NONEQUILIBRIUM THERMODYNAMICS 0F FLUID MIXTURES: PRINCIPLES, PERTURBATION METHODS, AND PARAMETER ESTIMATION Thesis for the Degree of Ph. D. MICHIGAN STATE UNIVERSITY JOHN LESTER BARTELT 1968 mm F) LIBRA 1’" ' I a §«ii.;"rxi.;;.r: Static ‘ . {scram-sir)! ( Lrantrtam-r This is to certify that the thesis entitled NONEQUILIBRIUM THERMODYNAMICS OF FLUID MIXTURES: PRINCIPLES, PERTURBATION METHODS, AND PARAMETER ESTI MAT I ON presented by John Lester Bartelt has been accepted towards fulfillment of the requirements for Ph.D. degree in Chemistry Wflw Major professor Frederick H. Horne Date ll November 1968 0-169 I lmomn or i I ms a sour : angx mm ma Ll ”IV .IInI-E ABSTRACT NONEQUILIBRIUM THERMODYNAMICS OF FLUID MIXTURES: PRINCIPLES, PERTURBATION METHODS, AND PARAMETER ESTIMATION BY John Lester Bartelt A fundamental nonequilibrium thermodynamic theory and powerful, accurate methods of exploiting it have been developed to deal with the intricate problems encountered when various, interacting transport phenomena occur simul- taneously. Critical examination of the fundamental theory has led to a reformulation which embodies a new approach to the specification of phenomenological relations through the rational use of the entropy inequality and proper enumeration of independent variables. The approach re- solves most of the ambiguities in the earlier theory with regard to validity of a Gibbsian differential equation in nonequilibrium situations, pressure and mechanical equi- librium criteria, kinetic energy of diffusion, and the appropriateness of the Newtonian stress formula for mix- tures. A perturbation expansion method has been developed to facilitate the solutions to a wide range of problems in John Lester Bartelt transport theory. Coupled with the weighted residual method and finite transform method, this approach is shown to be an extremely versatile technique for handling such difficult problems as the solution to partial differential.equations with variable coefficients and the convection problem of pure thermal diffusion. Some practical formulas are de- rived for the latter case. A detailed theory of estimation of nonlinearly interacting parameters from data with cor- related errors is presented together with a computational algorithm which has distinct advantages over the ordinary least squares method. NONEQUILIBRIUM THERMODYNAMICS OF FLUID MIXTURES: PRINCIPLES, PERTURBATION METHODS, AND PARAMETER ESTIMATION BY John Lester Bartelt A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1968 To Brooks ii ACKNOWLEDGMENTS I would like to express my thanks to the Dow Chemical Company, the Lubrizol Foundation, and the National Institutes of Health for the fellowships which provided my financial support. I am grateful to my colleagues Dr. T. G. Anderson and Mr. V. A. Nicely for the many hours of fruitful dis- cussion and for the numerous suggestions and comments which helped to make this thesis a reality. I would also like to thank Mr. D. Knirk for making his plotting program available and for his interest in the computer aspects of this work. I wish especially to express my sincere appreci- ation to Dr. F. H. Horne, whose counsel and guidance have been of immeasurable value. He has been encouraging, critical, and above all a conscientious teacher. Finally, I am deeply grateful to my wife Brooks, who endured my difficult times, and whose understanding and encouragement were unlimited. iii TABLE OF CONTENTS ACKNOWLEDGMENTS O O O O C O O O O O C C 0 LIST OF FIGURES . . . . . . . . . . . . . Chapter I. II. III. Iv. V. INTRODUCTION 0 O O I O O O O O O O A THERMODYNAMIC THEORY OF FLUID MIXTURES 1. Introduction . . . . . . . . . The Clausius-Duhem Inequality Thermodynamic Processes . . . Constitutive Assumptions for a Fluid Mixture . . . . . . . 6. Consequences of the Entropy Inequality . . . . . . . . . 7. Ordinary Binary Fluid Mixtures (The Linear Theory) . . . . 8. The Transport Equations . . . 9. Discussion . . . . . . . . . . £11wa 0 O PERTURBATION EXPANSION METHODS . . 1. Introduction . . . . . . . . . 2. The Experiment . . . . . . . . 3. Pure Thermal Diffusion Transport Equations . . . . 4. The Perturbation Expansions . 5. The Perturbation Equations . . 6. Summary . . . . . . . . . . . PARAMETER ESTIMATION . . . . . . . Introduction . . . . . . . . . Generalized Least Squares Adjustment of Data . . . . NH 3. An Alternative Approach . . . 4. Fringe Shape Analysis . . 5. Time Constant Analysis . . 6. Summary . . . . . . . . . . . CONCLUSION . . . . . . . iv Kinematics and Equations of Ba ances Page iii vi 18 26 3O 31 34 38 51 56 59 59 65 69 72 75 96 97 97 100 107 111 115 117 118 BIBLIOGRAPHY APPENDIX A. APPENDIX B. APPENDIX C. RESTRICTIONS ON PHENOMENOLOGICAL COEFFICIENTS IMPOSED BY POSITIVE ENTROPY PRODUCTION PARAMETER ESTIMATION PROGRAMS VELOCITY AND TEMPERATURE PROFILES FOR THE CONVECTIVE HEAT LOSS PROBLEM Page 122 125 130 148 LIST OF FIGURES Figure Page 1. Vertical velocity as a function of r at the center of the cell (2 = 0.5) . . . . . . 150 2. Radial velocity as a function of r at z = 0.25. O O I O O O O C O O O O O O O 151 3. Temperature correction versus r at z = 0.5 for (Tr - Tm + 1/2 AT) = - 5.0 o o o o o o o 154 4. Temperature correction versus r at z = 0.5 for (Tr - Tm + 1/2 AT) = 0.0 o o o o o o o o 155 5. Temperature correction versus r at z = 0.5 for (Tr - Tm + 1/2 AT) = 5.0 o o o o o o o o 156 vi CHAPTER I INTRODUCTION The mathematician has often been described as one who neither knows what he is talking about nor cares whether what he says is true. The physical scientist often prides himself on his practical understanding, thoroughness, and insatiable curiosity. Between such extremes of abstraction and realism it would hardly seem possible that there should be much progress. Yet the trend has been quite otherwise and today the sci- entist is increasingly aware of his need for mathematical insight and the mathematician proves more and more the stimulation of physical problems. The progress of the moment, at any rate in the science of materials, lies in the regions where mathematics and physical science have common interests. If the mathematician has little care whether what he says is true, it is only in the sense that his primary concern is with inner consistency and deductive conse- quences of an axiomatic theory. He is content with cer- tain undefined quantities and his satisfaction lies in the structure into which they can be built. If the sci- entist regards himself as dedicated to the progression of understanding, then he cannot rest content with par- ticular details. It is his understanding of the common features of diverse problems that allows him to progress and hence he must be concerned with abstraction and gen- eralization. It is the business of mathematical theory to provide just such an abstraction and generalization, but it will do it in its own fashion and use the axio- matic method. From what at first seem rather farfetched abstractions and assumptions, it will produce a coherent body of consequences. Insofar as these consequences cor- respond to the observable behavior of the materials the scientist handles, he will have confidence in the mathe- matical theory and its foundations. The theory itself will have been used to design the critical eXperiments and to interpret their results. If there is complete discordance between the valid expectations of the theory and the results of critically performed experiments, the theory may be rejected. Some measure of disagreement may suggest modification of the theory. Agreement within the limits of experimental error gives confidence in the math- ematical model and opens the way for further progress. In particular nonequilibrium thermodynamics provides a model of the real world in which the scientist can have a high degree of confidence. Although the early steps in the development of the theory of irreversible processes took place over a hundred years ago, the cornerstone of the modern theory was laid in 1931 by Onsager. Over the past twenty-five years a coordinated theory based on the fundamental work of Onsager has been developed by Prigogine (1947, 1955), de Groot (1945, 1955), de Groot and Mazur (1962), Meixner and Reik (1959), Kirkwood and Crawford (1952), Fitts (1962), and many others. It is also clear that over the last decade there has been a renaissance of interest in thermodynamics and rational mechanics in the mathematical world. It has been fairly widespread and attracted the attention of many mathematicians whose abilities are of the first order. If a few names are to be singled out, it would probably not be unfair to the others to select those of Truesdell and Toupin (1960) and Coleman and N011 (1963); their scholarship and extensive writing on the foundations of continuum mechanics have been of great influence. Even though the development of the theory has been extensive, the fact remains that, because of the diversity of phenomena and range of material behaviors that come within the scope of nonequilibrium thermodynamics, it is unlikely that every possible area for study has been examined. The investigation of one such area initiated most of the research in this thesis: convection in pure thermal diffusion experiments. Convection in liquids due to temperature, density, or composition gradients is an extremely delicate phenomenon. The pure thermal diffusion experiments of Thomaes (1951) and of Agar and Turner (l960),ampng others, indicate that proper interpretation of their results can only be made if the convection that is apparently present can be accounted for. Because of its simplicity and sensitivity the pure thermal diffusion method has the potential, if the convec— tion problem can be solved or eliminated, of being an impor- tant source of transport parameters of interest not only to physicists and chemists but engineers and biologists as well. Unfortunately, it appears that the rule of thumb in the past has been to neglect convection and its effects whenever possible. Because of the paucity of theoretical and practical work on convection (hydrodynamicists had never considered convection under conditions appr0priate to the thermal diffusion experiments), it was impossible to establish the convective behavior of a fluid mixture in a thermal diffusion experiment without a re-examination of the theory and without new (or better) mathematical methods. Besides convection, there were and are many other unsolved transport problems. Throughout the research which has led to this thesis, the emphasis has been on the devel- opment of a generalized theory and of generalized methods of using it. Most of the ambiguities of the previous theory have now been resolved, and powerful methods have been obtained for applying the theory to particular experi- ments. Although examples are given only for pure thermal diffusion, the theory and methods are sufficiently general that a great variety of other problems may be solved. A reformulation of the nonequilibrium thermo- dynamic theory of fluid mixtures is presented in Chapter II. The questions of theoretical interest which are dealt with in this chapter are: (l) validity of a Gibbsian dif- ferential equation in nonequilibrium situations; (2) neglect of inertial and viscous terms in the phenomeno- logical equations; (3) neglect of the kinetic energy of diffusion; (4) definition of mechanical equilibrium; and (5) appropriateness of the Newtonian stress tensor for mixtures. The resolution of each of these problems is accomplished in this chapter through the development of a phenomenological theory for "ordinary" fluid mixtures. Chapter III is concerned with the development of the methodology required to solve the extremely complex transport equations deduced in Chapter II. In particular it is shown that the perturbation expansion method is an extremely versatile technique for handling such difficult problems as the solution to partial differential equations with variable coefficients and the convection problem of pure thermal diffusion. Some practical formulas are pre- sented for the interpretation of pure thermal diffusion experiments including the possibility of convection and variable phenomenological coefficients. CHAPTER II A THERMODYNAMIC THEORY OF FLUID MIXTURES 1. Introduction The success of the thermodynamic theory of irre- versible processes in its application to a wide variety of problems in the physics and chemistry of materials has led to a renewal of interest in the fundamentals of the theory. The material presented in this chapter is in fact a result of the need for better understanding of the prin- cipal assumptions and approximations made in the theory. The purpose here is not the derivation of a new theory; it is instead a reformulation in which the nature of these assumptions is clarified. Moreover, the reformulation has uncovered many new features of the theory which previously were buried in assumptions too casually accepted. There is an extensive literature for the thermo- dynamics of irreversible processes (TIP) or nonequilibrium thermodynamics (NET) and its applications. Although a literature survey is not intended here, a reference to TIP or NET as it is currently practiced is generally meant to be a reference to the works of Fitts (1962), de Groot and Mazur (1962), Hasse (1963), and Meixner and Reik (1959). TNIere is in addition a large volume of material in the technical journals, much of which can probably be found most easily by consulting the extensive bibliography of Fitts; specific articles of particular importance from the literature will be cited individually as the need arises. The objectives of this chapter are (l) the develop- ment of a set of macrosc0pic transport equations from basic principles which describe the behavior of a liquid mixture, with an attempt at each stage of the develOpment to indi- cate clearly the introduction of any necessary assumptions or approximations; and (2) the comparison of the differ- ences between the results obtained here and those of the earlier theory (see Horne, 1966). In View of these objec- tives it will be necessary in the remainder of this section to give a concise exposition of TIP and to indicate at each opportunity the need for a more rigorous or objective ap- proach. There is no doubt that in such a short exposition the case for TIP may not be presented as fairly as possible and where matters of Opinion are involved it might be wise to consult the original sources. Typically, TIP begins with the fundamental princi- ples of conservation of mass, momentum, and energy. These basic postulates are usually given in integral form, and, in regions where the field variables are sufficiently smooth, the integral balance or conservation equations may be re- placed by local differential equations. The conservation of mass condition leads to the differentiad equation usually called the equation of continuity, api ss— + Y ‘(9191’ = 0 ' (1'1) where pi is the local density of component 1, Bi is the velocity of the local center of mass of component i, and it assumed that there are no sources of mass (i.e., chemi- cal reactions). The conditions of conservation of linear and angular momentum lead to an equation of motion for the fluid, 399 55— + V - (puu) = pX + V - c (1.2) where p is the total local density, r p E zpi , (1.3) i=1 and u is the local center of mass or convective velocity p.u. (1.4) ~ . 1 1~1 J. 'O :2 III II MB The quantity X in (1.2) is the external force per unit mass acting on the fluid system, and the tensor 0 is the stress tensor, the negative of the pressure tensor, which gives the surface force acting on a fluid element. Finally, conservation of energy is expressed by ———— + y - J = 0 , (1.5) where E is the total specific energy and J is the total T flux of the total energy. E It should be mentioned that the theory is already informationally deficient. Unless the external force, the stress tensor, and the total energy flux are independently specified there are more unknowns than equations. The pro— cedure in TIP is to provide phenomenological relations (sometimes called constituitive relations) for these quan- tities, although the origin of these phenomenological rela- tions is not determined with certainty. In some formula- tions of TIP a prescription is given for obtaining these phenomenological equations as will be seen later, but it is not clear that these prescriptions are more than arbi- trary. The develOpment given in this chapter approaches the determination of phenomenological relations in a more systematic if not more reasonable fashion. Another source of consternation is that TIP generally fails to consider the possibility of determining equations of motion for the individual components of the mixture. Clearly momentum is not conserved for each component of the mixture since there is exchange between components. This is not a disadvantage however, since it will be shown that equations of motion for the individual components can be formulated by the use of componentwise or partial stress 10 tensors without loss of information and with great increase in the sc0pe and the generality of the theory. This single aspect has led to many of the new results produced in this reformulation. At this point TIP makes a transition from the hydro- dynamic stage represented by equations (l.l), (1.2), and (1.5) to the thermodynamic stage. This is usually accom- plished by introducing the assumption of local equilibrium. For example Fitts (1962) uses the following postulate: For a system in which irreversible processes are taking place, all thermodynamic functions of state exist for each element of the system. These thermo- dynamic quantities for the nonequilibrium system are the same functions of the local state variables as the corresponding equilibrium thermodynamic quantities. The concepts defining the thermodynamic functions of state are not generally so restrictive as to exclude the possi- bility of extending the definitions to nonequilibrium situ- ations. It is another matter, however, to assume that the same functional relations exist outside of the domain for which they were established. Some attempt should be made to investigate the resonableness of this claim. Prigogine (1949) in fact made such an investigation and found that the postulate was reasonable for mixtures of monatomic gases sufficiently close to equilibrium. Little else is known about the validity of the postulate except by com— parison of the results which come from it with experiment. In this respect the postulate has performed very well for 11 systems whose gradients of the thermodynamic functions are small and evolve slowly. Despite these practical successes, there are still some inconsistencies which require clarification. One in particular was the stimulus for much of this research. According to the postulate of local equilibrium the spe- cific internal energy, E} is related to the specific entropy, S, by the equation of Gibbs; r + 2 u.d i=1 1 Di '5— r (1.6) d§=Td§-pd(%- where is the specific volume, “i are the specific 0 0 chemical potentials, and (7% are concentrations (mass fractions). With regard to the quantity p Fitts makes the statement: For a nonequilibrium system, the pressure p (in equa- tion (1.6)) is the same function p(p,E) as for an equilibrium system. In this case, p is not the pres- sure in the usual sense; i.e., p is not the normal force per unit are exerted by the fluid. In hydrodynamics the pressure or force per unit area is given in terms of the trace of the stress tensor. It seems strange that after stating a postulate to the con- trary, Fitts reaches the conclusion that the quantity p in equation (1.6) is no longer the pressure normally used while the temperature, chemical potentials, etc. all cor- respond to the concepts typically associated with them. Furthermore TIP in general fails to discuss the correspondence 12 between the quantity p and the trace of the stress tensor. This makes any critical analysis of the criterion for mechanical equilibrium within the context of TIP nearly impossible. At this juncture of the thermodynamic stage of TIP an entropy balance equation is deduced from the energy equation (1.6). It is always possible to write this bal- ance equation in the form So. . ¢ 32"“? :Ts‘r' ”-7) where Js is the entrepy flux and ¢ is the internal entr0py production. It is important to note that the separation of terms in this equation between a flux and source is fairly arbitrary. Furthermore the introduction of this equation to the list of transport equations does not in- crease its information content. The real need for this equation is for application of the second law of thermo- dynamics. The form of the second law most appropriate for nonequilibrium thermodynamics is Q 2 0 . (1.8) In TIP the entropy production and the inequality (1.8) play a far less important role than in the theory to be presented here. The inequality does serve to place restrictions on the phenomenological equations which are obtained according to the following ad hoc procedure: 13 After the energy equation (1.6), the equation of motion (1.2) and the equations of continuity (1.1) for the fluid have been introduced into the entropy balance equatiod and the resulting terms arbitrarily separated into flux and source terms, the entropy production can be_written as, r ¢=F:J+Z§-J,, (1.9) Where terms represented by the symbols F are called forces and those represented by J are called fluxes. This classi- fication into fluxes and forces is often arbitrary and it is impossible to show that any given assignment is correct. One of the more common assignments is F = (g + pl) . g = 2}} . (1.10) F0 = yzn T, J0 = - q, (1.11) if: u = - - Ei VT~ , Ji pi(ui u), (1.12) where Vu1=Vu-X +§VT (113) ~T i ~ 1 ~i i~ ’ ' and where the heat flux q is related to the total energy ~ flux of equation (1.5) by r q = J - DELI + 2J.fi . (1.14) 14 Also, Hi and Si are the partial specific enthalpy and entrOpy respectively. The linear relations for the fluxes of (1.10-13) are generally written (9 + pl) = (¢ - 2/3n) (Y°§)} + 2n sym(Yg) (1.15) where n is called the shear viscosity and ¢ is called the bulk or volume viscosity, r _ l - q — (zoown'r + 2 nojvTuj (1.16) 3—1 and r 1 + Ji = QiOVEnT + jgl flijvTuj 1=1,...r . (1.17) The coefficients Qij are called phenomenological or Onsager coefficients. The number of independent phenomenological coefficients is limited because of the linear dependence of the diffusion fluxes r 2 J. = o (1.18) and is further restricted by the entropy inequality (1.8). See Appendix A for a derivation of these restrictions. It should be made clear that the phenomenological expressions given here are not unique and that TIP generally considers many transformations among the forces and fluxes. The pro- cedure is quite constant however: the phenomenological 15 relations are always determined in terms of those particu- lar quantities appearing in the expression for the entrOpy production. It is certainly obvious that phenomenological relations must be used in order to make the theory deter- ministic and that it is convenient to write out the source terms of the entropy balance equation in order to make use of the second law inequality (1.8) for the entropy pro- duction. It is not so obvious, however, that the phenome- nological relations are to be written only for those quan- tities called fluxes in the expression for the entropy production or that they are only determined by linear combinations of only those forces of the same tensorial order. For example, TIP suggests that the same expression used for the stress tensor in a pure fluid is also appro- priate for a mixture of fluids. Furthermore it is not obvious that the heat flux should be independent of the velocities of the components. The motivation for this ad Egg procedure for obtaining phenomenological eXpressions in TIP needs to be examined and at the same time the pos- sibility for making the system deterministic by other procedures should be considered. The theory of mixtures to be developed in this_ chapter uses the thermodynamic methods of Coleman and Noll (1963) and Coleman and Mizel (1963, 1964). This approach combines the rational use of the entropy inequality (1.8) 16 and the principle of equipresence (An independent variable present in one phenomenological equation for a material should be so present in all, unless its presence is in direct contradiction with the assumed symmetry of the material or the laws of thermodynamics.) to obtain a sys- tem of transport equations consistent with ordinary thermo- statics but without the need for the second part of the assumption of local equilibrium. This approach to non- equilibrium thermodynamics, after careful extension to mixtures, eliminates the inconsistencies and removes most of the criticisms mentioned above. Coleman, Noll, and Mizel considered single thermo- elastic materials, and the extension of their methods has been attempted by only a few workers. Green and Naghdi (1965, 1967), Ingram and Erigen (1967), and Bowen (1967) have made such an extension, but have failed to produce a theory which is both well motivated physically and also consistent with classical thermostatics. Stimulated by the shortcomings of these authors, Mfiller (1968) succeeded in formulating a thermodynamic theory of mixtures of fluids which is consistent with both thermostatics and TIP in those circumstances in which they can be compared. Mfiller's major contribution was to show that the form of the entropy production could be deduced from the principles set forth by Coleman and Noll. The theory of this chapter extends Mfiller's work in such a way that the comparison to TIP is 17 easily made and so as to make the resolution of the incon- sistencies mentioned above most obvious. The development of the theory will be along the following lines: The conservation laws are introduced in the standard fashion to obtain the equations of continuity for each species, the energy balance equation, and in dis- tinction with TIP, equations of motion for each component in terms of partial stress tensors. The entropy inequality and balance equation for entropy are introduced next. The energy balance, equations of continuity and motion, and the entropy balance equation are all used to write the entrOpy inequality in great detail. The (Helmholtz) free energy is introduced as the energy variable at this stage of the formulation for convenience in writing the results. In order to make the information content explicit (with respect to number of unknowns versus number of equations) the concept of thermodynamic process is introduced at this point. It is obvious that the theory at this stage can not be developed further without phenomenological assump- tions about the fluid. Having introduced particular con- stitutive assumptions of an extremely general form (in particular not linear) the consequences of the entropy inequality are deduced. One result for this particular mixture is the establishment of the validity of the thermo- static functional relations applied in nonequilibrium situ— ations. Further results are obtained by restricting the 18 class of constitutive relations to those linear in the independent variables. Such a linear fluid is called an "ordinary" fluid mixture. After a considerable amount of algebra and careful application of the entrOpy inequality, it is found that the equations of transport for the "ordinary" fluid mixture are surprisingly similar to those generally found in TIP. There are differences how- ever; for example, the diffusion fluxes are determined from equations of motion rather than directly from phe- nomenological equations. There are some differences due to the fact that terms involving the kinetic energy of diffusion are included here but are usually neglected in TIP (although not necessarily). Finally a complete review of the problems brought out in this introduction is made to show that a better understanding of the nature of non- equilibrium thermodynamics for mixtures has been achieved. 2. Kinematiggand ‘ Equations of Balance1 In this section the kinematics of motion and the axioms of balance of mass, linear momentum, moment of momentum, and energy for a mixture of v components are considered. Nothing new is presented in this section; lCartesian tensor notation will be employed through- out this chapter. Vectors are denoted by a single subscript, vi, while tensors are doubly subscripted, tij. The summa- tion convention on indices will also be used i.e., 19 reference could simply be made to the papers of Mfiller (1968), Bowen (1967), or Truesdell (1957, 1960) were it not for the fact that notation and some of the well-known relations need to be introduced to the reader. Accord- ingly, the derivation of the balance laws is only sketched here. Every place in space, xn, is assumed to be occupied simultaneously by "particles" of all v components. The density of component a is denoted pa, and v: is the com- ponent velocity relative to some frame of reference. The total density of the mixture is defined as2 V p = {p (2.1) ov. = X o v9 (2.2) OI The diffusion velocity of component u: is defined as u. = v. - v. (2.3) 2For the remainder of this chapter the summation v over all species 2 will be denoted 2 . o=l o 20 The material, substantial, or convected time derivative, gL , i.e., the time rate of change an observer would detect if he moved through the mixture with velocity Vi, is re- lated to the time derivative at a fixed point, §%», by d_ a a as- ss” Vi Iii ”-4) Thegeneral balance equation.--The balance or con- servation laws will be postulated as integral equations. In regions where the field variables are sufficiently smooth the integral balance equations may be replaced by local differential equations. The integral equations of all the balance laws considered in this chapter are spe- cial cases of the general balance law d EE_]- pawadv = - jr ngsi + Jr paoadV , (2.5) V S V where V is a fixed volume with bounding surface S, F: is the influx of we through S, Go is a volume supply of a within V, and we can be any component of a tensor, vector, or scalar. a 8p 31) as. avo.‘ In regions where p w v? o a 0 «——£ ——£ are o’ a’ 1'5t '5t ’ o'ax.'3x. continuous, (2.5) may be replaced by the differential equation 3 8 a SE 9 W + 3x. (powovj a _ a C1 3 + fj) _ p O I (206) 21 where the convective contribution to the influx has been accounted for separately according to a _ a a By summing the integral equation of balance for an individual species over all species the total equation of balance is obtained, “‘f ——- pde = - If F.dS. + JI dev , (2.8) where the following definitions are used: ow = X puma (2.9) a _ a a Fi _ g (£1 + pawaui) (2.10) pS = 2 pa a - (2.11) a The differential equation associated with (2.8) is aplp+-§—F=OS . (212) RE axj j ’ .' Since appropriate Fj and 8 may always be defined for any given w, it may be said that all quantities w may be balanced, by definition. There remains, however, some indeterminacy in the definitions of the influx and volume 22 source since any arbitrary solenoidal field may be included in the influx without affecting the balance equation. Even so the general balance equation is still useful since in many cases the influx or source may be given a priori, or special forms for these quantities may be derived from in- formation about w. The conservation of mass.--In the absence of chem- ical reactions the mass of an individual component is assumed to be conserved. Thus the equation of balance of mass of species a is postulated to be d _ _ a _ This is a special case of (2.5) with w=1,f‘?‘=o,o=o. (2.14) Substitution of (2.14) into (2.6) gives so a 3 a _ _ W + K]. {)an — O (a-l,...\)) o (2015) Summation of (2.15) over all species results in apv. Go ;I_ 75_t+ 3X _ 0 . (2.16) 23 An alternative form of (2.15) can be obtained by intro- ducing the rate of change with respect to the barycentric velocity. Using (2.4), (2.15) may be written in the form dpa a apa a av? H'E-I-+ 11). 5357+ p #= 0 (0:1,...V) J J . (2.17) The conservation of linear momentum.--The equation of balance of linear momentum of species a is postulated to be 3—.j-pav.:dV'=I-j;(pav:v;.-Sij)dsj -+j;(pm:+pab.:)dV(o=l,...v). (2.18) This is a special case of (2.5) with II I m o o II a a pmi + pobi . (2.19) aj)= pmi H+p b: (a=l ,...v) (2.20) The source term here has been split into two parts. The external body force density pub: gives the part of the linear momentum source which would be calculated if the fluid consisted of only component a. The remainder, pmg, represents a force density due to interaction of component a with the other components. 89 lj is the partial stress tensor of component o. 24 Clearly the total linear momentum should not be affected by internal interactions; thus, 2 m: = o . (2.21) a Summation of (2.20) over all species yields by use of Z pau. = o (2.22) and (2.21) the following equation expressing the balance of total linear momentum: a a 3L t 3x. 3 j a 1 j i o. O. The quantity in brackets corresponds to the total stress tensor for the fluid as a whole: a Si. - p u.u. = Z pab. . (2.23) T.. = Z ( a - pugug) . (2.24) s.. 13 a 13 The conservation of moment of momentum.--The re- sults needed here are completely embodied in Truesdell's (1960, p. 546) derivation of Cauchy's second law of motion, expressed by When there are no assigned couples and no couple- stresses, a necessary and sufficient condition for the balance of moment of momentum in a body where linear momentum is balanced is =T.. ; (2.25) i.e., the stress tensor is symmetric. 25 Hence from (2.24) one concludes that the sum of partial stresses is also symmetric, although the partial stresses themselves need not be. The conservation of total energy.--The balance equation of total energy is postulated to be u .ijj 3% f pETdV = - f(q -v..T .+E ij)dS. J+f(pr+2pa VzbaMiV, (2. 25) V S where ET is the total energy density, qj is the heat flux vector, and r is the heat supply density. Equation (2.6) becomes _ o p E- + 5';- (qj - ViTij) - pr 'l' gpavgbi . (2.27) The internal energy is defined as the difference between the total and kinetic energy: E = E - ¥L 2 p v v (2 28) 2p a o o O C A balance equation for the kinetic energy is obtained by using equations (2.15), (2.20), (2.22) and (2.16): d l a a _ 8 o o _ a o 0 5E (53 g pavlvi) — Egj [g vi(S13 pauiuj) o l a o o a 1 + 2 g paululuj] - g 1] ij + p g m u + z p Vibe (2.29) OI &*4 ,, 26 Subtraction of (2.29) from (2.27) with the use of (2.28) yields a balance equation for the internal energy: dEI ahi a a a 3v: pa-t.—+.5_£3=pr-p§miui+gsijfi—j-, (2.30) where a new heat flux vector has been defined, a and l a a a -pau.uj)+§29 u.u.u. . (2.31) (S 1 a 1 1 3 330,113 3: The Clausius- Duhem Inequality_ The mathematical statement of the second law of thermodynamics takes the form of the Clausius-Duhem in- equality or the entropy inequality. The form of the in- equality used by Coleman and N011 for single materials is not sufficiently general. Even the various entropy inequalities prOposed in works of TIP include an influx of entropy different from that for single materials. Perhaps some of the shortcomings of the earlier theories of Green and Naghdi, Ingram and Eringen, and Bowen are' due to their particular choices for the form of the entropy production. Mfiller (1967, 1968), in fact, showed that the form of the entropy production could be derived explicitly within the domains of the principles set forth by Coleman and Noll. 27 According to the remarks made in the previous sec- tion any quantity can be balanced. Thus, the balance of entropy will be postulated to be 6 f pndV = - Jf (¢.+pnv.)dS.+.]— where on is the entropy density, 8 is the temperature, y gym;— dV , (3.1) is the entrepy production due to dissipative effects in V, and ¢j is the entropy flux through the surface. The main point in this formulation is that, in contrast to all previous work, the entrOpy flux is not to be specified in advance. The entropy flux, sj, is to be regarded as a quantity for which a constitutive relation has to be formulated, just as is usually done with the heat flux and the stress tensor. From (3.1) and (2.4), the result corresponding to (2.6) is dn —._1 £ 9 as" ax.“’°+pY+p ° (3'2) 6 J J The second law of thermodynamics is made mathe- matically precise by the following postulate: y 2 0 . (3.3) This postulate places restrictions on constitutive equa- tions of the type to be introduced in the next section. 28 It will be convenient for examining these restrictions to write the entrOpy production, 7, in detail using the follow- ing definitions and substitutions: The (Helmholtz) free energy is defined by wI = EI - 0n . (3.4) The difference between the entropy flux for mixtures and that normally used for single materials (viz., the heat flux divided by temperature) is defined by k. = ¢. - —i . (3.5) Substitution of (3.4) and (3.5) into (3.2) yields, upon rearrangement, as ’ 3h. aw 3k. h. _ I j _ _ I _ d6 j__ j 30 pay - p dt + Bx), pr pat pn dt+eaxj 8 ij ° (3'6) The equation of balance of internal energy (2.30) is inserted into (3.6) to give aw 3k. h. 8v? 3 xj D“ a Xj The partial stresses can now be decomposed into symmetric and antisymmetric tensors according to 89. = t8. + :9. (3.8) l] 13 1] ta. = i (8“ + 8“.) (3 9) ij 2 ij j1 ' 29 T.. = % (Sq - a ) 13 1j ji , . (3.10) which allows (3.7) to be written d4) 3k. h I de 36 a °9Y=“°ar—‘p”ar+esrl"aisi‘olmju§ 3' j a + 2 tijd ij + g Tijw ij . (3.11) In (3.11) the two quantities 3v: av? (3.12) _ l dij _ 2' 5x. + x. j 1 and Bug Bug 8p So a _ l 1 1 B B_ B B have been introduced with the help of the relation a .— 2 Ti. — o , (3.14) which is a result of the symmetry of the total stress tensor. Equation (3.11) now gives the entropy production in a convenient form for use in the entropy inequality (3.3). 30 4. Thermodynamic Processes By a generalization of the definition of Coleman and N011 (1963), a thermodynamic process for a mixture is defined as follows: A process is a thermodynamic process if it can be described by a set of (5v + 6) functions v“ = v°‘ (x t) E - E (x t) ' - (x t) 1 1 1' I ‘ I 1' n " n 1' ma = ma (x t) 6 - 8 (x t) <1) - (x t) i. i i’ ' i’ i- j. i’ a - a - for each a, o=1,...v, which satisfy a) the balance of mass for each component (p v.) = 0 (o=l,...v) (2.15) b) the balance of linear momentum for each component 0. 3%":- 1 a _ o o _ -Sij) — pmi + pobi (o—1,..An (2.20) c) the balance of angular momentum for the mixture 2 ri‘j = o (3.14) 31 d) the balance of energy for the mixture dEI ahj a a a 3v: p E?- + axj = pr - o g mini g Sij 3;; (2.30) for all xi and t. To specify a thermodynamic process it suffices to prescribe the (4v + 5) functions pa,v:ml H,S 1j’ ,6,hi,n,¢i . The remaining (v + 1) functions bi and r are then determined from equations (2.20) and (2.30). In the above definition no reference was made to the entropy inequality (3.3). Its part in the development will become clear in section 6. 5. Constitutive Assm mptions £6} Fluid Mixture To illustrate the implications of the theory pro- posed in sections 2, 3, and 4, it is useful to consider a special mixture, namely, a mixture of elastic materials susceptible of diffusion and heat conduction. This case is sufficiently generaly to illustrate the important fea— tures of the theory, and it is important since it includes by further specialization the thermodynamic theory of single materials and the theory of TIP for mixtures with diffusion and heat conduction. The equations of balance and the entropy inequality suggest that constitutive equations should be prescribed a o ,t ,T..,m. for each o,(o=l,...v). It is for wI,n, hi ,k. ij 1] 1 i assumed that the history of the motions of the components 32 and the history of the temperature within the body con- sidered determine these quantities as functions of xi and t. The functional relations that connect these functions with the histories of motion and temperature are called constitutive equations; their form characterizes a material. If, in particular, the constitutive functionals re- duce to ordinary function of So 36 pslfiglelfi.lugldgjlw§j (5.1) J 3 then for the purposes of this chapter, the material will be called a mixture of fluids. The constitutive relations are then - B uY 8 Y )IJI - ADI (pBIfiI Igi ui (dij twij) (5.2) = B uY B Y n n (p8,fi ,0,gi ui 'dij'wij) (5.3) _ B Y B Y kj - kj (OB'fi’e’gi’ui’dij’wij) (5.4) G. _ OI. . f8 uY B (DY = tij "' tij (DB'f i'e’gi, 'di 1]., (003-) (O. l'ooov) (5.5) o _ a B B wY _ _ Tij - Tij (p8,fi,6,gi,u ,di lj' %j) (o-l,...v 1) (5.6) a _ a B Y B wY _ _ mi - mi (pB,fi,e, g J_,ui, ,di 1j’ j) (o—l,...v l) (5.7) =q (p £889:i quB., 81(3) (58) qi 1 Bl iI I I I lj' 0 where Bo f] = 3x? ’ g] = %E. J J and 33 These constitutive equations satisfy the principle of equipresence, according to which the same independent variables should appear in all constitutive equations un- less their presence contradicts the entropy inequality (3.3) or the assumed symmetry of the material. It should also be noted that the functions (5.2)-(5.8) depend only on objective combinations of the velocities and velocity gradients. This choice was made because the principle of material objectivity (Noll, 1958) restricts any description of a material property of a material to one which does not depend on the frame of reference of the observer, no matter how he moves. An admissible thermodynamic process is defined as a thermodynamic process that is compatible with the con— stitutive relations (5.2)—(5.8). To every choice of the component densities pa, component velocities v?, and tem- perature 6 consistent with the equations of balance of mass of the components (2.15) and with 6 > 0, there cor- responds a unique admissible thermodynamic process. For when pa,vg and 6 are known for all xi and t, then clearly a ij equations then determine wI,n,ki,qi,m:,t: f:,g3i,u:,d , and ng are determined. The constitutive a and Ti Once J" 3" these fields are known, then b? and r are determined by the balance equations. Thus, at a given time t, it is possible to specify arbitrarily not only pa,vg, and 6 (subject to (2.15) and 34 de dg dug dd?. 6 > 0) but also the derivatives 3E , dt& , dE_ "—dEl I dw.. 3g. 1 i . .331" and gig-at a pOint xi and to be sure that there exists at least one admissible thermodynamic process cor- 1 It is not necessarily true, responding to this choice. however, that for this choice the entropy inequality will be satisfied. Extending the idea of Coleman and Noll (1963), it will be required here that the entropy inequality (3.3) be satisfied for all admissible thermodynamic pro- cesses. In other words, the constitutive functions are subjected to the requirement that the entropy production be non-negative, identically in the independent variables. The next section will be concerned with deriving necessary and sufficient conditions that the constitutive functions must obey in order that the entropy inequality be satis- fied for all admissible processes. g. Consequences of the Entropy Inequality According to the discussion in the previous sec- tion, it is necessary to find the restrictions imposed on the constitutive relations by the entrOpy inequality. The non-negativity requirement is most easily applied if the entropy inequality is written in terms of the independent quantities discussed in the previous section. From the assumed smoothness of pa,v:, and 6, it follows that 1It should be noted that doc and dfg are not in- dt dt dependent because of the relations (2.15). This fact will be used explicitly in the next section. 35 a - p ail -Zp (2::d .32 p .igl £1 d dt 8f? dt 1 awI dg. . awI dug 9 (as)% W( T)dti p 3ua '3?- i a d. awI ddii . awI dwl - Z 0 dt - 2 p ——5 dt (6.1) a 3d? a 3w.. 13' A 13 and 3kg 3kg a 3k2 a 3k 6 Si‘ — 2 e 33— fg + 2 e «at12 + e 56"9z 2 a a 3fi 8k . ak an? *9 6—2)912+59 ‘3‘ 3x—1 gi a Bug 2 3k2 3d: . 3k 369. + 2 e .3 + 2 e ——— ——£l (6.2) a 3x a 3x a 3d.. 2 a 3w.. 2 13 13 where for convenience O. BZDO. fiz = 6.767.; (6'3) 2 _ a e 912 _ 3x axg (6'4) ' v-1 2 = X (6.5) a a=l If now the first and third terms of (3.11) are replaced by (6.1) and (6.2) respectively, and terms containing the same independent quantities are collected, then 36 3w 3w BwI du: I gI 39 dt and dt1 1 WI 0‘3 ' 34:1 —-1+e3k + er 6 a adOI awII 13 13 where . 31de -2. 1“: 6(3__..2)_.,] Sp t a 9 '5— g, a a a afi 3k 3k 3k an? + 2 6 5_&, f: + Z 6 ——£ f“, 'e'——& §_£ a pa 8f? 1 Eu? xfi i i a a 3kg adlj . 3kg awlj + 6 3x + e a 3x a aa?. 2 0 3w.. 2 13 13 a a a a a a a Since each of the quantities a a a d6 dgi dui ddi' dwi. (6 8) EE ' EE‘ ' dt ' dt ' dt ' 912 can be chosen arbitrarily and independently of any other term in the inequality, the last with due respect to its symmetry, the argument given at the end of the last sec- tion implies that the necessary conditions for non-negative entrOpy production are: oeY = oevred 2 0 (6.9) 37 awI n = - (55—) (6.10) BWI 53E._ 0 (6.11) 3¢I Bui BWI — = O (6.13) (“3) 3w 0 I (6 l4 (XE-j "' o ) 3k 8k. 1 i _ L55; 4- 531- - 0 . (6.15) The inequality (6.9) can be further reduced, but the results will not be needed until section 7. Equation (6.10) is a well-known result of thermo- statics, and its presence here means that it is also valid in non-equilibrium thermodynamics for a fluid mixture. Equations (6.1l)-(6.14) imply that the functional dependence ofdkEmay be written as (1 WI = LI(pa'e’fi) 0:1,...V (6.16) In the next section, in which a linear theory is introduced, the free energy depends only on pa and 6. This together with some results established by reference to the equilibrium 38 state, yet to be defined, delineates precisely the domains of validity of the theory of nonequilibrium thermodynamics derived from the assumption of local equilibrium. 7. Ordinary Binarprluid Mixtures (The Linear Theory) Up to this point mixtures of v components have been considered. However, subsequent formulas would be even longer and more complicated than they are anyway if more than two components were considered. Since there is no conceptual difficulty in extending the theory to a mixture of any finite number of components, the treatment will be restricted, without loss of generality, to binary systems. The development of the theory is further restricted in this section to the special case of mixtures whose con- stitutive relations are linear in the variables Bo 3p 1 2 36 1 1 2 1 FT?" ’5— 73—; ' “2 ' dij ' dij ' wij (7'1) This special material will be called an ordinary binary fluid mixture. That such a material is special is obvious from the restrictions placed on the constitutive relations. A conviction that this material accurately represents the binary fluid mixtures commonly encountered by the chemist in his laboratory is the reason for the label "ordinary." The appropriateness of this label can only be determined by testing the ability of the constitutive relations to 39 represent accurately the experimentally measured responses of the material to variations of the independent variables. It should be mentioned, however, that under special condi- tions, according to which it is possible to neglect the effects of the so called inertial and viscous terms in the equations of motion, the ordinary binary fluid mixture is precisely the same material that TIP claims to describe. This correspondence will be made explicit at the end of this section. Material objectivity and linear representation theorems.--So far the principle of material objectivity has not been used to its full extent; material objectivity has been used only in the selection of independent vari- ables. A further consequence of the principle of material objectivity is that the constitutive functions must be isotropic scalar, vector, or tensor functions. Repre- sentation theorems have been worked out for these iso— tropic functions (see Truesdell and Noll (1965) and Smith (1965)). These theorems will not be needed here in their full generality, but for the case of functions linear in the variables (7.1) the following relations must hold for scalars s, vectors wi, and tensors tij: 40 B B l _ _ S (DBIfi 16:9 lluill Id lj’ wij) - 5(0816) B - lI2 f8 8 l B l wi (DB'f 2'9'92'u2'd2j ”(3') = .ngi + 821‘” p6f£+wn “1 2 3 f8 1 B l _ _ B d8 tij (DB! flierggruirdgjrwzj) P51] + le¢ dnnsij 2 + 2 2n8 {dB } B=l a B B l _ _ 1 tij (ps,f£,e,g2,ug,d£j, wzj) — “wij (7.2) The superscripts s and a on tij indicate the symmetric and antisymmetric parts of tij respectively. The coef— ficients WT, wa,wD, p, ¢, n, and 11 may depend on pa and 8; {dij} is the traceless part of d? ij' The functional relations given in (7.2) are usually obtained in TIP by application of Curie's theorem: enti- ties whose tensorial characters differ by an odd integer cannot interact in isotropic systems.l Interactions be- tween quantities of different tensorial orders are well known in the kinetic theory of gases. The separation of effects described by the writers on TIP are simply those that follow by linearization of isotropic functions 1This theorem has been attributed to P. Curie (1894), although apparently he neither stated it in this form nor proved it. 41 and should be considered only within the context of the linear theory. Constitutive relations in the linear theory.--In section 6 some general restrictive conditions on the con- stitutive relations (5.2)-(5.8) were obtained. These re- sults are summarized in equations (6.9)-(6.15), and they will be applied now to the most general possible linear constitutive equations. According to the representation theorems (7.2) the free energy in the linear theory does not depend on density gradients and thus for an ordinary binary fluid mixture (I = (I (91' 02. e) (7.3) Here, use of (6.ll)-(6.l4) has been made in leaving out the dependence on WI on f?, gi, ug, dgj' and ng. This means that for an ordinary fluid the free energy has the same functional dependence as that assumed in TIP on the basis of the assumption of local equilibrium. According to (7.2), the heat flux hi in the linear theory must be a linear function of the vectorial quan- tities in (7.1). Hence __ _ 1_ 2_ 1 hi — ATgi lei szi Quj (7.4) where the coefficients may depend on pl, 02, and 6. 42 Similarly, mi in the linear theory is given by 1 _ _ _ 1 _ 2 _ 1 mi - mogi m fi m fi mDu. l 2 l , (7.5) where again the coefficients may depend on p1, p2, and 6. A linear constitutive relation must also be for- mulated for the vectorial quantity kj; hence, from (7.2) and the restriction (6.15) already found for ki, the form _ l _ l _ 2 ki — + Kui lei lzfi (7.6) is expected. However, by further rearrangement of the reduced entropy inequality (this possibility was mentioned previously) it will be shown later that ki must be inde- pendent of density gradients also. Again from (7.2) it follows that the antisymmetric tensor Tij is proportional to w1.: lj l _ _ l Tij — uwij (7.7) where u may depend on p1, p2, and 6. The symmetric tensor tgj is given, according to (7.2), by ochB O. t..=—p(3..+ leqp dnndij 2 + 2 2n“8 {dEj} (6:1,2) (7.8) 43 a8 8 Here again the coefficients pa, ¢ , and nu are in general functions of pl, p2, and 9; {dzj} is the traceless part of dzj. Restrictions Imposed on the Linear Constitutive Relations by the Entropy Inequality.--In order that the entrOpy inequality be written in terms of independent quantities, the equations of balance of mass are intro- duced into the reduced entropy inequality in the form _ _ a a _ a aE—‘- uifi padijdij (7.9) At the same time several of the derivatives which are zero for the linear case are eliminated and terms involving the same independent quantities are collected. The resulting entropy inequality is from (6.7) e = IL 3w: f1 _ 1 361 f2 u1 + 6 3k2)_ fig 9 Yred 9 pl 53; 1 p2 3oz 1 D1 1 56‘ e 91 3k 3k + Z 6 g—& f: + 2 6 -—§ f: p(l mi+£ mi plui a pa a Bfi 2 pl 02 [3k. /3k. 3w + e -—% ml + e d). {p 3—5 pad9.6.. Bu. 1] Bu_ 1] a pa 1] 13 l l _ 2 Bo 8k da - 9 Bkfi £f1-£ f2 0 ul 6 p Bui 13 9 Eu: “5.1 D2 1 l 1 a a l 1 l 1 l + Z ijdlj (31 ij + —2le plwij 2 0 (7.10) 44 where the identity Bug d l B l B 5;? = “in + dzg— 3 g psdiz - 3 g uEfz (7.11) has been used. The quantities fgg are independent of all other terms in the inequality (taking proper account of their symmetry) and in the same manner as in section 6 the following restriction is needed: Bkz ki -—a- + —a- = 0 (7.12) Bfi Bfg This implies that 21 and 22 are both zero and thus the most general possible linear constitutive relation for ki is k. = Kn: (7.13) Equilibrium properties.—-Further conclusions may be drawn from the inequality (7.10) without assuming more particular constitutive equations. These conclusions refer to the equilibrium state, and they establish a consistency between this theory and thermostatics. The material is said to be in equilibrium if the independent variables , d.., d.., w.., u?) A=l,...21 (7.14) 45 are all zero. In order to write the entropy inequality completely in terms of these quantities the linear rela- tions are inserted into (7.10) to yield AT Cl 1 C2 2 _ l l peYred . If'gzgz + TT'ngQ + 7§1g2f£ + p — + — m p u 01 92) D 1 1 1 1 61+ 32)mo]°1“292 3W p ( I 6 (3K ) (1 1 ) 6K ] 1 1 + _ + — —— +;)—1 + — m - —— p u f [91 501 91 801 91 02 l 091 1 2 2 + B if; 6 25. + i + l m + 25 ulfz pol oz 2 002 D1 2 l + Z 2 ¢“Bd“ d8 + 2 2 2n"‘8 {dgj} {afij} C1 0. 8) 1 l u l l I _ l l + '61 + ‘2) 61 “’1:- ij +[ o + 901(7)}: P )du 8w er ] I 2 2 2 + _ - “— - d 7. 5 [902(3p2) p p 22. ( l ) The entropy production y may thus be considered as a function of the variables (7.14), and from inspection of (7.15) it is seen that Y (X1,X2,...X21) 2 O (7.16) and y (0,0,....O) = O . (7.17) 46 Hence y has a minimum value in equilibrium. A necessary condition for this is (g?) = o (7.18) A 0 where the subscript 0 indicates that the function to which it is attached is to be evaluated in equilibrium. Application of (7.18)to (7.15) results in the fol- lowing restrictions: Bw GKp l I 2 P = CD ‘—-'+ (7.19) l 391 0 8w 6K9 2 I 2 P = 992(§E—)- —Er— (7.20) If the pressure is defined as then from (7.19) and (7.20) aw I p = p 2 pa 53— (7.22) a a Eq. (7.22) is the usual definition of pressure in thermo- statics, and thus provides another example of an equilib— rium relation that carries over to nonequilibrium processes in an ordinary binary fluid mixture without need for local equilibrium assumptions. 47 If the chemical potential is defined as1 BwI “a ‘ ”’1 + p “—30 (7.23) a then the relation 2 Dana = pr + p (7.24) O. is easily obtained, from which the Gibbs-Duhem relation Zpa—3=-a£.-pn 39- (7.25) follows by use of (7.3). The results of the theory given here for the ordinary binary fluid mixture are thus completely consistent with the thermostatic theory and all the formalism of the latter can be taken over for use in nonequilibrium situations. There are still further consequences of non- negativeness for entropy production. If, for example, all variables in (7.14) are chosen to be zero except for 1 wij' then it is necessary that u 2 O (7.25) In the same way, if all variables except dgn are chosen to be zero then 1The symbol Ha used here for the chemical potential should not be confused with the symbol u used previously for the phenomeological coefficient for the antisymmetric part of the stress tensor in equation (7.7). 48 12 21 11 22 ll 22 1 ¢ 2 0 : ¢ 2 0 , ¢ ¢ - 4 (¢ +¢ )3 0 (7.26) One also finds using the same technique U11 2 0 ' r12 2 0 ' r111r122 __ %_ (n12+n21)3 o (7.27) If now the temperature gradient 92 is chosen to be some arbitrary vector Xi' f: and f: to some other arbitrary vector Yi’ and all other independent are both chosen to be equal variables are taken to be zero, then it can be seen by com- pleting the square that in order for the entropy production to remain non-negative for all choices of Xi and Yi it is necessary that A 2 0 (7.28) and C + C = 0 (7.29) Furthermore, if the density gradient ft is chosen to be equal to some arbitrary vector Xi' gfl and f: are both chosen to be equal to some other arbitrary vector Yi' and all other independent variables are taken to be zero, then by completing the square, it is necessary that C = 0 . (7.30) Eqs. (7.29) and (7.30) imply further C = O . (7.31) 49 Precisely the same technique used with the diffusive velocity ufi, instead of the temperature gradient, and the two density gradients yields the three necessary conditions 2 p 8 8K 1 1 8K ——— m + — + ——-p - —— = o (7.32) 0192 1 91 591 02 Q2 1 1 _E_z_m+§.§§.-.}_p2=o (733) p192 2 p1 3p2 p3 mD 2 o (7.34) Finally if all the independent variables except for the temperature gradient 92 and diffusive velocity u: taken to be zero, the non-negativeness of the entropy pro- are duction requires 9 meD 1 8K Q p z 0 . (7.35) —T‘z 9— 6" Summary of the linear constitutive relations.--A summary of the most general linear relations possible for an ordinary binary fluid mixture are h = - x g - Qul (7 36) i T i i ' l _ _ 1 _ 2 _ 1 mi — mOgi mlfi m2fi mDui (7.5) where 11 ll 0 \V \V \V \V \V Ku; 1 U ij - paé + E oaBdB 6. ij B= a=l,2 801 GKp2 + 001 g3; p BwI er2 pp '55;- o 9.11-2 (9.15.)- pi 01 3pl —.e_(3K+._l_.p 91532 03 0 0 1 3K Q 71'[6T)+6+ 0 22 \V (7.13) (7.7) (7.8) (7.19) (7.20) (7.32) (7.33) (7.28) (7.34) (7.35) (7.25) (7.26) (7.27) 51 These relations are necessary and sufficient for the entrOpy production always to remain non-negative for any admissible process in an ordinary binary fluid mixture. 8. The Transport Equations This section is intended as a summary of the dif- ferential equations governing the evolution of the state of the ordinary binary fluid mixture in thermodynamic processes. No new results are derived here. The equa- tions are rearranged and expressed in a form more convenient for applications. This includes replacing p1 and p2 in favor of W1 and O as independent variables, where the mass fraction w1 is given by wl = 01/0 (8.1) The balance of mass.--The two independent equations of conservation of mass to be used here are do _ _ l HIE" p — (8.2) which follows from (2.16) upon use of (2.4), and .1 dw Bji __1==_ ___ p t 8x. d (8.3) 1 which follows from (2.15) upon use of (2.4) and the defini— tion of the diffusion flux jj== u) (8.4) 52 The balance of momentum.--The overall equation of motion of the mixture is derived from equation (2.23): dv p(V'-V ) 3w 3n.. 1 _ a _ B 36 _ 1 8p _ l 2 l ij p at - 2 pubi. 8' 5x. 98' 5x. 8' 8x. + x. (8‘5) a 1 1 1 j and the following thermodynamic identity has been used for the pressure gradient: _§£ ____ B 88_1 a _ p(Vl-VZ) awl (8 7) 3x. FT'fiil 68‘ x. 8' 3x. ' ° 1. l l J. where 8 is the thermal expansivity, B=—.];§.B. 9 3T le1 8' is the isothermal compressibility, 811%. p 3p T,w and Va is the partial specific volume. The equation of balance of linear momentum (2.20) is written more conveniently (8.6) (8.8) (8.9) 613'1 p i _ D .1 38 3 - —— 1+ D §——+— (u u) l 8 l l l 3 2 2 -[§;3—x;(s +P513"6;5§;‘51j+1° 13’] J 3 9192 3"j p2 5"j ) + .1.1 1 8p .1 (9(92'91)+°192 awi); (8 10) Jijj EIEE'EET’ 9\ 2 2 8§.y ' where the coefficients D1 and D2 are related to mo, mD, and K by psz D1 = 2 (8.11) D192 2 p m _ 0 K 8 8K — _-— D2 - T— + 6- + B—(fi <8]. 52) (8.12) 0102 1 l The partial specific entropies, 85, have been used in (8.12). The special derivative of chemical potential used in (8.10) is defined by — _ .— -— 2 E3(V1"’2) 88 +[0(V1‘V2’ pull] 8“1 l 3 _ axi (“l-“2)T _ 8' 3x. 8' + p2 Exi + (Vl-VZ) 39 (8 13) 08' 3x. ° 1 where U = iii (8 14) 11 8x1 8,p 54 The heat flux.--A new heat flux is introduced by the definition — (fi'l - HZ) j? (8.15) Q- = Q: J 3 where Ha is partial specific enthalpy defined by - _ — l a H - 65a + “a + f uiu. (8.16) a 0'. 1 From the definition (2.31) of the heat flux hi and the constitutive relation (7.36), qj is seen to be 1 _ 80 .l a a a- - qJ - Bl axj + BZJj + g ui (Sij + p Gij) (8.17) where the coefficients B2 are related to AT and Q by B1 = IT (8.18) and B = 0/8 + 6(5' - §') - 95 (8 19) 2 1 l 2 p:L I ' The temperature equation.--In section 5 it was determined that a thermodynamic process could be specified by the fields wl, p, ji, Vi' and 6. So far differential equations have been given for all of these except 8. The temperature equation can be obtained from the following thermodynamic identity: — de_ 211 8.8-9.3- [—_— _ QCV 376“ 99 dt+p8' dt (38(81 52) 8' 55 The equations of balance of mass (8.2) and (8.3) will be substituted into the last two terms of (8.23) and the first term will be replaced by the entropy balance dn _ _ 8 p6 3E — 5E3'6¢j + @jgj + pay + or (3.2) The entrOpy flux to be used in (3.2) is determined from the definition (3.5) of ki together with the constitu- tive relation for ki, the definition (2.31) of hi and the definition (8.18) of qi: _ l — _— .1 a a a e¢j — qj + 8(sl $2) jj + g ui (sij + p Gij) (8.21) Similarly, from equation (7.10) for the entropy production, the relations (2.20), (7.12), (7.23), (7.19), and (7.20) may be used to write the entropy production in the more convenient form: _ 8 .1 a a a a NY 7 ‘ 45-93- [5:]— (“1 112)] 3j " g ”‘1 65; “1 §(Sij + p dij) 5;; (8.22) Substitution of these last two results into (3.2) gives dt 8x 3 1 2) 8x. 33 + or J 3 8v. .1 3 —v —v 1 —Jj §;—-(Hl - H2) + Fl] 3X (8°23) 56 where the system has been restricted to external fields derivable from a potential a 39a bi = - 5?- (8.24) 1 and where —I _ a - Ha + “a (8.25) The final temperature equation is now completed by inserting (8.26), (8.2) and (8.3) into equation (8.23). The result is 5 d8 = _ a _ 65(V1‘V2) a .1 + r p v a’E _3xj qj 8' ij Jj 9 av. av. 86 1 _ .1 3 ' _ ' 1 ' 87'52; 31 6§;"H1 Hz’ + "ij IE; (8'26) Summary of transport equations.--The complete set of equations which are necessary to describe any admissible thermodynamic process in an ordinary binary fluid mixture are equations (8.2), (8.3), (8.5), (8.6), (8.10), (8.17), and (8.26) together with the apprOpriate boundary conditions. 9. Discussion It is not surprising that the transport equations for the ordinary fluid mixture bear striking resemblances to the transport equations of TIP (see Horne, 1966) since hopefully they describe the same material. Briefly, a comparison of the equations yields: (1) The equations of 57 continuity of mass, (8.2-3), are identical with those of TIP; (2) The temperature equation (8.23) (derived from the energy balance equation) is identical in form to that of TIP except that the possibility of external sources of radiation has been included; (3) The phenomenological equation for the heat flux (8.17).contains viscous terms and a kinetic energy of diffusion not found in TIP; (4) The overall equation of motion (8.5) is identical in form to that of TIP; (5) The viscous pressure tensor (8.6) appearing in the equation of motion contains a contribu— tion from the kinetic energy of diffusion; (7) The phenom- enology of the partial stress tensors (7.7-8) provides completely new expressions for the determination of stress; and (8) The equation for the diffusion flux (8.10), usually given as a phenomenological equation in TIP, is determined here from the componentwise equations of motion. If the inertial and viscous terms of (8.10) are neglected, the equation is identical with that of TIP. On the surface it appears that the major differ- ences between the two formulations arise because of the use of the partial stress tensors and componentwise equa- tions of motion. It should be apparent, however, that many of the underlying inconsistencies have also been eliminated. For example, the careful application of Cole— man's thermodynamic methods has eliminated the need for the local equilibrium assumption as it is commonly used. 58 An extension of Mfiller's work has made it possible to determine the entrOpy flux by derivation rather than by making an arbitrary separation of flux and source terms as is done in TIP. It has been shown that the entropy production may be written as a quadratic form which allows the second law entropy inequality to be used with maximum effectiveness. Finally, the equivalence of the pressure appearing in the Gibbisian equation for the energy and the negative trace of the stress tensor has been estab- lished. One result of the identification of pressure and the ability to write the entropy production in a quadratic form is that is is now possible, for the first time to give an unequivocal criterion for mechanical equilibrium: A system is in mechanical equilibrium if and only if These restrictions also imply that at mechanical equilibrium the external forces are completely balanced by the pressure gradient in the fluid. CHAPTER III PERTURBATION EXPANSION METHODS 1. Introduction The particular stimulus for the research in this thesis has been the need for a better understanding and interpretation of the results of pure thermal diffusion experiments. The previous chapter is the culmination of an effort to better understand the fundamental theory of irreversible processes in liquid mixtures. The following quote from Bellman (1964, p. 1) sets the tone for this chapter: . . ., one of the major problems confronting the mathematician, after he has passed the first hurdle of achieving a reasonable analytic formulation of a physi- cal process, is that of deriving useful and meaningful approximations to the solutions of the equations de- scribing the process. In some cases, analytic ingenu— ity, alone or abetted by digital computers, will furnish the desired eXpressions; in other cases, a combination of "low cunning" and physical intuition will provide the essential key. In still other cases, a completely new interpreta- tion and formulation of the physical situation is re- quired. A major obstacle at the start of any research is ignorance of where the real difficulties lie. The greater part of the time, only perseverance and plodding effort yield this vital information. Indeed, there is no reason to believe that the research reported here has been any different. 59 60 In this section, prior to a discussion of particular problems, the simplest and most useful of all approximation techniques will be presented: the expansion of a solution as a power series in a parameter--the classical perturba— tion technique upon which much of the edifice of science rests. The basic idea of the perturbation technique can be exhibited most easily in abstract terms. The introduc— tion by Bellman (1964) is probably the best starting point: Suppose that we are required to solve the equation N(u) = v (1.1) which for any of a variety of reasons, is inconvenient to tackle in its original form. It may, for example be non- linear, be linear but of high dimension, or possess variable coefficients. Let L(u) = V (1.2) be an auxiliary equation that possesses a useful explicit solution u = T(v) (1.3) in general, the unique solution of (1.2). In practice, this means that L(u) is a linear operation on u. Let us henceforth assume that L is linear. 61 Returning to (1.1), we write it in the form L(u) = V + R(u) (1.4) where the function R has been defined as R(u) = L(u) - N(u). (1.5) To facilitate our discussion of particular solutions of (1.4), we introduce a parameter s and consider the new equation L(u) = v + eR(u) (1.6) In some situations, the introduction of e is solely a mathematical artifice that permits us to do various types of "bookkeeping" in a systematic fashion. For example, it allows us to group terms of comparable degrees of approximation in a methodical and convenient fashion. In a large number of situations, however, this parameter occurs naturally, representing some physical quantity. In the application to pure thermal diffusion experiments we shall have use for both interpretations. It is quite natural and sensible to begin with the study of those equations where s is close to zero. Let us then look for a solution of (1.6) having the form u = u + eu + ezu + ..., (1.7) a power series in s with coefficients that are independent 62 of 8. Clearly, the leading term uo, obtained by setting 6 = 0, is a solution of the linear approximation L(uo) = v (1.8) a solution that we shall write as u0 = T(v). To obtain the subsequent coefficients, ul, uz, ..., in a systematic fashion, we substitute u as given-by (1.7) into (1.6) and equate terms, obtaining thereby + ezu + ...) = v + eR(u0 + an + ...) (1.9) L(u0 + eul 2 1 Since L is by assumption a linear Operator, the left side becomes 2 L(uo) + sL(u1) + e L(uz) + ... (1.10) Assuming that R(u) is analytic in u so that we can expand the right side of (1.9) as a power series in e, we have 2 ._ R(u0 + eul + e u2 + ...) — R(uo) + eRl(u0,ul) 2 + e R2(u0,ul,u2)+... (1.11) where, as indicated, the coefficient of uk depends only upon the quantities un, n s k. Combining the expansions of (1.10) and (1.11), and equating coefficients of e, the single equation of (1.6) gives rise to the infinite system of equations 63 L(uo) = v L(ul) = R(uo) L(uz) = Rl(uo, ul) (1.12) L(“1;“) = Rk(“o' “1' °'°' uk) and so on. The important point to observe is that this system of equations can be solved recursively; that is, the deter- mination of uk involves a knowledge of un, n c k. The first equation yields u0 = T(v). (1.13) From the second, we derive the relation u = T(R(u0)) = T(R(T(v))). (1.14) 1 Continuing in this way, we see that we can express each u solely in terms of v. k The infinite series in (1.7), whose coefficients are determined in the foregoing fashion is called a formal solu- tion of the original equation in (1.6). To obtain a formal solution of (1.4), we need only set 6 = 1. Pure thermal diffusion experiments.--Historically, the application of nonequilibrium thermodynamics to the problem of interpreting pure thermal diffusion experiments 64 has been made by considering the magnitude and variability of the terms in the transport equations. Among the assump- tions that have been made in formulating working equations in the past are: (l) constant phenomenological coefficients; (2) a one-dimensional temperature gradient; (3) a density independent of composition; (4) no convective motion; (5) no viscous effects; and (6) dilute solutions. Undoubtedly there is some validity to these approximations in some cases, and their use sometimes provides an adequate first order theory. However, it has been impossible to provide corrections to the theory for the cases in which there is significant deviation from these assumptions and it has thus been impossible to interpret properly the experimental results. The use of the perturbation technique not only provides these corrections but at the same time facilitates the solution to the transport equations. It will be seen from the description of the pure thermal diffusion experiment in the next section that two characteristics stand out as natural possibilities for the application of the perturbation method: (1) The phenome- nological coefficients are fairly constant and any varia- tion might be taken as a perturbation. (2) No steady state convection is expected if it is possible to minimize heat losses; convective corrections are therefore introduced by the use of a heat loss parameter. In fact, as will be seen in section 3, we employ a double perturbation expansion to account for both effects simultaneously. 65 2. The Experiment The experimental device used in studies of pure thermal diffusion is usually described as a sandwich-type cell consisting of two thermostated, horizontal metal plates with an arrangement for containing a homogeneous binary liquid solution of known initial composition between the plates. A temperature difference is maintained between the plates by the external thermostatting mechanism with the warmer plate situated above in order to minimize con- vective effects. The temperature gradients which are thus present in the liquid result in more than just heat con- duction. The two components in the mixture will tend to separate or demix under the influence of the temperature gradient (thermal diffusion). Of course the resulting con- centration gradient produced by this effect immediately induces ordinary diffusion which tends to remix the com- ponents. The total diffusion flux will be observed as the resultant sum of these two Opposing fluxes. The funda- mental measurement in the pure thermal diffusion experiment is the composition distribution in the cell under steady state conditions (all local time derivatives are zero) for which the thermal diffusion and ordinary diffusion fluxes are as closely balanced as possible (total diffusion flux approaching zero) and convective effects are at a minimum. The boundary conditions.--Ideally the boundary con- ditions to be used with the temperature equation would 66 consist of the temperatures at the top and bottom of the cell, together with a condition that the heat flux at the side walls is zero. Practically these boundary conditions cannot always be obtained and any observations of convec- tion in the cell are probably for this reason. For the purpose of this treatment it will be assumed that the temperatures of the top and bottom plate can be controlled as accurately as desired, but that there is a slight heat loss at the side walls. There are two principal reasons for the constant temperature choice. First, the primary objective of this chapter is to show how a perturbation theory can be formulated for the heat loss problem, and the variable temperature problem only complicates the situation. Second, there are many practical experiments that can be done for which the constant temperature assump— tion is quite good. In any case if the inclusion of vari— able temperature at the horizontal surfaces is required, the procedure will not be very much different from that which will be introduced for handling the heat loss prob- lem. With these considerations in mind the boundary con- ditions for the temperature equation are: 8 = Tm + 1/2 (AT) at z = h (2.1) 6 = Tm - 1/2 (AT) at z = 0 (2.2) and §_8_=€(T —8) atr=a (2.3) 67 where it has been assumed that the z axis is vertical, with the height of the cell given by h, and that the fluid is contained between the plates by a cylindrical surface of radius a. The difference between the temperature at the top and at the bottom has been denoted by (AT) while the average of these two temperatures has been denoted by Tm' The heat loss coefficient is given above as e, and the sys- tem is assumed to be radiating heat into surroundings which are at a temperature of Tr’ Finally, we take it as an assumption that all of the prOperties of the fluid are axially symmetric; 1121' they depend only on the vertical position in the cell and on the distance from the center. Although s is a parameter which depends on the apparatus and must bedetermined experimentally in each individual case, it is helpful to have an order of magni- tude estimate. If the fluid is assumed to be contained within glass walls (thermal conductivity = 0.0028 cal/(sec) (cm2)(°C/cm)) and to lose heat through a stagnant layer of air (thermal conductivity = 0.000053 cal/(sec)(cm2)(°C/cm)) of thickness 0.5 cm, then s is given by (0.000053)/(0.5 - 0.0028) = 0.038cm‘l. (See Carslaw and Jaeger (1959) for more details concerning the use of heat loss boundary con— ditions.) In the absence of chemical reactions, the boundary conditions for density and mass fraction are simply state- ments of the conservation of mass: 68 Zn h a j, j” .J- p(r,z) rdrdzdw = M (2.4) 0 0 0 and 2“ h a ‘[ —[ -f- p(r,z) W1 (r,z) rdrdzdw = M1 (2.5) 0 0 0 where M is the total mass contained in the cell and M1 is the total mass of component one. These two masses are easily obtained by weighing or density determinations at the beginning of the experiment. Boundary conditions for the velocity and diffusion flux are easily written to insure that no material either flows or diffuses through a rigid wall. These conditions are 2 jlz = 0 at z = o, h (2.6) ur = jlr = 0 at r = a (2.7) It will be further assumed that the fluid sticks to all rigid walls. This is the usual assumption in all hydro- dynamic treatments. This leads to the further conditions ur = 0 at z = 0, h (2.8) u2 = 0 at r = a (2.9) Under certain circumstances it might be possible that the number of boundary conditions is insufficient in order to determine properly all the solutions to the dif— ferential equations. If these circumstances arise, it will always be possible to ask the experimenter to provide the 69 additional measurements of concentration or concentration gradient at one or more points in the cell. Pure Thermal Diffusion 3. Tansport quatibnsl The equations to be used for the description of the pure thermal diffusion experiment are those listed at the end of Chapter II. There are a few modifications and specializations of these equations which make them more appropriate for pure thermal diffusion. The equations are written here in such a way as to make the comparison with the standard treatments of nonequilibrium thermodynamics convenient. In particular this involves the following major assumptions: (1) The viscous terms in both the phenomenological equation for the heat flux and the equa- tion of motion which gives the diffusion flux are neglected. (2) The viscous stress tensor is assumed to be given by the Newtonian formula; 1121! the stress depends only on the convective motion. The following simplifications have also been included: (1) Since the description is to be of asfieady phenomena, all local time derivatives are set equal to zero. (2) The only external force to be considered is gravity. (3) It is assumed that there are no external sources of radiation. 1For the remainder of this chapter direct, as dis- tinguished from Cartesian or component, tensor notation will be used. Vectors are denoted v, tensors by t, and the gradient operator by V = 8 ml 70 The steady state equations of transport, modified as above, are v ° Vp + pV - v = 0 ~ _JL_ . - ° _ p p Y Y21 + D1.3.1 + YT(“1 1 2 [p(vl-V2)gh pullo] + _ AT p2 pEvv ° VG + V . q' + 23 V - v - 2 :VV 88(V'-7') l 2 . + BT Y 0 21 + _ ' = ° 3 3139 + B2V21 (1:) Notice that the phenomenological coefficient D pv ° Vv + Vp - pgA - V -'n= 0 “2) V6 = 0 = (8 - 2/3n)(y '9) l + Zn sym (Y9) (3.1) (3.2) (3.3) (3.4) (3.5) (3.6) (3.7) of equation (II.8.10) has been eliminated in favor of the coefficient 0 which is defined by _ "2(Vi’vz)gh 02132 (AT) “11 pu11 (3.8) 71 The coefficient 0 will be called the thermal diffusion parameter. It is related to the thermal diffusion factor a used in many nonequilibrium thermodynamic studies of 1 thermal diffusion. For example, from the phenomenological equations given by Horne and Bearman (1967) it is easy to see that wlwzal = O _ "2(V1"V2’gh (3 9) 5 011(AT) ° The last term in this equation is typically very much smaller than 0, and the behaviors of 0 and al are there- fore very strongly correlated. Notice also that in equation (3.7) the_common coefficient of shear viscosity, n, and the coefficient of bulk or volume viscosity, 0, have been used. Finally, in the equation of motion (3.4) the exter- nal gravitational field has been assumed to be of magnitude g and to act in the direction 1. In all that follows, it is assumed that this direction is vertical and 1.: (0,0, - 1) (3.10) For completeness, the thermodynamic expansions for the pressure gradient and chemical potential gradient needed in equations (3.3) and (3.4) are repeated from the previous chapter: Vp=-B-TY8+—-VQ+-——ET——le (3.11) Vw (3.12) 4. The Perturbation Expansions In this section the beginnings of the perturbation method for the pure thermal diffusion method are developed. In the introduction to this chapter it was suggested that in a typical pure thermal diffusion experiment the diffusion flux is small, that if there is only a small heat loss the convective velocity is small, and finally that the vari- ability of the phenomenological coefficients is very slight. While it might be true that the phenomenological coeffi- cients are fairly constant in experiments in which the gradients of temperature, pressure and composition are small, the neglect of their variation in the transport equations constitutes a serious introduction of systematic error in many cases. Of course the differential equations are much easier to solve if they have constant coefficients and not infrequently it might be suspected that the cri- terion of constancy is more strongly influenced by the re— duction of work in solving the equations rather than by any physical evidence of constancy. In any case, the perturbation method is particularly well suited to handle problems of this type. Its use focuses attention on the major processes while at the same 73 time making possible corrections for the minor effects of variable coefficients. This is accomplished by expanding each of the coefficients in a thermodynamic series in the manner of Hurle, Mullin, and Pike (1965) and introducing the parameter y for the purpose of ordering the terms. The typical coefficient is expanded in the following man- ner: L = L0 + y[LT(6-60)+Lc(wl-wlo)+Lp(0'00)] + 0(y2) (4.1) Where L0 = L(GO'WIO'DO) (4.2) LT = 8%) 0=00,w1=w10,p=p0 (4'3) LC (381') e0'w10'90 (4.4) and L0 = (3%) eo'wio'po (4.5) Corresponding to the ordering of terms by the parameter y, each of the independent variables is given by a power series expansion in y, so that the contribution of each term in the series to the total is proportional to the effects of the perturbation at that order. It is easy to see that if the differential equa- tions are solved with the parameter Y set equal to zero, 74 then the solution is that for constant coefficients, whereas if Y equals one, then the full variability of the coeffi- cients is taken into account. This same analysis can be applied to the heat loss a as a parameter. If all the independent variables are eXpanded in a power series in s, then the solutions to the differential equations with 6 equal to zero correspond to the zero convection problem. The perturbation expansions for the independent variables thus become double power series expansions in Y and e of the following form: 00 oo.k 8 = T + 2 1 le T. (4.6) r i=0 k=0 1k M °° °°ik p = —7—— + 2 2 6 Y d. (4.7) a uh i=0 k=0 1k M 00 oo 1 i k w = ——- + 2 2 e Y C. (4.8) 1 M i=0 k=0 1k . co mik j = Z Z 8 Y J. (4.9) ”1 i=1 k=0 ”1k ” m i k v = Z X 8 Y v. (4.10) ” i=1 k=0 ”1k where the first term on the right in (4.6), (4.7), and (4.8) has been introduced for the purpose of making the boundary conditions homogeneous. The summations on i in (4.9) and (4.10) begin with one in order to indicate that in the case of zero heat loss no convective or diffusive effects are expected. expanded 5. begins with the transport equations listed in section 3. The expansions (4.6) 75 according to (4.1) in the following way: 0' + 80 + 86 + Y[OT(6-TO)+0C(wl-wlo)+op(0-00)] YIDT(6-T0)+Dc(wl-wlo)+Dp(0-00)] YlKT(6-T0)+Kc(wl-w10)+Kp(0-00)] YIQT(9-TO)+QC(w1-wlo)+op(0-00)] Y[VT(6’T0)+Vc(w1-w10)+vp(9-90)1 Y[HT(6-To)+HC(wl-wlo)+Hp(o-po)] Y[BT(6-RO)+BC(wl-wlo)+8p(0-00)] + + + + + The phenomenological coefficients to be used are 0(Y2) 0(Y2) 0(Y2) 0(Y2) 0(Y2) 0(Y2) 0(72) Y[8&(e-To)+sé(wl-wlo)+85(p-po)1 + 0(y2) Cv0 + Y[CVT(6‘T0)+Cvc(w1‘w10’+CVp(9‘90)1 + 0(Y2) “0 + Y[0T(8-TO)+uC(wl-wlo)+up(0-00)] + 0(Y2) no + Y[0T(6-T0)+nc(w1-wlo)+np(0-00)] + 0(y2) (0 + Y[¢T‘9'To’+¢c‘wl‘wlo)+¢p‘p‘po’1 + 0(y2) The Perturbation Equations (4.11) (4.12) (4.13) (4.14) (4.15) (4.16) (4.17) (4.18) (4.19) (4.20) (4.21) (4.22) The analysis of the pure thermal diffusion experiment to (4.22) are substituted into each of the equations in section 3 and into the boundary conditions of section 2. Terms of the same order in6:and~onr each i and 76 k are collected and separated. The equations at each order may then be solved independently of the solutions of any higher order. The zeroth-order equations.--After the substitu- tions indicated above, the terms of the equations contain- ing neither parameter are rearranged to give 2 _ v T00 — 0 (5.1) th - _ o A_T, Ycoo " UOYToo ‘(ZTT “—0. [YToo + (11h) (5'2) _ M I _ Ydoo — d00 + —§——)[8091 (V000+80)YT00] a Nb 2 V gh 0 M AT + "TH T doo + "2— [YToo + (Ta—)5] (5'3) 0 a nh The boundary conditions become T = T — T + l (AT) at z = h (5 4) 00 m r 2 ° T = T - T - i (AT) at z = 0 (5 5) 00 m r 2 ' 3T 00 _ _ 3r — 0 at r — a (5.6) for temperature, and Zn h a -I- .f .[ doordrdzdw = 0 (5.7) O O 0 77 C rdrdzdw = 0 (5.8) for composition and density. The zeroth-order solutions.--The zeroth-order equations have been arranged so that the composition and density distributions may be easily found from the tempera- ture distribution. The equation for temperature to zeroth- order (5.1) has a general solution consisting of linear terms together with periodic terms appropriate to the boundary conditions (5.4-6). The nature of these boundary conditions eliminates the possibility of periodic solutions to (5.1). The temperature can therefore be given by a linear function which according to (5.4-6) is TOO = 1%21 (z - h/2) + (Tm - Tr) (5.9) The gradient of the zeroth-order temperature solution is most conveniently written YToo = ' (%§)§ (5'10) The density distribution is next determined from equation (5.3). The solution is easily seen to be an exponential which satisfies the boundary condition (5.7). This solution is 78 d = M M exP [A(z - h/2)] - 1 (5.11) 00 .72.. 2 .....438. where _ a _ A: A — 80g (Vooo+ BO)(h.) (5.12) Finally, if the temperature solution is substituted into equation (5.2) and the density solution used in the boundary condition (5.8) the following solution can be found for the composition distribution: (2 - h/2) + % - — hcoth(2) (5.13) 00(AT) C =— 00 h In most practical problems the quantity A has a -2 magnitude of about 10 , and the hyperbolic functions can conveniently be replaced by their expansions: sinh(%P) = (fig) + %(§E)3 ... (5.14) coth (1:31) (‘33—)-l + 1(5h—)- 1 (53f (5.15) In using these expressions the cubic terms (as well as all higher terms) are negligible with respect to the first. Insertion of (5.14) into (5.12) and (5.15) into (5.13) yields 800 = 2M exp [A(z - h/2)] - 1 (5.16) a fih and 00(AT) 2 _ Ah . 79 after the higher order terms of the eXpansions have been dropped. It must be remembered that these are only approx- imate expressions for use in practical calculations where the magnitude of A is small. Equations first order in Y.--After the perturba- tion expansions have been introduced into the transport equations the terms of these equations which multiply Y represent the corrections to the basic processes due to variability of phenomenological coefficients. The intro- duction of the perturbation expansions (4.6) to (4.10) into the thermodynamic expansions (4.1) of the phenomeno— logical coefficients gives for the typical coefficient 2 L = LO + YL1 + 0(Y ) (5.18) where Ll has been introduced for compactness of notation and is defined as M _ _ _l... L1 ‘ LT(TOO + Tr To) + LC (coo + M w10 + L d + M — p (5 19) p 00 2 0 ° a Uh The transport equations of first order in Y are 2 -(A—T) cm (5.20) 1__ 01 KO __1 dz VCOl == 0 lVTO m[— AT 930 YTOl (5.21) 80 2 M 8' v0gh Yd01 ‘ ‘ Adelé + doo + 22;; 195 ‘ V000 + Bo ' EETKTT'YT01 - (v100 + voo1 + 81) YTOO (5.22) From (2.1) to (2.8) the corresponding boundary conditions are T01 = 0 at z = O (5.23) Tol = 0 at z = h (5.24) 3'1.01 = 0 at r = a (5.25) 3r for temperature and 2w h a J( ,[ df~ delrdrdZdw = 0 (5.26) 0 0 0 2H h a J[ _[- _[. Cood01 + doo + 34 C01 rdrdzdw=-o (5.27) 0 0 o a nh for density and composition. Solutions first order in y.--In the same way as in the zeroth-order case the boundary conditions on tempera- ture admit no periodic solutions to (5.20). Under these conditions the only possible solution for the temperature is K 2 Tol =(%§)K§ ggig 2 -(%§) 3%; {KT + KCOO} z(z—h) _A% +(93) KOMe (1 - eAZ) (5 28) 2 . h 2K0a n51nh A? 81 while the gradient is most conveniently written K 2 YT01= " (éf1$)fi 221E; -‘él712 1T1; {KT + KCOO} (2 ’ g) h -A— %) K pMAe Zih eAz l (5.29) 2Ko a 2nsinh A7 Substitution of this expression for the temperature gradient into equation (5.22) for the density together with the detailed expressions for the phenomenological coeffi- cients, allows equation (5.22) to be written dd 01 _ AZ M dz ... Adol + [B + 2C2 + ADe ](d00 + 3; (5.30) where A has been defined in (5.21) and AT h B = [8'9 ' (HT) (00" T+OTV0+BT)][Tm ‘ To ‘ (T '2'] M I ._AT AT-D__1._E Ell- +[Bcg (Tm—“Go;+°cV0+BC)H’0M(TN 2+A 2 com A2 +M W10! Vzgh V0 AT AT h -VO’+B- ATM ---—— 2-——(K+I a12 a22 Y2 127 is positive (q 2 0) for any Y if and only if both all 2 0 (A7) and 2 ‘ a12 all a22 2 0 . (A8) Proof. Equation (A6) may be rearranged to give a12 2 312 2 q = a ‘Y + ——— Y + a - ——— Y . (A9) 11 1 all 2 22 all 2 Thus if for any Y1 and Y2, all 2 0 and alla22 - a12 2 0 then q 2 0. On the other hand if q 2 0 for any Y, and Y2 = 0: a 12 2 Y1 + a 11322 " 5‘12 3 0' then all 2 0, while if Y2) = 0 then a Definition. Let V 36 5 2 add 6:0,..." (A10) d=l and \J 05 = 2 06a 6:0,...n . (All) d=1 Lemma II. 0:0,...0 (A12) Proof. For each 0, 0:0,...0, choose 128 §0 = §d = 0 ’ a > v ’ §a = §Y , a=1,...v, y+6, d+5 . (A13) Then by use of (AlO-ll) l . ,.__1~ “66 2 (36+C6)‘968 X6 0 = xtS x .1__;l 1 v . (A14) 7(36+C5)-(266 SE135+Q55-BG-C xY Use of Lemma I and (A1) requires V 1 2 066 £138 2 4 (135 + Ca) , 6=0,...n . (A12) 3 = The final theorem which gives (A5) is now: Theorem. Given a scalar, invariant, positive, bilinear function 0, defined by (Al) Q Ill IIM=I IL: w o Q where the X0 are a linearly independent set of vectors which determine the vectors Ja uniquely according to the linear relations TT ~a = 8:0 QanB , a=0,...n , (A2) then if \) ) Q08 = 0 , 8:0,...n , (A3) (A1) requires Proof. and therefore or By hypothesis 129 (1:0,...71' o 6=0,...n , (1:0,00011' 0 (A5) (A13) (A14) (A15) (A5) APPENDIX B PARAMETER ESTIMATION PROGRAMS l. Subroutine MINIMIZE The subroutine described here was written in FORTRAN for use with a CDC3600 digital computer. It is designed to minimize a function of up to ten variables by choosing conjugate search directions. This assures that a quadratic function of n variables will be mini- mized in at most n steps. (If the number of variables exceeds ten, the program must be redimensioned.) For a theoretical description see the article by Powell (1964), "An efficient method for finding the minimum of a func- tion of several variables without calculating derivatives." There are three considerations for the use of the subroutine which must be tailored to the individual pur— pose. 1. Calling statement. The subroutine may be called from a FORTRAN program by means of the following statement: CALL MINIMIZE(X,N,EPS,ENDNORM,ITMAX,IPRINT,SUCCESS,FNORM) 2. Parameters. a. X = A linear array dimensioned for the number of variables. The program should be called 130 131 with a set of initial guesses for the variables stored in X. The solution will be returned in X. N = The number of variables (less than ten). EPS = A convergence criterion parameter. The change in each variable from the-last step is compared with EPS times the current value, and convergence is assumed if the change is smaller. ENDNORM = A convergence criterion parameter. The function value at the current point must be less than ENDNORM to obtain convergence. ITMAX = The maximum desired number of iterations. IPRINT = An option. If IPRINT equals unity, the program will cause the results to be printed. If IPRINT is zero, no results will be printed. SUCCESS = A logical variable to indicate con- vergence. If success is unity, the process has converged. If SUCCESS is zero, the method has failed to converge, and a statement will be written to indicate the reason for termination. (SUCCESS must be declared TYPE LOGICAL in the calling program.) FNORM = The name of the function subroutine which calculates the function to be minimized. (See the next section. FNORM must be declared EXTERNAL in the calling program.) 132 3. Required subroutines. a. QUADMIN. This is a routine required by MINIMIZE and a listing is given together with MINIMIZE. b. FNORM. This is a subprogram in which the function to be minimized is calculated. It must have the following form: FUNCTION FNORM(X,N) DIMENSION X (N) (any necessary calculations) FNORM =- f(X(1), X(2), ..., X(N)) RETURN END where f is the function to be minimized, and X and N have their previously indicated meanings. Listing of MINIMIZE and QUADMIN 134 SUBROUTINE MINIMIZE(XQNOEPSIQEPSZQITMAXQIPRINTQSUCCESSQFNORM) DIMENSION X(N)9XO(10)9Y(10)9R(10910) COMMON/MIN/LASTNORMoKOUNT EXTERNAL FNORM TYRE REAL NORM. LASTNORM TYRE LOGICAL SUCCESS IF(N.GT¢10)GO TO 5000 ITER=O KOUNT=O D01 I=19N X0(I)=X(I) P(IQI)=OoI*XO(I) IE(X0(I)¢LT¢(1¢OE-7))P(I:I)=OoOI L=I+l DO 1 J = LoN 1 P(IQJ)=P(J91) =OoO LASTNORM = FNORM(X9N) KOUNT=KOUNT+1 NM=N~1 IF(IRRINT)PRINT lOOoLASTNORMqX 100 FORMATIIHIQ*THE INITIAL VALUES ARE*O//O5XQ*NOQM*QIOX9*XO(1)OOQOOXO(N)*9 I(N)*o//09E1506o/1(15X08E1506)) IF(IRRINT)RRINT 110 110 FORMAT(1H60*ITER INC*95X9*NORM*9lOXo*X(1)oooooX(N)*o//) 1000 ITER =ITER+1 IF(ITER.GT¢ITMAX)GO TO 3000 95 DELTA=1.0E-IOO M=O F1=LASTNORM DO 2000 I=IQN DO 2 J=IoN 2 Y(J)=P(J9I) CALL QUADMIN(X CY ONORMQN gFNORM) IF(IRRINT)PRINT lOloITERoIqNORMcx 1(31 FORMAT(2I598E15.60/o(25X97E1566)) IF((LASTNORM-NORM).GE.DELTA)3,4 3 M=I DELTA 8 LASTNORM-NORM 4 LASTNORM = NORM 2000 CONTINUE 2004 F2=NORM IF(ITER.GT.N)15916 155 IF(NORM.GT¢EPSZ)16¢I7 1‘7 IF(ABS(FI-F2)oGToABS(ERSl*F2))16919 1‘? DO 18 I=loN IF(ABS(X(I)-XO(I))oGToABS(ERSI*X(I)))16918 153 CONTINUE GO TO 4000 15> DO 5 =10N £5 Y(I)=2.0*X(I)-XO(I) 135 F3=FNORM(Y9N) KOUNT=KOUNT+I IF((F3.GEoFl)o0Ro(((F1-2.0*F2+F3)*(F1-F2-DELTA)** 12).GE.(DELTA*((F1-F2)**2)/2.0)))6o7 6 DO 8 I=1¢N 8 XO(I)=X(I) GO TO 1000 7 DO 9 I=19N 9 Y(I) =X(I)-XO(I) CALL QUADMIN(X0YoNORMoNoFNORM) DO 10 I=IQN IO XO(I)=X(I) DO 11 I=M9NM DO 11 J=10N 11 P(JoI)=P(JQI+1) DO 12 I=19N 12 P(ION)=Y(I) LASTNORM = NORM GO TO 1000 3000 PRINT 102 102 FORMAT(1H49*THE MAXIMUM NUMBER OF ITERATIONS HAS BEEN EXCEEDED*) SUCCESS =0 PRINT 5004gKOUNT RETURN 4000 PRINT 1039ITER 103 FORMAT(1H49*THE PROCESS HAS CONVERGED IN*91693X9*ITERATIONS*) SUCCESS =1 PRINT 50049KOUNT 5004 FORMAT‘lH-O*THE NUMBER OF FUNCTIONAL EVALUATIONS WAS*,110) RETURN 5000 PRINT 5001 55001 FORMATC1H40*MORE THAN 10 VARIABLES. PLEASE REDIMENSION.*) SUCCE55=0 RETURN END SUBROUTINE QUADMIN(X9P.NORM9N9FNORMI DIMENSION PHII3)9VT(3)0X(N)9P(N) COMMON/MIN/LASTNORMQKOUNT TYPE INTEGER UPPER TYPE REAL NORM vLASTTo LASTNORM DO 9 I=loN 9 X(I) =X(I) +P(I) LASTT = 1.0 T3000 ITER =0 1(3 ITER = ITER + l NORM = FNORM(X¢N) KOUNT=KOUNT+1 IF((ABS(T-LASTT)oGTo(001*ABS(T))oANDoITERoLEoEO)oORoI 111912 11 13 14 15 16 17 18 19 20 21 22 23 24 136 IFIITERoEOoIII3oI4 VT(11=OOO VTI3) I“1.0 PHIII) =LASTNORM PHII3) =NORM IF(PHI(I)oGToPHI(3))10? T=200 LOWER=I MID=3 UPPER=2 K=2 GO TO 1000 T=-l.0 LOWER=2 MID=1 UPPER=3 K=2 GO TO 1000 PHI‘II)=NORM XW =VT‘21‘VTI3) XX=VT(3)‘VT(I) XY =VT(1)-VT(2) XW =-(PHI(1)*XW+PHI(2)*XX+PHI(3)*XY)/(XW*XX*XY) xx=(PHI(11-pHI(2))/XY-XW*(VTII)+VT(2)I LASTT =T IFIXWOGTQOQO)ISQI6 T=-XX/(2.0*XW1 GO TO 19 IF(PHI(UPPER)¢GT.PHI(LOWER))17918 T:3.O*VT(LOWER)‘2.O*VT(MID) GO TO 19 T=3.0*VT(UPPER)‘2.O*VT(MID) IF(T.GT.VT(UPPER1)20921 I=LOWER LOWER =MID MID = UPPER UPPER 31 K=UPPER GO TO 1000 IFCT.LT.VT(LOWER))22923 I=UPPER UPPER =MID MID =LOWER LOWER=I K=LOWER GO TO 1000 IE‘T.GT¢VT(MID))24025 I=LOWER LOWER =MID MID=I 25 1000 1001 12 137 K=MID GO TO 1000 I =UPPER UPPER =MID MID=I K=MID II=K VTIK)=T DO 1001 J=19N X(J)=X(J)+(T-LASTT)*P(J) GO TO 10 IF(NORM.LE.LASTNORM)RETURN NORM=LASTNORM DO 7 I=10N X(I)=X(I)‘LASTT*P(I) RETURN END 138 2. Program FOTOFIT The fringe shape analysis described in section 4 of Chapter IV is performed by the program FOTOFIT and its associated subroutine FRNGFIT and FRNGSTAT. FOTOFIT is an executive routine which handles the reading of data, transformation of data, and calling of auxiliary programs. The actual calculations of fringe shape and weight func- tions are done in FRNGFIT while the estimates of standard deviation are made from the moment matrix calculated by the subroutine FRNGSTAT. Of course the minimization of the difference between the calculated and experimental fringe shapes is done by the routine MINIMIZE which must accompany these routines. The data input format is given by FORMAT statement 3 in the listing that follows. The first card of the data deck contains an identification code, the number of data cards to be read, the number of derivatives to be estimated, and the parameters A, b1, K and u1 as specified in equa- TI tion (4.6) of Chapter IV. The remaining data cards are read according to FORMAT statement 2 and each card contains one (xi, zi) point. As many fringes as desired may be analyzed at one time. The only requirement is that a blank card be placed at the end of the data deck. Listing of FO'I‘OFIT, FRNGFIT, and FRNGSTAT 1000 10 999 140 PROGRAM FOTOFIT DIMENSION X(lO) COMMON/TRNS/Z(29100)oNOPT¢BloA9TK9UT1oWTIlOO)9F(100)9TEMP9FOTOLBL TYPE LOGICAL SUCCESS EXTERNAL FRNGFITQWTFIT READ 3OFOTOLBLONOPTONOAQBIOTKQUTI F0RMAT(A802I594E15.6) IF(NOPT¢E000)GO TO 999 D0 1 I=IQNOPT READ 292(19I19Z(29I) FORMAT(F6¢39F5.3) 2(19I13200*(Z(10I)+0035)/404 CONTINUE DO 10 I'IQN X(I13000 CALL MINIMIZE(X0N00000101.00259IQSUCCESSQFRNGFIT) CALL MINIMIZEIXQNQOOOOOI01000501IOSUCCESSQWTFIT) CALL FRNGSTAT(X9N) GO TO 1000 CONTINUE END FUNCTION FRNGFIT(XQN) DIMENSION X(N) COMMON/TRNS/Z(29100)oNOPToBlvoTKoUTIoWT(IOO)0F(100)9TEMP9FOTOLBL DO 1 1819NOPT WTIII=100 LUSE=1 GO TO 2 ENTRY WTFIT LUSE=0 CONTINUE M=N-1 DO 3 I'IQNOPT YP=Z(IQI)+BI YL=YP‘20*B1 TP=TK*(YP**2-1001/40+YP/20 TM=TK*(YL**2-1001/40+YL/20 IFILUSE)GO TO 5 TEMP=A*UT1*TK+BI DO 4 J31 QM TEMPITEMP+A*XIJI*(J+I,*(Tp**J*(TK*YP+IOI“TM**J*(TK*YL+Io)1/20 WTIII=IoO/(100+TEMP**2) CONTINUE DN=UT1*(TP*TM) DO 6 K'IOM DN=DN+X(K)*(TP**(K+1)“TM**(K+1)1 FCI)=X(N)-Z(29I)+A*DN TEMP=000 DO 7 I=19NOPT TEMP=TEMP+WT(I)*F(I)**2 3 2102 3000 2106 2101 3001 3002 301 3003 211 141 FRNGFIT=TEMP RETURN END SUBROUTINE FRNGSTAT(X9N) DIMENSION X(N)¢D(10010)oS(10v100) COMMON/TRNS/ZCZQ1001CNOPTQBI.AoTKvUTIoWT(IOO)OFIIOOIOTEMPQFOTOLBL M=N-1 DO 1 I319NOPT S(NQI)=Io DO 1 J=19M YP=Z(IOI)+BI YL=YP-20*BI TP=TK*(YP**2-1oO)/4o+YP/2. TM=TK*(YL**2‘IoO’/4o+YL/20 S(JoI13(TP**(J+11-TM**(J+1))*A DO 2 K=IoN DO 2 J'ION D(K9J)‘000 DO 2 ISIQNOPT DIKOJ)8S(KQI,*WTII)*S(JOI)+D(KOJ, CALL INVERSE(D¢N9N010E~7QDET010010) IF(DET0E00000)PRINT 3 FORMAT11H6¢*MATRIX IS SINGULAR*) PRINT 21020(IOPQIOP=10N) FORMAT(/25X0*SIGMA MATRIX*/169121101 PHI=NOPT~N BV=TEMP/PHI D0 3000 I=1QN DO 3000 J=IoN D‘IQJ)=D(IOJ,*BV DO 2106 J=10N PRINT 21019J0(D(JOI)OI=19N) FORMAT(* *9I2913E1002) DO 3001 I’IQN S‘IOII'SORT(D(IOI)) PRINT 30020FOTOLBL FORMAT(1H~¢*FINAL RESULTS FOR RUN*9A8q//9*PARAMETER NUMBER*95X0*ES ITIMATE*913X0*STDo DEV.*) DO 301 I819N PRINT 3003910X(I)9$(191’ FORMAT(IH 95X.I5c5X9E150605X0E1503) BV=SQRT(BV) PRINT ZIIIBVOPHI FORMAT(* STANDARD ERROR OF Y *9E15o69* 1F8.2) PRINT 90769(19F(I)91=19NOPT) 9076 FORMAT(4(4H FO(13q3H) =E15.7.5X)) RETURN END C6H12294 29 5 2.394998E+5 2.272727E-1 -1.4001o504 DEGREES OF FREEDOM *9 1.200610E-3 *1.8500C -1.3001o257 -1o2001o062 -1o10000894 ~1.0000.74S -0.9000o619 -0.8000o514 -O.7000o405 —0.6000o333 -Oo3000.104 -0.20000034 -0.1000o014 00.0000o000 0010000009 0.20000025 0.3000o074 0.4000o159 0.5000o225 0.6000o324 0.7000o460 0.8000o569 009000.691 1o0000o845 1.10010046 1.2001o215 1o3001o524 01.4001o834 142 143 3. Program EXPCURVE EXPCURVE is a program similar in nature to the FOTOFIT program and used for the time constant analysis of pure thermal diffusion experiments. Again the program is simply an executive routine which reads data and calls on the auxiliary routines EXPO and EXPOSTAT. The expo- nential functions are calculated together with weighting functions in EXPO while the subroutine EXPOSTAT handles the statistical calculations. The data input is as follows: The first card to be read contains the number of data points in the time series to be analyzed. The second card contains the ini- tial estimates of the parameters a, b, and c in equation (5.2) of Chapter IV. Each subsequent card is punched with one (di' ti) data point which is to be read according to FORMAT statement l03. As listed the routines analyze only one set of data at a time, but they may be easily repro- grammed to handle multiple data decks. Listing of EXPCURVE, EXPO, and EXPOSTAT 101 102 103 145 PROGRAM EXPCURVE DIMENSION X13) COMMON/TRNS/ZI2QIOO)ONOPTOTEMPOWT(100) EXTERNAL EXPOQWTEXPO TYPE LOGICAL SUCCESS READ IOIQNOPT FORMAT(IZ) READ 1029(X(I)91=103) FORMATIF6029E4039F502) DO 1 I=IQNOPT READ 103921201)!Z(101) FORMAT(F7029F502) CALL MINIMIZEIXOBOOOOOOI0005910.IQSUCCESSOEXPO) CALL MINIMIZE1X930000001900591091OSUCCESSQWTEXPO) CALL EXPOSTAT(X9N) END FUNCTION EXPO(X9N) DIMENSION X(N)$F(1001 COMMON/TRNS/Z129100)QNOPToTEMPQWTI100) DO I I=10NOPT WTIII=100 LUSE=1 GO TO 2 ENTRY WTEXPO LUSE=O CONTINUE DO 3 I‘IQNOPT S=X(1)*EXP(-X12)*Z(EOI)) IFILUSE1GO TO 4 WTII1310/(Io+(X(2)*S)**2) F11)=Z(1911-$-X(3) CONTINUE TEMP=OOO DO 5 I=IoNOPT TEMP=TEMP+WT(I)*E(I)**2 EXPO=TEMP RETURN END SUBROUTINE EXPOSTAT(X9N) DIMENSION D(30319X(N)OS(391001 COMMON/TRNS/Z129100)QNOPTcTEMPoWT(100) DO 1 I=IQNOPT 51101)=-EXP(-X(2)*Z(ZQI)) 5(291)=Z(2.I)*X(1)*EXP(-X(2)*Z(2.I1) $139113-100 DO 2 K=103 DO 2 J=193 D(K9J)=Oo0 DO 2 1:1.NOPT D(K9J)=S(K01)*WT(I)*S(J91)+D(K¢J) 146 CALL INVERSE(D.3.3.1oE-7.DET.3o3) IF(DET.E0.0.0)PRINT 3 3 FORMAT(1H6.*MATRIX IS SINGULAR*) PRINT 2102.(IOP.IOP=1¢3) 2102 FORMAT(/25Xo*SIGMA MATR1X*/16o12110) PHI=NOPT-3 BV=TEMP/RHI DO 3000 1:1.3 DO 3000 J=103 3000 D(IoJ)=D(IoJ)*BV DO 2106 J=1.3 2106 PRINT 21019J0(D(J9I10I=193) 2101 FORMAT<* *012913E1002) DO 3001 1:103 3001 5(1.1)=SORT(D(IcI)) PRINT 3002 3002 FORMAT(1H-.*FINAL RESULTS*9/.*PARAMETER NUMBER*o5X.*ESTIMATE*.5X.* ISTD. DEV.*) DO 301 1:193 301 PRINT 3003oIvXII)oS(IQI) 3003 FORMAT(1H .SXo15.5x9E15.6.5X.E15.3) BV=SORT(BV) PRINT ZIIOBVQPHI 211 FORMAT(* STANDARD ERROR 0F Y *.E15.6o* DEGREES OF FREEDOM *. IFS-2) RETURN END 33 F6037.06833.032-0.2501.7500.9766934.950 -OS.50.01610.07 0025.0005.71 0030.0005.97 0035.0006.22 0040.0006.54 0045.0006.75 0050.0007.13 0055.0007.30 0060.0007.63 0065.0007.90 0071.0008.16 0075.0008.39 0080.0008.45 0090.0008.66 0100.0008.78 0110.0009.11 0120.0009.31 0130.0009.41 0140.0009.46 0150.0009.74 0160.0009.58 0170.0009.82 DISC-0009.72 019000009085 020000010007 0212.0009o99 0222:0009-82 0240-0009o84 0252.0009o97 0266.0009.96 0280.0010o01 030000009099 0314.0010007 0364-0010007 147 APPENDIX C VELOCITY AND TEMPERATURE PROFILES FOR THE CONVECTIVE HEAT LOSS PROBLEM l. Velocitnyrofiles The detailed solutions for the vertical and radial velocity components in a pure thermal diffusion experiment with radial heat loss have been given in Chapter III, section 5. It is shown there that the velocities are determined by the potential ¢ according to -1 _ M 1 30 V10x "' [doc + F] P ’5? “3'“ -l _ _ M 1 34> VlOz " [doc + 17'] a? a? “-2) where r and 2 have been scaled to unit length and where, by equation (III.5.16), -1 M _ w [doo + T] " M The potential ¢ is given in equation (III.5.62) as 2 2 N M O = r(r-l)z(z-l) r(r-l) + 2 2 C i=0 j=l ..zl 1] 148 exp [A(1/2 - 2)] (C.3) sin1rjr (C.4) 149 The coefficients Cij have been determined by the minimiza- tion of the integral in equation (III.5.65) using the MINIMIZE routine (see Appendix B). It_is clear that their values depend only on the magnitude of the quantity A. The numerical values determined with A equal to 0.01 are 0.12007 - 0.00597 0.00105 = - 0.17734 - 0.00409 0.00197 (C.5) 0.17785 - 0.00396 0.00201 Having determined these coefficients, one can easily calculate vlOr and v10 for any r and z. The 2 vertical velocity lez has been plotted versus r for 2 equal to 0.5 and is shown in Figure 1. For other values of z the vertical velocity has the same shape but de- creases in magnitude almost parabolically from the center towards the horizontal plates. The radial velocity vlOr has been plotted versus r for z = 0.25 and the results are shown in Figure 2. Again the shape of vlOr remains essentially the same for all values of z. The magnitude of v r decreases 10 as 2 approaches 0.5, where ler changes sign, and then increases in magnitude. It should be recognized that the signs on the velocities as they have been plotted are not absolute, but should be multiplied by the sign of the heat loss coefficient. Thus if there is a heat loss, the fluid 150 102 Figure l.-&Vertical velocity as a function of r at the center of the cell (2 = 0.5). vlOr 0.25 0.20 0.15 0.05 151 Figure 2.--Radial velocity as a function of r at z = 0.25. 152 near the wall will be relatively more dense than that in the center and the velocity near the wall will be downward. 2. Temperature Profiles The temperature correction, T10 (r,z), for the convective heat loss problem is determined according to equation (III.5.73) by 00 T10 (r,z) = nil Tn(r) sin (nnz) (C.6) where Tn(r) = ClI0 (nnr) + Tnp (r) ’ (C.7) and the coefficient C1 is given by equation (III.5.83). The zeroth order modified Bessel function-has been denoted by I0 (mrr). The particular solution, '1' (r), is given by np Tnp (r) = 220 Gng’r2 (C.8) and the coefficients Gn2 are given by equations (III.5.77-8). The temperature correction depends on the phenome- nological coefficients for the system through the three parameters A, pl, and p2. The coefficient A is again taken to be 0.01 while pl is taken to be 100 and p2 to be 500. These are values typical to common liquids. The shape of T is also affected by the value of the quantity lO 153 (Tr - Tm + 1/2 AT) which enters through the boundary conditions. The sign of this quantity determines whether there is an overall heat loss or heat gain (- and + respectively). The temperature correction has been plotted versus r at the center of the cell (2 = 0.5) under the three conditions (Tr - Tm + l/2 AT) = - 5.0, 0.0, and 5.0 in the Figures 3, 4, and 5 respectively. It is interesting to note that the lepe of T10 versus r has the same sign as (Tr - Tm + l/2 AT) as r approaches l.0 indicating heat loss or gain as the case may be. The temperature corrections appear to be largest near the center of the cell (2 = 0.5) and to decrease in either direction towards the horizontal plates. 10 -6.0 -700 -800 154 Figure 3.—-Temperature correction versus r at z = 0.5 10 155 Figure 4.--Temperature correction versus r at z = 0.5 for (Tr - Tm + 1/2 AT) = 0.0. 10 -405 E Figure 5.--Temperature correction versus r at z 156 for (Tr - Tm + l/2 AT) 5.0. 0.5