the two expressions for m: which were discussed in Section 4. Substituting (8.13) and (8.12) into (8.6) and rearranging, we arrive at an entrOpy balance equation 1 j .1 (q; — jj(Hl H2)}89.nT/8xj + Hij(8vi/8xj) ac __. ,_, T o. (9.1) If the coefficients ¢ and Cfiu depend only on p1, p2, and T, it would seem that this expression places a restriction on among the variables p1, p2, T, 8p1/8xi, and 8p2/8xi. At what point did our variables lose their independence? We have assumed that thermodynamic quantities are the same functions of state variables in nonequilibrium thermodynamics as they are in thermostatics. If this postulate is not true, we have a whole new problem. Trues- dell actually gives a chemical potential tensor for each component. Of course, we can define any combination of terms by any name we desire. An interesting point to investigate will be the nine viscosity coefficients that appear in rational mechanical theories. Only two coefficients have ever been measured. The problem of actually measuring stress 74 in a fluid due to a diffusion velocity gradient seems in- surmountable and much work remains to be done in this area. Certainly there are many implications of rational mechanical theory which remain to be investigated. How- ever, the theory needs also to be extended, particularly to include second order terms in the constitutive equations. This is an immense problem which would more than quadruple the number of terms in each constitutive equation. It would produce a more realistic theory, but attainment of relationships between constitutive coefficients would be much more difficult since (4.18) and (4.19) tell us nothing of terms with more than two of the 21 nonequilibrium vari- ables. One would also be tempted to try to formulate a theory by adding only those second order terms which he feels are likely to be important. This would be highly subjective, however, for who among us knows how inter- action forces depend on a velocity gradient dotted into a temperature gradient and so forth? Nevertheless, it will not be possible, except by experiment, to assess the accuracy of the linear theory until a second-order theory is formulated and the contributions of terms added are evaluated. We have seen that terms which are second order in u: have caused the greatest problems in comparing the rational mechanics theories to the theories of TIP. Ideally, one would minimize these contributions in an 75 experimental situation. This, however, implies that the diffusion flux itself will be minimized and that we are approaching the study of a binary mixture behaving as a pure fluid in nonequilibrium. We would be more certain that our experiments were correct as evaluated by the theory of rational mechanics or TIP, but problems of diffusion would be unanswered. If a complete non-linear treatment of rational mechanics is done, many, many more terms which are second order in u: will no doubt arise. It may well be that these terms in combination with the terms which come out of the linear theory will vanish, or will at least form clearly nonlinear combinations. The foundations of TIP are the same as those of rational mechanics except that TIP does not treat the balance of diffusion momentum, and we have established that the theories are identical ex- cept for the second order terms in diffusion velocity. In thermostatic derivations, a change of state variables gives formal results different from the original equations, but each set of results equally well describes reality and gives the same calculable results. One's choice of state variables is usually directed toward giv- ing the simplest formulation for the problem at hand. In TIP one can change his reference frame (reference velocity to which component velocities are compared) and derive several formally different theories. A comparison of the 76 entropy productions of these theories allows one to obtain readily relationships among the various phenomenological coefficients because the entropy production is invariant. We would suggest then that, since all the rational me- chanics theories have been developed from prOper sets of Objective independent variables applied to the same bal- ance equations and since all the theories have been linearized in exactly the same rigorous manner, the en- tropy productions of the theories should be equivalent and any formal difference occurring between them can be equated to zero. This procedure would hopefully give us new insights into the meaning of the constitutive co- efficients or at least giVe us relationships among the coefficients of various theories. CHAPTER III THERMAL CONDUCTIVITY 1. Introduction A great deal of research on thermal conductivity in fluids has been done over many years, particularly be- cause of the diverse technological applications of heat conducting liquids. A review of theories of conductivity, all of which are microscopic, is given by McLaughlin (1964). However, a fully rigorous mathematical treatment of thermal conductivity as a macroscopic process has not been presented in the literature. Solutions of the macro- scopic heat transport equations for liquids have, in gen- eral, been found by assuming that physical properties such as heat capacity, density, and thermal conductivity are constants. These solutions then resemble solutions for heat conduction in solids as discussed by Carslaw and Jaeger (1959). Many experimental techniques have been developed for measuring thermal conductivity in liquids, but most of these depend on calorimetry measurements and measure- ments of heat flow and, therefore, either require calibra- tion or are subject to large experimental error. 77 78 These problems are emphasized in a recent review (Tree and Leidenfrost, 1969) of the methods that have been used to measure the thermal conductivities of carbon tetrachloride and toluene. The values measured by these methods show a variation of more than ten percent for each of the fluids. Such findings suggest that one needs a better theoretical treatment for liquids and more reliable ex- periments. We have completed a theoretical investigation of heat conduction in a flat plate cell (thermal diffusion cell). The theory does not treat coefficients in the thermal conductivity equation as constants, and thus an analysis of the error introduced by such an assumption is possible through comparison with this rigorous theory. The attainment of general macrosc0pic solutions for tem- perature as a function of thermal conductivity makes possible the design of new time dependent experiments based on temperature measurements rather than on measured heat flow. The time needed for a liquid to adjust to new temperature boundary conditions is a relaxation time de- pendent directly on the liquid's thermal conductivity. Experiments which measure these relaxation times can be done with greater accuracy than those which depend on calorimetry and heat flow. In order to determine the thermal conductivity of a pure fluid, one needs a completely general 79 description of temperature variation as a function of space, time, and thermal conductivity in a cell. Then, after determining the temperature or its gradient at specific positions and times, one can fit the general temperature solution to the measured temperatures by varying the value of the thermal conductivity. If one's solution is correct and his experiments accurate, the value which gives the best fit will be the best experi- mental estimate of the true thermal conductivity. Horne and Anderson (1970) have recently developed such a temperature solution for binary mixtures of fluids in a pure thermal diffusion cell. Using a perturbation scheme, they solved the transport equations for both the heat conduction and the mass fraction as functions of position and time under the initial and boundary condi- tions of the pure thermal diffusion cell. Their solutions are superior to those of others in that Horne and Anderson account for convective heat and mass transfer, the depen- dence of density and other properties upon composition and temperature, and warming-up effects. Their perturbation scheme is readily applicable to many problems, and we use it here to solve the transport equation for heat conduction in a pure fluid. The transport equations for a pure fluid are actually simpler than those for mixtures, but the solutions of interest in such a case must be known at very short times, during which the temperature gradient 80 is being established in the cell, whereas the equations of Horne and Anderson are primarily intended to describe dif- fusion effects after this warming-up period is over. The classical thermal diffusion cell is a rectangu- lar parallelepiped bounded above and below by metal plates in contact with heat reservoirs which can be maintained at any temperatures. The area of the plates is made much greater than the distance between them in order to mini- mize edge effects. The cell often has glass walls so that refractive index changes resulting from temperature changes in the fluid can be measured. After the plates, reservoirs, and fluid in the cell have equilibrated at some temperature, a positive temperature gradient is applied to the cell by simultaneous heating of the upper plate and cooling of the lower plate through their respective reservoirs. This temperature gradient produces a negative density gradient without convection (for the normal fluid, whose thermal expansivity is positive). The establishment of the den- sity gradient occurs within the time necessary to reach the temperature steady state in the cell. This is a much shorter time than that within which a density steady- state is established in a pure thermal diffusion experi- ment. As in the work of Horne and Anderson, we use empirical parameters to characterize the times needed for the plate temperatures to reach their steady values. 81 Horne and Anderson assumed in their solutions that the upper and lower plate parameters are equal, while our solutions include a different parameter for each plate. In Section 2 we present the complete set of prac- tical macrosc0pic transport equations for a pure fluid and the conditions which the solutions of these equations must satisfy in a typical cell. In Section 3 we introduce the approximations which must necessarily be made in order to solve the equations of Section 2. The solutions of the temperature equations are presented and analyzed in Sec- tion 4. 2. Transport Equations Our fundamental transport equations correspond to equations (II.5.1), (II.5.5), and (II.8.22) with (II.5.23). These equations are greatly simplified when the fluid in question is assumed to be a one component liquid subject to only a gravitational external force. The above equa- tions under these conditions are also those of TIP (Horne and Bearman, 1967). We assume that the temperature gra- dient is applied only in the vertical direction with no horizontal effects and that the cell walls are adiabatic so that there is no horizontal heat flux through the walls. Heat and matter flow only in the vertical direction and there is no steady state convection. 82 The hydrodynamic equation of continuity of mass is (dp/dt) + p(8V/8z) = 0, (2.1) where p is the density, v is the barycentric velocity, t the time, and z, the vertical position coordinate, and where the substantial time derivative, d/dt, is related to the local time derivative by (d/dt) = (8/8t) + v(8/8z) = 0. (2.2) For the equation of energy transport we have pEé(dT/dt) - TB(dP/dt) = ¢1 - (8q/8z), (2.3) where E? is the specific heat capacity at constant pres- sure, T is the temperature, 8 is the thermal expansivity, ¢l is the entrOpy source term for bulk flow, and q is the heat flux. The linear phenomenological equation for the heat flux is -q = K(3T/32). (2.4) where K is the thermal conductivity of the fluid. The initial conditions in the cell are V(z,0) = 0, T(z,0) = TM, (2.5) and the boundary conditions at all times are 83 V(a/2It) 0 = V(-a/21t) T(a/2,t) ¢U(t) T(-a/2,t) = ¢L(t), (2.6) where g.is the cell height. Warming-up effects are taken into account by the time dependent functions for plate temperatures, 8 and °L' U One can empirically determine the time required for a plate to achieve its steady state. It will generally depend on the plate material and thickness, on the reservoir heat capacity, and on the means of supplying heat to and re- moving it from the reservoirs. A typical functional form for the plate temperatures is ¢U(t) = TM + (TU-TM)[1 - exp(-t/YU)] ¢L(t) = TM + (TL- TM) [1 - exp(-t/‘YL)]) (2.7) where TU and TL are the temperatures applied to the upper and lower plates, respectively,and YU and YL are the characteristic relaxation times of the upper and lower plates. These times are best obtained experimentally. Typically (Anderson, 1968), YU and YL are of the order of 40 to 60 seconds. 3. Approximation Methods. In order to solve the transport equations given above we must make approximations on three levels: (i) the equations are simplified by suppressing demonstrably 84 negligible terms; (ii) a self-consistent, well-ordered perturbation scheme is introduced to take into account variable coefficients; and (iii) a Fourier sine or cosine solution is obtained for each perturbation equation. It is essential to recognize that at no point do we assume that coefficients are constant or that the barycentric velocity is zero. Simplification of Equations.--We can neglect terms in the transport equations (2.1), (2.3), and (2.4) if they are very small compared with other terms. Ob- viously equations (2.1) and (2.4) cannot be simplified; but in equation (2.3) we examine the pressure and entropy source terms. The magnitude of these-terms in a pure fluid is comparable to their magnitude in a mixture, and the magnitude of the heat flux in a pure fluid is compa- rable to its magnitude in a mixture, when both the pure fluid and mixture are subjected to the same experimental conditions. For this reason, the arguments of Horne and Anderson for the suppression of these two terms are valid for pure fluids as well as for mixtures. The time depen- dence of pressure is neglected because we assume that the steady-state pressure distribution in the cell is attained instantaneously. Horne and Anderson found that ¢1 is at most lo‘lsJ'm-Bs-l, whereas aq/Bz is on the order of 25J -38-1 m , so ¢1 in (2.3) is certainly negligible in com- parison to the heat flux gradient. 85 These approximations simplify our energy trans- port equation (2.3) so that with (2.4) we have pEé(8T/8t) = - pCév(8T/8z)4-8[K(8T/8z)]/8z. (3.1) By equation (2.2) we have for (2.1) (8v/8z) =.- (82np/8t) - v(8£np/8z), (3.2) where density derivatives are simply related to tempera- ture derivatives by the chain rule, dznp =-BdT, (3.3) when pressure derivatives are neglected as assumed above. Perturbation Scheme.--The perturbation scheme of Horne and Anderson is used to take the non-constancy of coefficients in the transport equations into account. It is certain that the coefficients vary with temperature (and pressure) in the cell, but it is also true that fairly reasonable solutions to the transport equations have been obtained using constant coefficients. Following this reasoning, we treat every coefficient as a constant plus perturbation terms which depend on temperature fluctua- tions. This scheme modifies the usual solutions of the energy equation and also allows the density of the system to vary, thus permitting the barycentric velocity to change. 86 The most general expression for the representation of a coefficient which varies with temperature (or other state variables) about some value for a fixed set of state variables is the Taylor series; therefore, we write each coefficient, L, as a Taylor series with successive higher order derivatives being preceded by higher orders of the ordering parameter, 6: 1’3 = Lo + em - TM)LT+ 52(1/2)(T - TM)2LTT + NJ), (3.4) where _ _ _ 2 2 Lo - L(TM) LT - (8L/8T)TM LTT — (8 L/8T )TM. (3.5) When a = l, L = L and we have the exact Taylor series ex- pression for L (neglecting pressure dependence in accor- dance with our previous assumption). The increasing orders of 8 indicate the decreasing contribution of those terms to L. Accordingly we write our solutions as per- turbations in s also so that A _ 2 3 T — TM + To + 5T1 + a T2 + O(e ), 9 = v + evl + 52v + C(63), (3.6) 0 2 where T = T and G = v if e = 1. All of the solutions Tn and vn are functions of space and time. The total solu- tions are actually Taylor series in a about a a 0. 87 The parameter s is a dummy device used to order subsequent manipulation of equations in accordance with the tempera- ture derivatives of the coefficients. Thus, the zeroth- order problem and solution involve no temperature deriva- tives, the first-order problem involves first derivatives, the second-order problem and solution involve second de- rivatives and products of first derivatives, and so forth. As in any perturbation method, one cannot be certain a priori whether the solutions (3.6) converge when a a l. A reasonable procedure is to find the solutions through the second-order in e and to compare the zeroth-, first—, and second-order solutions. If the importance of the re- sults diminishes relatively rapidly with increasing order, then one may stop. If, on the other hand, some higher- order result appears to be more important than a corre— sponding zeroth-order result, one must adjust his per- turbation scheme so that the offender is included at zeroth order. The scheme proposed in equations (3.4)- (3.6) works well for thermal conduction in a pure fluid. The first application of equation (3.6) is to equation (3.4), which becomes " _ 2 2 3 L _ Lo + eToLT-l-e [(1/2)TOLTT+T1LT] + 0(6 ). (3.7) wherein all of the subscripted L's are constants as de- fined in (3.5) and all of the T's vary Spatially and temporally. 88 The second application of the perturbation scheme is to the velocity equation (3.2). The initial and bound- ary conditions at each order of e are, by equations (2.5) and (2.6), vn(z,0) = 0 vn(ia/2,t) = 0, n = 0,1,.... (3.8) Putting L = lnp in (3.3) and (3.7) we have for the zeroth- order velocity 8v 0 (3.9) .._..._.=O' 8t which by (3.9) implies V0(z,t) = 0. (3.10) By equations (3.2), (3.3), (3.7), and (3.10), the velocity equations for higher powers of s’are all of the type (8vn/82) = 80(8Tn_1/8t) + Vh(z,t), n = 1,2,... (3.11) where for the first and second powers of e Vl(z,t) = 0, V2(z,t) = TOBT(8To/8t) + v180(8T0/az). (3.12) The result of our perturbation scheme is thus a differ- ential equation for each order whose inhomogeneous part contains only terms of lower order. 89 Substitution of equations (3.6), (3.7), and (3.10) into equation (3.1) and separation of the results accord- ing to the power n of 6 yields the following temperature equations: 1 2 2 (8 Tn/az ) = Un(z,t) n = 0,1,... (8Tn/8t) - KO(DCP)O (3.13) with U0(z,t) = 0 _ - -l Ul(z,t) — (pCp)o {8[TOKT(8TO/8z)]/8z} — - -1 - (pCp)T(pCp)o TO(8TO/8t) - v1(8To/82). (3.14) The similar but very long expression for U2(z,t) is omitted for brevity. The initial and boundary conditions for each power n of e are, by equations (2.5), (2.6), and (3.6), Tn(2,0) = 0 n = 0,1,..., Tn(ia/2’t) = o n = 1,2,..., T0(a/2.t) = ¢U(t) - TM. T0(-a/2,t) = ¢L(t) - TM. (3.15) Fourier Transforms.--Inspection of equations (3.11) --(3.15) reveals that the equations must be solved in the progression To, v1, T1, v2, T2, etc., and that the tem- perature equations are all of one type. For this reason the method of solution will be the same, and we use the method of Fourier transforms for the temperature solutions 90 as did Horne and Anderson. Whether we choose a sine or cosine transform solution is dictated by our boundary con- ditions, 122;! whether we know the solution or its Spatial derivative as a function of time at the boundary. In the first instance, we use a sine transform; in the latter, a cosine. Now, for our temperature equations, we know the behaviour with time of the temperature at the walls rather than the behaviour of its derivative, so we can write for every order, T , n Tn = Z gnm(t)sin(mwx/a), n = 0,1,..., (3.16) m=l where x = z + a/2 (3.17) and where a g (t) = (2/a) ] T (x,t)sin(mnx/a)dx, m = 1,2,... nm 0 n n_= 0,1,2,... (3.18) We multiply each term of equation (3.13) by (2/a)sin(mwx/a) and integrate each of them over x from 0 to a. This pro- cedure gives us a (twat - ( '6 )'1(2/ > [a (32 T (tn/3x2} 1 (mvrx/a)dx gnm Ko 0 p o a 0 1'n s n a = (2/a)(pE ) 1 f u (x,t)sin(mnx/a)dx, n = 0,1,..., p o 0 n I“= 1,2,... 0 (3.19) 91 The second term is readily integrated by parts using the boundary conditions (3.15) to yield 2 dgnm(t)/dt + (m /T)gnm(t) = ¢Om(t)60n + (2/a) [a Un(x,t)sin(mwx/a)dx, n = 0,1,..., 0 m = 1,2,..., (3.20) where 50n is a Kronecker delta, where - - - m _ - - ¢0m1t1 - (2m/1rT){( 1) [¢U(t) TM] [¢L(t) TMJ}. (3.21) and where T is a thermal relaxation time, _ 2 — 2 T — a (pCp)o/w KO. (3.22) The differential equation for each gnm(t) is easily solved, in principle, by means of the integrating factor, exp(m2t/T). The initial conditions are readily found by inserting the intial conditions (3.15) in equation (3.18). Thus we have 0,1,..., 3 ll gnm(o) = OI n 1,2,... 0 (3.23) 4. Practical Results We can now find our temperature solutions to any desired degree of accuracy by solving equation (3.2) re- peatedly for successively higher values of n. We solve the first two orders rigorously and estimate the 92 contribution of the solution which is second-order in e from its steady state solution. By equations (3.16)--(3.23) we find the solution for zero-order temperature. T0 = (z/a +‘l/2)(TU r TM)[1 — exp(-t/YU)] + (z/a + 1/2)(TM - TL)[l - exp(-t/yL)] + 21 (-l)n(nn)-1{[4n2(YU/T)'1]-1(TU-TM1 n: x [exp(-t/YU) - exP(-4n2t/T)] + [4n2(YL/T)—1]'1(TM- TL) [exP(--t/YL) - exp(-4n2t/T) 1} x sin(2nnz/a) so + g (2/n)(-1)1(22+1)‘1 [(22+1)2(yU/T)-1]-1(TU - T M) x {epr-(2(+1)2t/rl«-eXP(-t/YU>} - [(22+1)2(YL/r)-11‘1(TM - TL) x {exp[-(2£+l)2t/T] - exp(-t/YL)})008[2£+1)fl22/a]. (4.1) If the initial temperature, TM, is the average of the applied temperatures, T and T U L' the form of this equa- tion is reduced to T0 = AT(z/a)[2 - exp(-t/YU) - exp(-t/YL)] + (AT/4)[eXP(-t/YL) - exp(-t/YU)] + (AT/2n) 2 (-1)nn‘1{[4n2(Y_/T)—11‘1 n=l U’1 93 X [exp(-t/YU) - exp(-4n2t/T)] + [4n2(YL/T)-1]-1[exp(-t/YL) - exp(-4n2t/T)]} X sin(2nnz/a) as + (AT/n) 2 (-1)1(21+1)‘1 [(21+1)2(yU/T)-11' (=0 1 X {epr-(22+l)zt/Tl-exp(-t/YU)} - [(22+1)2(YL/T)-l]{exp[-(2£+l)2t/1]-exp(-t/YL)} X cos[(2£+l)wz/a] (4.2) where AT = TU - TL. (4.3) If the plate relaxation times are the same, but the initial temperature is not the same as the average of the applied temperatures, we have the reduced form To = (Z/a)(TU - TL)[1 - exp (-t/Y)] + l/2(TU + TL - 21M)[1 - exp(-t/Y)] (-1)“(1rn)"1('r 1 - TL)[4n2(v/r)-11‘ 1 + U IIM8 n X [exp(-t/Y) - exp(-4n2t/t)]sin(2nnz/a) co 2 (2/11)(—1)1(22+1)('rU + T + 2 —l L - ZTM)[(2£+1) (Y/T)-1] X {expl-(2£+l)2t/r] - exp(-t/Y)}cos[(2£+l)nz/a], (4.4) 94 where Y = YU = YL. (4.5) Finally, if both plates have the same relaxation time and the initial temperature is the average of the applied tem- peratures, we have the greatly simplified form T = AT{(z/a)[l - exp(-t/Y)] + 2 (-1)n(n/n)[4n2(y/r)—1l'1 1 0 n: X [exp(-t/Y) - exp(-4n2t/T)]sin(2nnz/a) . (4.6) The complex forms of equations (4.1), (4.2), (4.4), and (4.6) might discourage one from attempting numerical evaluation for specific experiments. The series converge rapidly for times of ten seconds or more after the applica- tion of the temperature gradient and could be evaluated on a calculator, but evaluation by a digital computer is faster and can be carried further with a greater degree of accuracy. At short times the series term of (4.6) is nearly cancelled by the leading term, thus it needs to be known to a very high degree of accuracy. All evaluation of such equations has been done on an CDC 6500 computer. The series have been calculated until changing their length by ten percent fails to change their value by approximately 0.1% or more (212;! 100 terms of a series are calculated and summed and then flhe next ten terms are calculated and summed; if these ten (xenms do not change the initial sum of 100 terms by more "I a—u- ‘u 95 than 0.1% of its original value, the series is truncated after these 110 terms and all further terms are assumed negligible). In higher order solutions where multiple series appear, this criterion is applied to each of them. A plot of T0 versus cell position at various times as calculated from equation (4.6) is shown in Figure l. The approximate parameters for carbon tetrachloride shown in Table l have been used. The solution is antisymmetric about 2 = 0. Table l--Approximate physical properties of CC14 j. e=1.2 x 10‘3K'1 I< =1.1 x lO-J'J' m'1s‘1x‘1 K-1(3K/3T)=- 1.2 x 10'3K’1 p'c‘ = 1.4 x 106.1 m‘3K‘1 p (pEé)-l[8(pE§)/8T] = 7.1 x 10‘1K"1 In every case, our temperature equation depends only on time, position, length of the cell, temperatures of the plates, relaxation times of the plates, heat ca- pacity of the fluid at its initial temperature, and the thermal conductivity of the fluid at its initial tempera- ture. The experimentalist can determine all of these factors prior to and during his experiment, with the 96 Figure l--Plot of T0 versus (z/a) for CCl4; a = 303, b = 605, c = 90s, d = 1205, e = 1508. 97 GdCb 0.4 )- —0.4 - - 0.2 0.0 0.2 0.4 — 0.4 2/8 Figure l 98 exceptions of the heat capacity and of the thermal con- ductivity. If, however, the heat capacity of the fluid is known from other experiments, the thermal conductivity can be determined by monitoring the temperature of the fluid (or some parameter which is simply related to tem- perature) and fitting it to whichever of equations (4.1), (4.2), (4.6), or (4.4) is applicable by variation of the thermalifionductivity. Such an application of these equa- tions has been made by Olson (1972). In order to make such measurements, however, we have assumed that only To in the perturbation solution T - TM = TO + T1 + T2 +... is important. To test this hypothesis the solution for T1 is found and evaluated for typical parameters in order to determine its magnitude. We have a complete solution T1 for y =4YU =4YL and T + U T = 2T Minor variations from these conditions do not L M' affect the magnitude of T1. The Fourier solution of (3.20) with n = l and subject to the above conditions is very com- plex, involving triple sums of infinite series. It is given in full detail in Appendix C. This solution has been evaluated for typical values for carbon tetrachloride. The results are shown in Figure 2. It is symmetric about 2 = 0; therefore, its derivative is zero at z = 0 for all t and will add nothing to the temperature derivative moni- tored at z = 0. T1 is zero initially, rises to a maximum of 5 x lO'SK at a time of approximately one-half T, and 99 Figure 2--Plot of T1 versus (z/a) for CCl4; a = 303, b = 605, c = 90s, d = 1203, e = 1503. 100 4.0 - 3: $2 x c \, J 0.0 0.2 0.4 2 /a —— 0.2 —— 0.4 Figure 2 101 then falls steadily, going negative until the steady-state solution, T1‘”’ = (1/2)(AT)1(KT/Ko)[(1/4)-(z/a)21. (4.7) is achieved. The steady-state value in carbon tetrachloride at z = 0 is approximately - 1.6 x 10-4K. The behaviour of T1 at t < 21 is shown in Figure 2 and may be compared to the values of To shown in Figure 1. We can also investigate the behaviour of T It 2. too will have a maximum absolute value in the steady state where 12(w) = (1/2)(AT)3[(KT/Ko)2 - (KTT/3K011 x ((z/a)3 - (z/4a)1. (4.8) This value has been shown to be small everywhere by Horne and Anderson, and it is zero at z = 0 while its gradient is on the order of lO-GK. Comparison of the values of T0' T1, and T2 shows that, indeed, the solution To gives a very good approxi- 1mation to the total temperature solution particularly at short times, and thus, that the thermal conductivity as fitted by equation (4.1), (4.2), (4.4), or (4.6) should be a very good approximant to the true thermal conduc- tivity. The estimates above indicate that the theoreti- cal error caused by truncation of the temperature solution after To is probably less than experimental error. CHAPTER IV THE. DUFOUR EFFECT 1. Introduction When a concentration gradient is imposed on an originally isothermal fluid mixture, the phenomenological equation for the heat flux predicts that a temperature gradient will develop as diffusion occurs. This phenome- non is called the Dufour effect, after its discoverer (1873), or the diffusion thermoeffect, to indicate that it is the reverse of thermal diffusion. The effect is of interest in liquids for three important reasons: (1) be- cause it can be used to verify the heat-matter Onsager reciprocal relations, (2) because the temperature varia- tions could cause complications in diffusion experiments, and (3) because it has never been unambiguously observed. The purpose of this work is to predict theoretically the magnitude of the Dufour effect when liquid mixtures are allowed to diffuse into one another and to determine where the Dufour effect can be best measured in a geo- metrically well-defined cell. Interest in the Dufour effect rose early in World War II when Clusius and Waldmann (1942) and Waldmann 102 103 (1939, 1943, 1946, 1947a, 1947b, 1949), who were inter- ested in determining thermal diffusion factors for gases, developed kinetic and phenomenological theories and ex- periments to measure the Dufour effect in gases. The equations of the phenomenological theory were solved for adiabatic diffusion and diffusion between cylindrical chambers with diathermic walls. Waldmann's final results depend heavily on the ideal gas law, and they therefore cannot be extended to liquids. The work also requires constancy of thermal conductivity and the heat of trans- port within each chamber of the experimental apparatus, and the solutions are based on the assumption that the temperature at the boundary of the two chambers does not change during the courSe of the experiment. Miller (1949), noting the great fluctuations of the temperature at this boundary, suggested that they might be due to heat of mixing as well as to variations in density and Specific heat. He did not, however, develop any theory to take these into account. Further kinetic theory has been develOped by Grew and Ibbs (1952), and the phenomeno- logical theory is presented by Meixner (1943) who assumed that the reciprocal relations hold. Waldmann (1947a) used his phenomenological theory to estimate, for the liquid mixture hexane-octane, that the variation of temperature in one chamber would be a maximum of 0.04 K. He neglected heat of mixing. The only 104 investigations of the Dufour effect in liquids have been done by Rastogi and Madan (1965) and Rastogi and Yadava (1969, 1970). Their work is primarily experimental and their results are based on ill-founded mathematical as- sumptions about the behaviour of their system. We shall discuss their work in more detail at the end of this chap- ter. Thus, for many years the Dufour effect in liquids was ignored because it was felt to be too small to measure. Its presence in liquids is masked by the heat of mixing which is not present (at least to any great extent) in gases. Any measurement of the Dufour effect is complicated by the fact that no steady-state other than the equilibrium state is reached by a system which initially has a con- centration gradient, and thus the measurement of the tem- perature gradient arising from the Dufour effect must be time dependent (Fitts, 1962). The work reported here is a theoretical examina- tion of the temperature effects which result when two binary fluids of different composition are allowed to mix. These are the effects predicted by the macroscopic trans- port equations, which include in the temperature equation terms for the transported enthalpy. These are the terms that take into account the heat capacities of the two fluids and their heat of mixing. Terms which are demon- strably small are omitted from the transport equations, 105 and transport coefficients are allowed to vary with the thermodynamic variables of the system by expressing them in Taylor series in composition and temperature. Two ideal experiments which could be approximated in the laboratory are defined. The results of any real experiment should extrapolate to the results given here. We determine optimal conditions for (1) measurement of and (2) neglect of the Dufour effect. We define our first system to be an adiabatic cylinder or rectangular tube of length a. The value of a will be chosen to get maximum temperature variations as predicted by the solutions of the transport equations with consideration of experimental practicality. The length would probably never exceed ten centimeters. The cell is divided into an upper and a lower chamber, each of which contains a mixture of the same two liquids but in different proportions. In the extreme case each half could contain a pure liquid. The more dense liquid is placed in the lower chamber and the boundary between the two mixtures is assumed to be infinitely sharp initially. The two mixtures, originally at the same temperature, are allowed to diffuse into one another. Diffusion occurs only in the vertical direction and causes vertical tem- perature variations in the cell. At infinite time the composition in the cell will be uniform (except for sedi- mentation effects) and the temperature will again be 106 uniform throughout the cell, though not necessarily at the initial temperature. The second system is identical to the first except that the ends of the cylinder are assumed to be diathermic. The lateral walls of the cylinder are still adiabatic, or the cylinder is of sufficient diameter that variations in edge temperature will be negligible,and diffusion will hence occur only in the vertical direction. 2. Transport quations The equation of energy transport in a binary liquid is Lg£.(II.8.22) with (11.5.6)] pCé(dT/dt) - TB(dp/dt) = ¢1 - (8q/8z) - j1[a(fil-fiz)/321. (2.1) where C? is the Specific heat capacity, p the density, T the temperature, 8 the thermal expansivity, p the pressure, 01 the entropy source term for bulk flow, q the heat flux, j1 the diffusion flux of component 1 relative to the bary- centric velocity, and ii the partial specific enthalpy of component i. The time variable is t, and z iS the vertical position coordinate (the initial boundary is at z = 0, the ends of the cell at i a/2). The conservation of mass equations for the fluid are [g£. (II.5.l) and (11.5.2)] (dp/dt) = - p(8v/8z) (2.2) 107 and p(8w1/8t) + (8j1/8z) + pv(8wl/8z) = 0, (2.3) where v is the barycentric velocity and W1 is the mass fraction of component 1. The linear phenomenological equations for the fluxes of heat and mass are (Horne and Bearman, 1967) -q = K(3T/32) + 0DQi(8wl/Bz) + DDQiSIW1W2(8p/8z) (2.4) and = pD(8wl/8z) - pDa w w T-1(8T/8z) ‘31 112 + pDSlw1w2(8p/8z), (2.5) where K is the thermal conductivity of the mixture when there is no concentration gradient (at infinte time), D is the mutual diffusion coefficient, Qi is the heat of transport of component 1, al is the thermal diffusion factor of component.1,and S1 is the sedimentation coef- ficient of component 1. The solutions to the transport equations for con- centration and temperature must fit the initial and boundary conditions of the systems described. Initially we have w1(z<0,0) = w1A + wlA ' NEH NIH ego 15% v(z,0) = 0, wl(z>0,0) , T(z,0) = T , w1(Q,0) = w (2.6) lA' 108 where w1 is the mean of the mass fractions in the two A chambers of the cell and 1 (1/2)w$ are ‘thé‘initial diaplace— ments of wl fromw1A in each half of the cell. At the ends of the cell we have no matter flux and either no heat flux in the adiabatic case or constant temperature in the diathermic case, so q(ia/2,t) = 0, (adiabatic) v(ia/2,t) = 0, j1(ia/2,t) = 0, T(ia/2,t) = T . (diathermic)M (2.7) In order to solve the problems which have been posed, we assume that the equations above are correct and can be applied to the idealized experiments described above. We must make further assumptions about terms which can be shown to be essentially constant in order to solve these equations. By using the same arguments as Horne and Anderson (1970) as our first approximation, we neglect pressure terms where now we consider I8T/8zl 5 1 deg cm-1 and |8wl/82| 5 10 cm-l. The size of the two derivatives will diminish simultaneously. Secondly, we neglect the thermal diffusion term in equation (2.5) because [alwlwz/TI 3 l, and magnitude of the thermal is on the order of 10- deg- diffusion term is therefore at most 0.01% of the diffusion term. Finally, we neglect the entropy source term of equation (2.1) since (Horne and Anderson, 1970) it is proportional to the square of the velocity gradient, 109 (8v/az)2, which we assume to be unmeasurable in a dif- fusion experiment. Subsequent results verify this assump- tion '. The equations remaining are then: -q = K(8T/8Z) + pDQi(8w1/8z), (2.8) pCp(8T/8t) = pD[8 (Fl-8'2) /8z] (8wl/8z)+8 [pDQi(8wl/8z) ]/8z - pCév(8T/az) + 8[K(8T/8z)]/8z, (2.9) p(8w1/8t) = 8[pD(8w1/8z)]/8z - pv(8w1/8z), (2.10) (8v/8z) = -(82np/8t) - v(8£np/8z). (2.11) 3. Perturbation Scheme The coefficients in equations (2.9)--(2.ll) are functions of wl, T, and p. In order to solve these equa- tions we must devise a perturbation scheme which allows us to take into account the nonconstancy of the coef- ficients in Space and time. For this purpose we use the perturbation scheme of Horne and Anderson, introduced in Chapter III of this thesis. Here, we expand the coef- ficients in w1 and T about w1A and TM. Variations with pressure are neglected in accordance with our previous assumption, and successive derivatives with respect to w1 and T are multiplied by higher powers of the ordering parameter, 8. Thus, we have for any coefficient, L, 110 L = Lo + e[(T-TM)LT + (wl-wlA)Lw] + ez[(1/2)(T-TM)2LTT+(T-TM)(wl-wlA)LwT+(l/2)(wl-wlA)2wa] + C(63), (3.1) where L = L(T ,w ) L = (8L/8w) L = (8L/8T) o M 1A w l TM,w1A T TM,w1A as) L = (82L/8T2)T 2 L = (8 L/8w 2) l TM'wlA wT L = (82L/8w WW 1 M’wlA (3.2) The first and third terms on the right hand side of equation (2.9) arise because we are studying mixtures. Without these terms equation (2.9) would simply be the thermal conductivity equation for a pure fluid. Because we do not expect thermal conductivity in a mixture to dif- fer very much from that in a pure fluid, we can treat the two terms which arise in the thermal conductivity of mix- tures as perturbations on the pure thermal conductivity equation. We recognize these perturbations by multiply- ing each of the two terms by a second perturbation pa- rameter, 6. Thus, we have pCé(8T/8t) = 0pD[8(fii-fié)/8z] (8w1/8z) - pC§v(8T/82) + 8[K(8T/8z)]/8z + 68[pDQi(8wl/8z)]/8z. (3.3) Our solutions for T, wl, and v are the perturba— tion solutions 111 2 A 2 T=T+T+T+ T o 6 6 T + 6Te + 86 86 + 56 T8 M 6 66 66 + 0(53) + 0(52) A - _ 2 2 W1 - w1A - w + 60)6 + 6 (066 + ewe + £60086 + 86 w o 866 + 0(53) + 0(22) v=v +6v +62v +ev +e6v +e6zv o 6 66 6 86 £66 + 0(53) + C(82) , (3.4) Each term in equations (3.4) (except w1 and TM) is a A function of space and time. When 8 = 1 and 6 = l, T = T, 81 = wl, and G = v. The second perturbation parameter 6 has the effect of making all mass fraction derivatives appear in lower order equations than their counterpart temperature deriva- tives. We select this scheme because coefficient varia- tions with temperature are in general considerably smaller than variations with mass fraction and because we are dealing with small temperature variations and large mass fraction variations. In Chapter III we investigated the behaviour of zeroth-, first-, and second-order solutions in e and found for the case of thermal conductivity that the magnitude of these solutions decreased rapidly with increasing order. We assume that the behaviour here will be similar and terms of order 82 should be very small in- deed. Terms of order a Should be smaller than terms of zeroth-order in 6, although inversion of these 112 contributions will occur in a system with very large heat of mixing and small heat of transport. We have not pre- viously investigated the behaviour of terms which are zeroth-, first-, and second-order in 6 so we investigate 2 our temperature solutions to order 6 . We first apply (3.4) to (3.2) to obtain A: + L Lo €(TOLT + woLw) + €6(Dw(05 + DTT6) 2 3 2 + 56 (Dwu)66 + DTT66) + O(6 ) + O(e ). (3.5) By inserting equations (3.4) into equations (2.9), (2.11), and (3.3) and by using equation (3.5), we obtain a set of differential equations which can readily be solved for each order of e and 6. Very good approximations to T, wl, and v can be found by solving for the first few terms of equations (3.4), provided that the series is convergent. We are, of course, primarily interested in the equation for T, but it is necessary to have lower order terms in w1 and v in order to evaluate terms in T. 4. Zeroth-Order Solutions for an Adiabatic System Before stating general equations for all orders of e and 6, we solve our zeroth-order equations for v0, To' and mo. These results can then be used to simplify the general equations. The initial and boundary conditions for the bary- centric velocity are 113 vn(z,0) = 0, vh(ia/2,t) = 0, n = 0,6,€,... . (4.1) The zeroth-order equation for velocity is (8vo/8z) = 0, (4.2) so from equations (4.1) and (4.2) we find vo(z,t) = 0. (4.3) The zeroth-order mass fraction equation has the form (80) /8t) - 0 (320) /8z2) = 0. (4.4) O O O The solution of this equation can be obtained through Fourier transform methods as discussed in Chapter III. We first change our space variable to x such that x =12 + (a/2), (4.5) Our boundary conditions dictate that a cosine transform be used because we know the gradient of the mass fraction at the boundaries from equations (2.7) and (2.8) (after neglecting the last two terms): I o s (8wl/8x)o't = (8wl/8x)a't - (4.6) so that II 0 s n = o,6,e,... . (4.7) (8wn/8x)o,t (800n/8x)a’t 114 We assume wo has the form 00 wo(x,t) = (l/2)foo(t) + m£1 fom(t)cos(mnx/a), (4.8) where a fom(t) = (2/a) é wo(x,t)cos(mnx/a)dx, m = 0,1,2,... (4.9) The Fourier transform of equation (4.4) using (4.8) and (4.9) is [8fom(t)/8t] + (m2/0)fom(t) = o, (4.10) where 0 = az/(NZDO). » (4.11) This equation (or any similar inhomogeneous equation) can readily be solved using an integrating factor, exp(mzt/O), and we find - _ 2 fom(t) - A exp( m t/O) (4.12) where A is a constant not yet determined. Our initial conditions on fom are found by requiring that wo(x,0) = w1(x,0) - wlA' (4.13) and leaving (4.14) II 0 s 5 II 0‘) ‘ m s I wn(x,0) 115 Applying the initial conditions on mass fraction, we find 0 m is even f0111(0) = (m-l)/2 2w§(nm)’1(-1) m is odd. (4.15) Thus, we have 1 wo(x,t) = Zwifl-l 2 (-l)£(2£+l)_ exp[-(22+1)2t/0]cos[(22+l)nx/a]. 2 0 (4.16) as our zeroth-order solution for composition. Our zeroth-order equation for temperature has the same form, 1 — - 2 2 _ (8To/8t) - (<0(pCp)O (8 TO/8z ) - 0. (4.17) A solution is again found using a finite Fourier cosine transform: To = (l/2)goo(t) + mil gom(t)cos(mnx/a) (4.18) because equations (2.7) and (2.4) (consistently consider- ing the heat flux in a mixture as a perturbation of the heat flux in a pure fluid) we have [K(8T/8z) = - 6pDQi(8wl/8z)]z=ia/2, (4.19) so that in the zeroth-order case we have from (4.7) (8To/8x) = (8To/8x)a't = 0. (4.20) 0,t 116 The time contribution to the solution is gom(t) = B exp(-mZt/T), (4.21) where r = (pE§)oa2/(Kon2). (4.22) However, we have B = 0 because T(x,0) = TM (4.23) from (2.6), so Tn(x,0) = 0, n = o,6,e,... . (4.24) Thus, To(x,t) = 0. (4.25) 5. Higher Order Solutions for an Adiabatic System We use equations (4.3) and (4.25) to simplify the differential equations for higher order terms. The ve- locity equations have the form (8vn/8z) = Vn(z,t), n = 5,8 (5.1) where the Vn(z,t) are explicit functions determined from previous solutions: V6(z't) 0, V = 0, 66 V€(Z)t) - (inp)w(8wo/8t)r 117 V€6(z,t) = - (£np)w(8w6/8t) - (inp)T(8T6/8t) - v5(£np)w(8wo/8z), V56612’t) = - (£np)w(8w56/8t) - (£np)T(8T66/8t) - v5(£np)w(8w6/8z) - v6(2np)T(8T6/8z). (5.2) We immediately see from (5.2) and (4.1) that v6 = 0, = 0. (5.3) V55 The equations for mass fraction have the form (8w /8t) - D (326 /8z2) = w (z t) n = 5 e (5 4) n o n n ’ ' "'°’ ' where W6(z,t) = - V5(8wo/82), . W65(z,t) = - v55(8wo/8z) - v6(8w6/8z) W€(z,t) = [Do(1np)w + Dw](8wo/8z)2 + wao(82wo/8zz) - v€(8wo/8z), W86(z't) = [2Do(£np)w t 2Dw](8wo/8z)(8w6/8z) + [Do(£np)T + DT](8T6/8z)(8wo/8z) + wao(82w5/8z2) + wa6(82wo/8zz) + DTT6(82wo/8zz) — v€6(8wo/8z) - v6(8w€/8z) - v€(8w5/8z), W566 = [(2np)TDO+DT](8T66/8z)(8wo/8z) + 2[(2,np)wDo + Dw](8w66/8z)(8wo/8z) + [(2np)TDO+DT](8T5/8z)(8w5/8z) 118 + [(2np)wDo+Dw](8w6/8z)2 - v€56(8wo/8z). - v€(8w66/8z) - v€5(8w6/8z) (5.5) The temperature equations are very similar with Un(z,t), n = 6,6,..., - 2 2 (pCp)o(8Tn/8t) - Ko(8 Tn/8z ) (5.6) where U5(z,t) = (pDQi)o(82wo/8z2), U55(z,t) (05§)ov5(8T5/32) + (0DQi)o(3w5/32). U€(z,t) = 0 085(z.t) = - (pap)wwo(aT5/at)+((p0)o(fil-fiz)w + (puczgpwl(awe/32)2 - (pC§)ove(8T5/8z) (pCé)ov6(8Te/8z) + Kw(8wo/8z)(8T6/8z) + wao(82T6/8zz) + (pDQi)wwo(82wo/8zz) + (pDQi)o(82w€/822L The similar but very long expression for 0566 is given in Appendix E. The initial conditions for equations (5.4) and (5.6) are given by equations (4.13) and (4.23), respec- tively. It is obvious that the equations must be solved in a certain order, although there is some flexibility. We choose to solve them in the order T5, v5, w6' T55, Te’ Ve’ we' Te6' v86' ”66' T666’ v666’ we66' V85' 856' The order in which the solutions are found is immaterial 119 unless an unknown function is required in the inhomogeneous part of some equation, in which case solution is impossible. Obvious solutions of the above equations include V5 =85 ”1'55 =V5o j6 o j6 2 2 = * (pDQl)o(8 wj/8z ) + Yj(z,t) j = 0,8 (5.14) where 0) Yo(z,t) Y€(z.t) (pE§)wwo(aT5/at) + [(0D)o(fii-fié)w+(0voi)wl(awe/az)2 (pCé)ov€(8T5/8z) + Kw(8wO/8z)(8T6/8z) + wao(82T6/8zz) + (pDQi)wwo(82wo/822). (5.15) The equation for T€56 is given in Appendix E along with its solution. The boundary conditions for each order are found from (4.18): [Ko(3Tj5/32) + (DD011018wj/3211ia/2,t = Cj(ia/2.t). j = 0.8 with I o Co(ia/2,t) c€(:a/2.t) 119D011ww018wo/az)1ia/2,t' (5.17) 122 The Fourier cosine transform of equation (5.14) is atg. (t)1/at + (mz/r)g. (t) = c (t)(p6 )‘1 36m 36m jnl p o 2 3 - '1 2 a + (Zn /a )(pDQ11019Cp10 m g chos(mflx/a)dx + (2/a)(pC )-1 [3 Y.(x,t)cos(mnx/a)dx, (5.18) p o 0 3 where = - m - - ij(t) (2/a)[( 1) Cj(a/2,t) Cj( a/2,t)]. (5.19) The temperature solution T6 is then given by 15 = (zwi/n)(p00;)o K;1(1-(r/0)1'1 x Z (-1)1+1<22+1)‘1{exp[-(21+1)2t/0]-exp[-(22+1)2t/r]} 2=o X coS[(22+l)nx/a]. (5.20) and the expression for T56 is given in Appendix C. 6. Solutions for a Cell with Diathermic Ends The temperature solutions for the cell whose ends are diathermic differ radically from the adiabatic solu- tions of Sections 4 and 5. However, the equations to be solved and the initial conditions are the same; the only difference is in the boundary conditions. This, fortu- nately, causes all those solutions which were zero in the adiabatic case to again be zero in the diathermic case; i’namely, equation (5.8) is true in the diathermic case also. 123 Our solutions are found in a manner similar to that described in Sections 4 and 5. The boundary condi- tions for temperature now are Tn(ia/2,t) = 0, n = 6,66,..., (6.1) and all of the temperature solutions have the form 00 Z 9nm(t)sin(mnx/a). n = 5,e5,... . (6.2) T (x,t) n m=l For order 6 we have T6 = <8wi/"2)(°DQE’OK81 g 2 (-1)1+1(21+1)2mt4m2-(21+1)21' m=l 2—0 1 x [4m2-(r/0)(22+1)21'1{exp[-(22+1)2t/01-exp[-(22+1)Zt/r]} X Sin(2mnx/a) (6.3) and T86 is given in Appendix D. We expect T855 to be very small as it is in the adiabatic case. 7. Calculated Temperature Variations In order to examine the behaviour of the tempera- ture solutions, we assign typical values for organic liquids to all of the coefficients and calculate the values of the solutions at various times during the ex- periment and at various positions in the cell. The ther- mal conductivity, heat of mixing, heat of transport, and initial concentrations are varied in order to demonstrate their respective contributions to the temperature 124 distribution Table 2 gives the values which have been as- signed to those coefficients which we have not varied in our calculations. These coefficients would not be correct for all systems, but they represent typical behaviour in any system at relatively short times (t < 20). Table 2--Typical values of physical properties for a binary mixture of organic liquids - 3 -3 - O - 0.8 x 10 kg m Dw/Do — 0.1 (inp)w = 0.1 (0DQi)w/(0D0i)o = 0.1 _ _ -3 _ -3 ()an)T - 1.0 x 10 (DDQI1T/(OD0110 - 2.0><10 —- _ 6 -1 -3 _ _ pCP - 1.4 x 10 J K m Ko/Kw - 0.1 _ _. _ _ _ -3 (pCP)W/(pCp)o — 0.1 Ko/KT — 2.0 x 10 (pC)/(OC) =-10><1o‘3 (H—fi') =20x1020k‘1x'1 p T p o ' , 1 2 T ’ g -9 2 -1 D = 1.4 X 10 m s The coefficients (Hi - fiZ’w have been calculated from typical heats of mixing (Rock, 1969; Lewis, Randall, Pitzer, and Brewer, 1961) as is shown in Appendix F. Heats of transport have never been measured in liquids, but by assuming that the heat matter Onsager reciprocal relation is valid so that (Horne and Bearman, 1967) 125 Qi = " alW1(3)Jl/3W1)T,Pp (701) we have calculated values from thermal diffusion factors. Thermal conductivity does not vary radically among systems which could be studied, but, because it has a marked ef- fect on the speed with which temperature variations in the cell relax, we consider two thermal conductivities. In all cases the results are presented in reduced time (t/O) and reduced cell coordinate (z/a). Because the length of the cell enters into the solutions only through these two factors, the results are the same for any length cell. The factor I depends on 3 also, but it can be re- placed in all solutions by (T/O)0. The factor (1/0) is independent of a. From the definition of 0(4.ll) we see that it is proportional to a2 and, thus, for longer cells much larger absolute times will be needed to reach the temperature distributions which are rapidly attained in Shorter cells. This fact will cause problems for the experimentalist in that, since his cell cannot be truly adiabatic, deviations from predicted behaviour due to non- adiabaticity will be larger at longer times. For this reason short cells should be used. In a 3 cm cell with D of Table 2, 0 is 108 minutes; for a 10 cm cell, 0 is 1200 minutes. All computations have been done on a, CDC 6500 computer with series truncation criteria similar to those discussed in Chapter III. More terms are needed at Shorter 126 times, and with double and triple sums the computational cost becomes quite high at these very short times (t/O i 0.01). Computation at forty—nine points in the cell was done simultaneously for each t/O because the time depen- dent parts for every point z/a are identical. The results presented here have been determined from T = T6 + TeG' (7-2) Since T855 is very small in the adiabatic cell, with -2 -1 -1 — - 4 '1 pDQi = 3 x 10 J m S , (Hl - H2)w =.5 x 10 J kg , K = 2 x lo-lJ m-ls-lK-l, and w0 = 0.8, equation (E.4) of 1 Appendix E yields a maximum value of 4.32 x 10-4K at t/O = 0.09. The value becomes more negative as t/O increases and asymptotically approaches its equilibrium value. The T566 term contributes less than 1.0% to the total temperature perturbation and is symmetric about z = 0, so that it does not affect the temperature dif- ference between any two symmetrical points in the cell. Because T266 is so small, we have omitted it from further work without thereby introducing any measurable error. The symmetries1of the solutions T and Te make 6 6 it possible to observe the Dufour effect. T6' which is due solely to heat of transport, is antisymmetric about 2 = 0. The T6 solution is symmetric about 2 = 0. The 6 'value of this solution depends on the variations of 127 physical properties and on the heat of mixing. It is generally small except for the contribution of a large absolute heat of mixing. However, if one is interested in measuring the Dufour effect, he must measure the tem- perature difference in the cell. The solution, TeG' be- cause of its symmetry adds nothing to the temperature dif- ference which is measured at two points symmetric about 2 = 0. These symmetries can be seen in the next twelve figures. Note that the ordinate scale of these figures changes from one to another. Figure 3 shows calculated temperature distribu- tions in an adiabatic cell with no heat of mixing, pDQi = 3 x lO-ZJ mfils-1 (a1 = - 0.9), and thermal con- ductivity of 2 x 10-1J mmlsle-1 at various times. The most important feature of this plot is that a temperature difference of approximately 0.12K exists between the right and left hand sides of the cell. This difference is due almost entirely to the Dufour effect, and we thus predict that the Dufour effect in liquid mixtures is measurable. The slight deviation of the curves from true antisymmetry about 2 = 0 is due to our inclusion of Spatial variations of coefficients in the cell. We observe that the temperature fluctuations arise in the center of the cell where there is a concentration gradient and that the temperature throughout the cell changes as heat is con- ducted and convected. We also see that the temperature 128 Figure 3—-Temperature variations in an adiabatic cell for 11111“ = 0.0, pDQ1= 3 x 10-2J m-ls_1, K = 2 x 10‘1J m'1s‘1K-1, and w: = 0.8 at (t/e) = 0.01(ooooo),(t/0)=0.05( ), (t/e) = 0.09 ( —-), (t/O) = 0.13 ( ————— ), and (t/0) = 0.17 (——-----). 0.12 1— 0.08 *- 129 0.04 8 l-g 0.0 I L'. -0.04 -0.08 *- - 0.12 *- l l I l J - 0.4 - 0 2 0.0 0.2 0.4 2 /a Figure 3 130 difference initially rises and then falls again, with the maximum difference being observed at 0.05-0.09 t/O. The fact that the temperature is quite stable for z > .35 and z < -.35 for these times Should enable the experimentalist to make accurate temperature measurements well inside the cell; 112;! end effects will not matter. We note that, Since 82.np/8w1 is positive, the end of the cell which originally has the greater density is that which has the greater concentration of substance 1, i;g;, the end for which z < 0. The temperature gradient arises to oppose this density gradient so that the end for which 2 > 0 is the warmer end. This is what would be ordinarily expected although a negative heat of transport would reverse the effect. The fourth figure shows temperature distributions in a system identical to that considered in the first figure except for the thermal conductivity. The lower thermal conductivity produces much larger temperature ef- fects because heat is not conducted so rapidly. Measure- ments will be easier for systems with low thermal conduc- tivities. We have again studied the first system in Figure 5 where in this instance (Hi - E21w = 5.0 x 104 J kg_1. which corresponds to a heat of mixing (Afih) of - 6.5 x 103 J kg-l. We see that this causes a general rise in the tem- perature of the entire cell. The temperature difference 131 Figure 4--Temperature variations in an adiabatic cell for AH = 0.0, pDQ* = 3 x 10‘2J m'1s'1, K = 91 -l 1 1--l o l x 10 J m s' K , and wl = 0.8 at (t/0) 0.01 (00 o co ), (t/G) = 0.05 (—————) (t/O) = 0.09 ( —), (t/O) = 0.13 ( ______ ), and (t/0) ='0.l7 (—-—---—-—-). I 132 /’::€::=-.- _ _ _ _ _ _ _ 2 8 4 O 4 8 2 m. 0. O 0. O 0. .I 0 O O 0 n_u n_u _ AXVEEhIC 0.0 0.2 0.4 2 /a -0.2 -0.4 Figure 4 133 Figure 5--Temperature variations in an adiabatic cell 1 for AH = - 6.5 X 103 J kg- , pDQi = 3 X J m_15—11K = 2 X lO-lJ'm_ls_1K-l, and mi 0.8 at (t/O) = 0.01(OO 00.), 10‘2 (t/O) = 0.05 ('———-—————-)p (t/G) = 0.09 ( —~ '-'-' 0.13 ( _____ )r and (t/e) = _ ). 0.17 ), (t/G) (T-Tm)/(K) 1.2 0.8 0.4 134 / / — / —— / ’ —————————— w a I ’1 L— -"” /./-~-‘-‘ ’/ / -/‘ -——--—-'— h- o... O C .0. O .0 '0. ‘0. ‘..' ‘..". "0¢boo .. Figure 5 135 due to the Dufour effect is remarkably constant, however, at about 0.12K for t/G = 0.05-0.13. Because the overall heating of the cell will cause slight errors in constants used even though the constants can be calculated correctly at any temperature, it would be best to measure the tem- perature difference as soon as it is well established (t/e = 0.09) and before it begins to decay. Our next three figures are also for the adiabatic cell, but the effects of varying various constants are illustrated at a fixed time (t/e = 0.1). Figure 6 illus- trates the variations of the temperature distribution with (El — fié)w or heat of mixing. Values used for (ii - fié)w are 1.5 x 105, 5.0 x 104, o, and - 5.0 x 104J kg-'l which represent heats of mixing ranging from 3 1 l -19.5 x10 to 6.5 x 103 J kg- . We see that large J kg- heats of mixing cause much greater effects on the tem- perature than do heats of transport. Illustrated in Figure 7 is the variation of the temperature difference with heat of transport. Values 2 2 3 2 for pDQi are 6 x 107 , 3'x 10‘ , 3‘x 10‘ , and - 3 x 10‘ J mfls-a. The small heat of transport is swamped by the heat of mixing and has a temperature difference of 0.012K. In every case we have seen that the temperature differ- ence at its maximum is directly proportional to the heat of transport. 136 Figure 6—-Temperature variations in an adiabatic cell for (t/G) = 0.1, pDQ* = 3 x lO-ZJ m-ls_l, _ -l -d.-l -l o _ . K - 2 x 10 J m s K , and ml - 0.8 With Afim= - 19.5 x 103 (——-—-), - 6.5 103 ( —————— ),o.0(————),and 6.5 x 103 J kg’l (—————). 137 , / 4m , , . _ I .2 . o. . _ — _ _ r. n _ 4 _ .0. _ : _ _ __ _ __ _ _ 4. 0. AN 2 8. 4. 0. 4. 8. 0‘ 0‘ 1| 1| AU AU nu JV mu 3: \ “Eh ..t 0.4 0.2 z/a Figure 6 . J mum-"I. .__ 138 Figure 7-—Temperature variations in an adiabatic cell for (t/e) = 0.1, Aim = - 6.5 x 103J kg'l, K = 2 x 10‘1J m‘ls'lx'l, and w? = 0.8 with '2 ( ), 3 x 10’2 (-—-—-——-—-), and (T - Tm)/(K) 0.9 139. Figure 7 140 The last graph for an adiabatic system, Figure 8, shows how greatly temperature variations in the cell are reduced as the two initial mixtures become more similar. An 0.075K difference can still be seen with an initial difference of 0.5 in the weight fractions. The heat of mixing effect has been greatly reduced, however, as it is proportional to wiz. Some sacrifice of temperature difference for suppression of the heat of mixing might be advisable in experimental systems with large heats of mixing. Interesting to note are the very small variations of temperature (0.005K) in the region of mi = 0.05—0.01. Such variations would not affect diffusion measurements at these relative concentrations. Our final six figures deal with a cell whose ends are diathermic. These figures present a much greater challenge to the eXperimentalist because the temperature fluctuations are much smaller and there is no temporal region in which the temperature distribution remains relatively constant. These six figures correSpond to the Figures 3-8 except that Figure 14 has a heat of mixing. Ianigure 9 we see that our temperature difference falls rapidly with increasing time. At t/O = 0.05 the differ— ence is 0.07K and at t/0 =_0.17 it has fallen to 0.04K. The maxima occur between 0.15 and 0.25 z/a and the minima between - 0.15 and - 0-25 z/a. The curves are again not symmetric, but they are more so than in the adiabatic 141 Figure 8--Temperature variations in an adiabatic cell for (t/O) = 0.1, Afih = 0.0 J kg-l: K = 2 x 10'1J m‘ls‘lx'l, and one; = 3 x 10' J m-ls-lwith w‘1’= 0.3 (o o o o u), 0.5 < -—-), 0.2 (--————-), 0.05 (————--——— ), and 0.01 (-————-—-—-). 2 0.12 0.08 0.04 (T- Tm)/ (K) .0 o I .0 C) a. -0.08 -O.'|2 00.00.00. +— Figure 8 f " ----_--' 142 0.0 0.2 0.4 143 Figure 9--Temperature variations in a cell with diathermic ends for Afih = 0.0, pDQi = 3 X 10.2 J m-ls-l I K = 2 x 10-1 J m'ls‘lx-l, and mi = 0.8 at (t/G) = 0.05 (———-*-———-), (t/e) = 0.09 ( ______ ), (t/e) = 0.13 (———— (-————-———). ), and (t/O) = 0.17 144 0.06 f" _\ 0.03 .— / - _\\ 3 E I t: -0.03 P'— \ ‘_/l -0.06 ‘- J J l L 1 -0 4 '0 2 0.0 0.2 04 Figure 9 145 case. Halving the thermal conductivity again almost doubles the observable temperature difference as seen by comparison of Figures 9 and 10. The introduction of a heat of mixing in Figure 11 changes the temperature distribu- tions little from the original curves of Figure 9. The curves are raised only a few hundredths of a degree due to rapid replacement of the heat of mixing through the end walls so that the overall temperature of the cell can- not rise appreciably. Figure 12 better emphasizes the greatly reduced effect of the heat of mixing, especially compared to Figure 6. That the diffusionist needs worry even less about difficulties arising from temperature fluctuations in a diathermic cell than about those in an adiabatic cell is attested to by Figure 14. 8. Discussion We have attained temperature solutions both for adiabatic cells and for cells with diathermic ends when initially a sharp concentrations boundary exists in the center of an initially isothermal cell. From these solu- tions we find that the Dufour effect is measurable, that an adiabatic cell produces larger temperature differences than a cell with diathermic ends, that high heats of mix- ing do not cause significant variations in temperature changes due to the Dufour effect in a cell with diathermic ends, that low thermal conductivities and low heats of 146 Figure lO--Temperature variations in a cell with diathermic ends for Afih = 0.0, pDQ* = 3 x 10—2 J m-ls-l, K = l x 10-1 J m_ls-1K‘}, and w: a 0.8 at (t/O) = 0.05 (———---———-), (t/G) = 0.09 (—~—-—-—-—-)p (t/@) = 0.13 (‘—__‘ —_—-), and (t/e) = 0.17 (————-—-). (T-Tm)/(K) 0.06 0.03 0.0 -0.3 147 Figure l0 I/\ \ £7“. _ M 'l .I/ ,’ I // I’l’ __ \ I," 0‘ I \K, /, ,/ \ l __ \:~ / 1 V , 1 I -0.4 -0.2 0.0 0.2 0.4 z/a 148 Figure ll--Temperature variations in a cell with diathermic ends for Afih = - 6.5 X 103 J kg-l, pDQi = 3 x 10.2 J mfls-l, K = 2 x 10-1 J m-ls-lK-l, and w‘1’= 0.8 at (t/O) = 0.05 ( — ). mm» = 0.09 ( —————— ), (t/e) = 0.13 (———— ). and (t/O) = 0.7 (-—————-—-). 149 0.6 r 332:; :5 -0.03 '— -0.06 - 0.4 0.2 0.0 z/a -0.2 -0.4 Figure 11 ‘. 'hffltcfl Figure 12-—Temperature variations in a cell with diathermic ends for (t/O) = K = 2 x 10- AK m 6.5 x 10 guishable. 3 1 J kg— J m- — 19.5 x 103, - 5,5 x 103, 0.0, and l 0 1s- 150 .1, pDQi = 3 x 10"2 J m'ls‘l, 1K‘1, and mi = 0.8 with The curves are indistin- 151 0.0 0.2 0.4 z/a - 0.2 -0.4 0.03 ~— 0.02 —- 0.0] ._ 0 D 0 CSIEFIC '-C)CH P— -0.02 .— ‘0 03 Figure 12 152 Figure 13--Temperature variations in a cell with diathermic ends for (t/O) = 0.1, Afih = - 6.5 x 103 J kg'l, K = 2 X 10”1 J m-ls-lK-l, and mi = 0.8 with pDQg=6x10'2( —),3><1o‘2 (——— ),3x10‘3(-———).and-3x 10-2 J m-ls-l ( ------ ). 0.6 0.3 - 0.03 - 0.06 153 Figure 13 g ,/ ‘\\ ' \\ ’ ‘ ’ \ \\ / \ _ \ r“ / \ / \ / \/ L... I I l l 1 -04 -O.2 00 0.2 0.4 154 Figure l4--Temperature variations in a cell with diathernnil: ends for (t/G) = 0.1, Ail-m = - 6.5 X 103 J kg-J': K = 2 X lO-1 J m-ls-lK-l, and pDQi = 6 X 10"2 J'm-ls-l with w‘1’= 0.8 (o o o ...), 0.5 (—-——-),0.2( ------ )p0.05 (—_—)I and 0.01 (——————). 155 oooouvsg 1A 0 no. \ xx \ 0 \ \ _ ~ . _ Ila/m / ., 1 o OOOOOI/ z/I ooooyu% nu. ..zbbfiooooo .10 Izle/luooooo ’ .u /. 12 nu \ \ \\oooo.a 4 3.2.1.... r P _ . _ _ m. m m. w m. m m o o o ... ... ... z/a Figure 14 . 156 mixing are conducive to easier measurement of the Dufour effect, and that the Dufour effect becomes very small as the initial concentration difference decreases. Using our results we can analyze the work of Rastogi gt_gl. (1965, 1969, 1970). First, they neglected heat of mixing. We find this a very poor assumption in that the temperature at all points of the adiabatic cell depends strongly on the heat of mixing. However, our re- sults do show that, if the points between which the tem- perature difference is measured are symmetric about 2 = 0, the heat of mixing does not appreciably affect the tem- perature difference measured. The assumption by Rastogi gt_§1. that coefficients are constant is not correct but does not introduce large errors. Their cell design is not really comparable with our idealized cell in that their initially sharp concentration boundary was achieved by a stop cock whose diameter was only half that of the cell. Thisconstriction was not taken into account in their results. Their cell was 23 cm long and measurement was made at z/a i 0.24. This falls in the region which does not attain a steady temperature difference as dis- cussed in Section 7. The maximum temperature difference between these two points was reached at very low reduced time; and the concentration difference between the points was assumed to be the initial concentration difference. From Figures 3 and 4 we observe that the maximum 157 temperature difference varies with the z/a used while the assumption that the concentration difference is the initial one is still reasonable. Rastogi and Yadava (1970) recog- nized this and took the variations into account with their experimentally determined parameter 6. Unfortunately their results are incomprehensible because one given distance between their thermocouple and stop cock is greater than the given half length of the cell. As we have shown, it is the value z/a, not 2, which is important to the tempera- ture distribution. We hope that the results of our work will provide a useful guide to the experimentalist. The solutions given are such that, if all diffusion coefficients, heat capaci- ties, 232;! are known, they can be substituted into the temperature equation, and the heat of tranSport can be adjusted to fit the measured temperature variations. Our results also indicate that the Dufour effect causes no complications in diffusion experiments with very small concentration differences. BIBLIOGRAPHY BIBLIOGRAPHY Bartelt, J. L., Ph.D. Thesis, Michigan State University, 1968. and F. H. Horne, Pure Appl. Chem., 22, 349 (1970). Bearman, R. J. and J. G. Kirkwood, J. Chem. Phys., 33, 136 (1958). Carslaw, H. S. and J. C. Jaeger, Conduction of Heat in Solids (Oxford U. P., London, 1959), 2nd ed. Clusius, K. and L. Waldmann, Naturwiss., 32, 711 (1942). Coleman, B. D. and V. J. Mizel, Arch. Rational Mech. Anal., 13, 245 (1963). and v. J. Mizel, J. Chem. Phys., 4_o, 1116 (1964). and W. Noll, Arch. Rational Mech. Anal., 12, 167 (1963). Doria, M. L., Arch. Rational Mech. Anal., 3, 343 (1969). Dufour, L., Poggend. Ann. Physik, 148, 490 (1873). Fitts, D. D., Nonequilibrium Thermodynamics (McGraw-Hill, New York, 1962). Grew, K. E. and T. L. Ibbs, Thermal Diffusion in Gases (University Press, Cambridge, 1952). de Groot, S. R., L'Effet Soret (North Holland Pub. Co., Amsterdam, 1945). , Thermodynamics of Irreversible Processes (North Holland Pub. Co., Amsterdam, 1952). and P. Mazur, Nonequilibrium Thermodynamics (North Holland Pub. Co., Amsterdam, 1962). Gyarmati, I., Nonequilibrium Thermodynamics (Springer- Verlag, New Yofk, N.Y., 1970). 158 159 Haase, R., Thermod namics of Irreversible Processes (Addison- WesIey P55. Co., Reading, Mass., 1969). Horne, F. H. and T. G. Anderson, J. Chem. Phys., 53, 2332 (1970). 7" and R. J. Bearman, J. Chem. Phys., 46, 4128 (1967). Kestin, J., A Course in Thermodynamics, Vol. I. (Blaisdell Pub. Co., Waltham, Mass., 1966). Kirkwood, J. G. and B. Crawford, J., J. Phys. Chem., 56, 1048 (1952). Lewis, G. N., M. Randall, K. S. Pitzer, and L. Brewer, Thermodynamics (McGraw-Hill, New York, N.Y., 1961) 2nd ed. McLaughlin, E., Chem. Rev., 34, 389 (1964). Meixner, J., Ann. Phys., 22, 244 (1943). and H. G. Reik, Handbuch der Physik, Vol. III/2, S. Flugge, Ed. (Springer-Verlag, Berlin, 1960). Miller, L., z. Naturforsch, 43, 262 (1949). Mfiller, 1., Arch. Rational Mech. Anal., 36, 118 (1967). , Arch. Rational Mech. Anal., _2_§_, 1 (1968). Olson, J. D., Ph.D. Thesis, Michigan State University, Onsager, L., Phys. Rev., 31, 405; 38, 2265 (1931). Prigogine, I., Etude Thermodynamique des Phenomenes Irreversible (Editions Desoer, Liege, 1947). , Introduction to Thermodynamics of Irreversible Processes (Interscience Pub., New York, N.Y., 1955). Rastogi, R. P. and G. L. Madan, J. Chem. Phys., 43, 4179 (1965). and B. L. S. Yadava, J. Chem. Phys., 51, 2826 (1969). and B. L. S. Yadava, J. Chem. Phys., 52, 2791 (1970). 160 Rock, P. A., Chemical Thermodynamics (MacMillan Co., London, 1969). Tree, D. R. and W. Leidenfrost, "Thermal Conductivity Measurements of Liquid Toluene," Thermal Conduc— tivit , Proceedings of the Eighth Conf., Ho and Taonr, Eds. (Plenum Press, New York, 1969). Truesdell, C., Rend. Lincei (8) 22, 33 (1957). and W. Noll, Handbuch der Physik, Vol. III/3, S. Flugge, Ed. (Springer-Verlag, Berlin, 1965). and R. A. Toupin, Handbuch der Ph sik, Vol. III/1, S. Flugge, Ed. (Springer-Verlag, Ber 1n, 1965). Waldmann, L., 2. Phys., 114, 53 (1939). , 2. Phys., 121, 501 (1943). , Z. Naturforsch, 1, 59 (1946). , 2. Phys., 124, 2, 30 (1947a). , z. Naturforsch, 23, 358 (1947b). , Z. Naturforsch., 43, 105 (1949). APPENDICES APPENDIX A SOME EQUATIONS OF THERMOSTATICS AI = EI - TS (A.1) dEI = TdS - Pd(1/p) + (01 - u2)dwl (A.2) HI = EI + VP (A.3) - dP + wldml - p2) + SdT = 0 (A.4) pHI = {paid (A.5) Ha = 1.1a + TS“ (A.6) s = - aAI/BT (A.7) Ewan“ = AI + P/p (A.8) 11a = AI(l - wa) (aAI/aw1)p,'r+92(3AI/3°)w1,'r (A.9) pa = AI + p(8AI/8pa)pB’T (11.10) (“1"“2/9)‘=(3AI/3°1)p2,m"(aAI/392)pl,T .= 1/°(3AI/3w1)m,p (A.ll) P = 991(3A1/391)p2,m + 992(3AI/892)91,T = pz‘aAI/3°)w1,m (A.12) 161 APPENDIX B A RATIONAL MECHANICAL DEVELOPMENT OF THE TIP THEORY The heuristic but very successful TIP results of nonequilibrium thermodynamics have often been criticized for their lack of mathematical rigor by the prOponents of rational mechanics. It is, however, quite possible to de- rive the results of TIP with the formalism of rational mechanics by redefinition of thermodynamic process for a mixture in such a way that there is no difference between the specific partial stress tensors and there are no in-‘ ternal forces among components. The thermodynamic process is then defined by the Bv + 5 functions which satisfy the balance of mass for each component, the balance of linear momentum for the mixture as a whole, and the balance of energy. Now we can further assume that the external forces, bf, may be solved for using the linear momentum balance equation. There are no interaction forces. With this limited definition of a mixture, the TIP results are imme- diately derivable from a linear rational mechanical for- malism. Our balance of energy equation is 162 163 p(dE/dt) = (avi/axj)Tij-(3q;/3xj) + pr + (0102/0)ui(bi - b? l). (3.1) This can be placed in the entropy balance equation using the thermodynamic identity A = E — TS. (B.2) The entropy balance equation which results is pT(dS/dt) = - p(dA/dt) - pS(dT/dt) + (avi/ij)Tij - (aqg/axj) + (9192/9)ui(bi - bi).(B.3) Until this point we have not needed to distinguish among the independent variables and the functionals determined by constitutive equations. Studying our definition of thermodynamic process and those terms which appear in the energy balance equation we let our independent variables be 3p1 3p2 8T 3V1 91' 92' T' 3x.’ 8x.’ ax.' 3x.’ 3 J J J (E.4) and our constitutive relations in terms of these variables will define ui, Tij, qg, fi, A, s. (3.5) Following the principle of equipresence in the linear limit, we obtain constitutive relations of the form 164 A = A(pl, 02: T) (8.6) Tij = - Paij + ¢(8vk/axk)sij + 2n(3vj/axi) (B.7) * = _ n _ n _ n qj qu(3p1/3xj) Cq2(892/8xj) CqT(3T/3xj) (8.8) kj = - Cfll(3p1/3xj) - C£2(392/3xj) - C£T(3T/3xj) (8.9) uj = - C31(apl/3xj) - C32(apZ/8xj) - C3T(8T/8xj), (3.10) where k3. is defined by q? k. = f - -l (3.11) The derivatives dA/dt and Bkj/axj can be expanded in terms of the independent variables and dpa/dt can then be re- placed by the balance equations: (do /dt> = - (92/9)u (39 /3x ) - (92/9)u (89 /ax ) l 2 j 1 j l j 2 j - 9192(8uj/3xj) - 991(3vj/3xj) 2 2 (dpz/dt) (92/9)uj(391/3xj) + (pl/p)uj(392/3xj) + Bu. 3x. — 3v 3x. . 8.12 9192( J/ J) 992( j/ J) ( ) An immediate consequence of this formalism is kj = 0. (B.13) The total entropy production is readily rearranged to the form 165 9T9 = (“1 - 02)[3(9192/9)uj/3xj] * + (Tij + P)(8vi/3xj) + qj(8T/3xj) + (0102/0)ui(b% - bi). (9.14) The coefficient P is identified as the thermostatic pressure, P = ptpl(aA/ap1) + 02(8A/302)] (3.15) and the usual thermostatic definition of chemical potential has been used. The result 8 = - aA/aT (8.16) is a consequence of the Clausius-Duhem inequality. Re— arrangement of the entrOpy balance equation (8.3), then leads to the TIP result 9T¢ = 3[(ui - 95)9192uj/p]/3xj (9192/9)uj[3(ui - u§)/3xj] + q§(8T/3xj) + (Tij + Pdij)(3vi/8xj). (8.17) .-- 3 In: ml.‘-—'.al—-._t.vdn -.' I. h :L APPENDIX C THE TEMPERATURE SOLUTION, T FOR THERMAL ll CONDUCTIVITY OF A PURE FLUID IN A FLAT PLATE CELL The solution of equation (III.3.13) for n = 1, subject to the conditions Y = Yu = YL and Tod-TL:=2TM, is T1(z,t) = (4/fl32(AT)2££0[(KT/Ko)(2£+1)-l (2£+1)-2{1 - exp(-(21+1)2t/T]} 2x[(22+1)2x-11-1{exp(-t/y) exPI-(22+l)2t/T]} + x[(22+1)2x-2]-1{exp(-2t/Y) - epr-(2£+l)2t/r]} + 2 -+ Z](-l)2cos[(2£+l)nz/a] (c.1) l 2 ‘where x = Y/T. (0.2) 2 = Z (4n2x - 1)'1{2(30(22+1)‘l l 1v=l - [2(KT/Ko) - 301(22+1)[2n+22+11’1[2n-(22+1)1‘1 - [(pCé)T/(QC§)O - (KT/Ko)] 166 + X 167 16n(22+1)[2n+22+1]‘2[2n-(22+1)1’2} X[(2£+1)2X - l]_l{exp(-t/Y) - exp]-(22+l)2t/r]} [(21+1)2-4n21'1{exp[-4n2t/r] - exp[-(2£+1)2t/T]} x[(22+1)2x - 21'1{exp(-2t/y) - exp[-(22+1)2t/T]} x{((22+1)2—4n21x - 1}-l{exp[-4n2(t/T)-(t/Y)] exp [-(22+1)2t/T]} ( 2 (4n2x - l)-l(4m2x - l)-l m=l (480(2n-(2£+l)]-l[2n+(2£+l)]_14-{[(KT/Ko)-Bo]4(2£+l) [4n2+4m2+(2£+l)2] [(pEé)T/(pfié)o - (KT/Ko)]32(2£+l)m2 } {[2(n+m)+2£+1][2(n+m)-(2£+l)][2(n-m)+(2£+1)] [2(n-m)—(22+1)1}"l (x{[(22+1)2-4m21x - 1}‘1{exp[-4m2(t/r)+4m2(t/y)1 exp[-2(2£+1)2t/T]} [(22+1)2-4(n2+m2)1‘1{epr-4(n2+m2)t/rl exp(-(22+1)2t/rl} xt(22+1)2x-21‘1{exp[-2t/y1 - exPI-(2£+l)2t/T]} x{[(2£+1)2-4n2]x - 1}‘1{epr-4n21 exp[-(22+l)2t/T]} C.3) (C. 4) APPENDIX D THE TEMPERATURE SOLUTIONS, T66' FOR THE DUFOUR EFFECT IN ADIABATIC CELLS AND CELLS WITH DIATHERMIC ENDS Adiabatic In the adiabatic cell, the solution T86 from Eqs. (IV.5.18) and (IV.4.17) has the form 2 _ oz -2 -1 ” - T86 - 2ml n Ko 2: [(22+1) B1(£) o + Z (-l)mBz(£,m)cos(2mnx/a)] m=l where 31(2) = c1(2) + c2(1) (v.1) B2(£,m) = C3(£,m) + C4(£,m) + C5(£,m) + C6(2,m) + C7(£,m), with 1/0 = y C1(2) = 0.5{(95§)W(DD03)0[(95') K (l‘Y)]- p o o (9.2) 1 -1 —- - —| -1 + (£np)w(pDQi)o[Ko(1-Y)] - (9D)O(Hl-H2)w[(ocp)onol } x {exp [-2(22+1)2(t/e)] - 1} 168 rd “Ali—m‘l' ... ' ”I. .n .- I t .... 169 _ _ - - _ -l C2(£) - { (90%)w(pDQi)o [(pCp)o(1 y)] - (2np)w