TRACER AND MUTUAL DIFFUSION IN SEVERAL [SSTHERMAL NON - IDEAL LIQUID NON ~ ELECTROLYTE SYSTEMS Thesis for the Degree of Ph. D. MESHIGAN STATE UNWERSHY C. MICHAEL KELLY] 1970 LIBRARY Michigan Stats? University P {HR-“S This is to certify that the thesis entitled TRACER AND MUTUAL DIFFUSION IN SEVERAL ' ISOTHERMAL NON-IDEAL LIQUID NON-ELECTROLYTE SYSTEMS presented by C . MICHAEL KELLY has been accepted towards fulfillment of the requirements for Ph . D . CHEMICAL ENGINEERING degree in Major professor Date February 25, 1970 0-169 ABSTRACT TRACER AND MUTUAL DIFFUSION IN SEVERAL ISOTHERMAL NON—IDEAL LIQUID NON-ELECTROLYTE SYSTEMS By C. Michael Kelly Hydrodynamic theory is used to develOp equatiOns predicting the effect of intermolecular association upon tracer and mutual diffusion. On the basis of simple assumptions about the volume of associated complexes, it is shown that Onsager's Reciprocal Relation should be valid in certain associated systems. An experimental study is made of tracer and mutual diffusion in several systems. It is found that the association characteristics of a given system may be determined from plots of the tracer diffusivity-viscosity product vs. composition. It is further shown that several systems which are non-associated, as can be seen from the D*n products, fail to obey the Hartley-Crank equation. Possible reasons for this failure are presented. C. Michael Kelly A study has been made of the method currently employed for measuring ternary diffusivities. Weaknesses in the current method are pointed out, and suggestions are made for improvements. Within experimental precision, however, ternary measurements support both the predictions of hydrodynamic theory, and the Onsager Reciprocal Relations. TRACER AND MUTUAL DIFFUSION IN SEVERAL ISOTHERMAL NON-IDEAL LIQUID NON-ELECTROLYTE SYSTEMS By c. Michael Kelly A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemical Engineering 1970 r: ézaOl- 7-/-7o ACKNOWLEDGMENTS The author wishes to express his appreciation to Dr. Donald K. Anderson for his guidance and assistance throughout the course of this study, and to the chemical engineering faculty in general for their support and encouragement. The author is grateful to the National Science Foundation for financial support during his graduate studies, and to the Division of Engineering Research for support of this work. The patience and encouragement of the author‘s wife Ruth is deeply appreciated. ii TABLE OF CONTENTS Page ACKNOWLEDGMENT. . . . . . . . . . . . ii LIST OF TABLES. . . . . . . . . . . . V LIST OF FIGURES . . . . . . . . . . . vi INTRODUCTION . . . . . . . . . . . . I BACKGROUND . . . . . . . . . . . . . 5 Hydrodynamic Flow Equations . . . . . . 5 Solution Thermodynamics . . . . . . . 9 Nonassociated Solutions . . . . . . ll Associated Solutions . . . . . . . 15 THEORY . . . . . . . . . . . . . . 18 Solution Thermodynamics . . . . . . . 18 Nonassociated Solutions . . . . . . 18 Associated Solutions . . . . . . . 22 Tracer Diffusivities . . . . . . . . 26 Nonassociated Systems . . . . . . . 26 Associated Systems . . . . . . . . 28 Binary Mutual Diffusivities . . . . . . A0 Nonassociated Systems . . . . . . . 40 Associated Systems. . . . . . . A2 Ternary Mutual Diffusivities. . . . . . A5 Nonassociated Systems . . . . . . . 45 Associated Systems . . . . . . . . SA EXPERIMENTAL . . . . . . . . . . . . 66 Tracer Diffusivities . . . . 66 Calculation of Tracer Diffusivities . . . 69 Mutual Diffusivities . . . . . 72 Analysis of Binary Mutual Difquion Experiment. . . 7A Calculation of Binary Mutual Diffusivities . 78 Analysis of Ternary Mutual Diffusion Experiment. . . . . . . . . . 79 iii Page RESULTS AND DISCUSSION . . . . . . . . . 87 Tracer Diffusivities . . . . . . . . 88 Mutual Diffusivities . . 101 Error Analysis - Mutual and Tracer Difqui on 106 Ternary Diffusivities . . . . . . 109 Error Analysis - Ternary Diffusion. . . . 114 SUMMARY . . . . . . . . . . . . . . I24 FUTURE WORK. . . . . . . . . . . . . I26 APPENDICES . . . . . . . . . . . . . 129 A. Procedure for Tracer Diffusion Experiment. . . . . . . . . . 130 Procedure for Mutual Diffusion Experiment. . . . . . . . . . 138 C. Computer Programs . . . . . . . 147 D. Ternary Intermediate Data . . . . . 153 E. Experimental Results . . . . . . . 166 F. Thermodynamic Data. . . . . . . . 171 G. Nomenclature. . . . . . . . . . 17A BIBLIOGRAPHY . . . . . . . . . . . . 179 iv LIST OF TABLES Table ' Page 1. Summary of Hydrodynamic Theory Predictions of Tracer Diffusivities . . . . . . . 39 2. Predicted and Experimental Ternary Dif- fusivities in the System Chloroform (C) — Acetone (A) - Benzene (B) at 25°C. . . . 110 3. Reduced Sensitivity Coefficients, 8: in the . » i System Acetone (A) - Benzene (B)- Chloroform (C) for all Choiced of Coordinates . . . . . . . . . ', . 119 A. True Initial Times for Ternary Runs, as Predicted by Equations (187) and (188) . . 122 LIST OF FIGURES Figure Page * 1. Din vs Mole Fraction for Associating Components . . . . . . . . . . . 89 x 2. Din va Mole Fraction for Nonassociating Components . . . . . . . . . . . 91 3. Mutual and Tracer Diffusivities for the System 2-Butanone — Carbon Tetrachloride at 25°C a o c o o o o o o o o o 914 A. Tracer Diffusivity - Viscosity Products for the System 2—Butanone - Carbon Tetrachloride at 25°C . . . . . . . 95 5. Mutual and Tracer Diffusivities in the System p-benzoquinone (A) - benzene (B) at 25°C . . . . . . . . . . . . . 97 6. Mutual and Tracer Diffusivities in the System Diethyl Ether - Carbon Tetrachloride at 25°C . . . . . . . 99 7. Mutual and Tracer Diffusivities in the System Chloroform (A) - Carbon Tetrachloride (B) at 25°C . . . . . . 103 8. Mutual Diffusivities in the System Benzene (A) - Chloroform (B) at 25°C . . . . . 105 x 9. Din vs Mole Fraction in the System Acetone Chloroform at 25°C . . . . . . . . 111 x 10. D n vs Mole Fraction in the System Acetone Benzene at 25°C . . . . . . . . . 112 vi INTRODUCTION Interest in molecular diffusion in liquids has increased considerably in the past few years, from both experimental and theoretical points of view. Several techniques have been developed for measuring ordinary (both binary and multi—component) and tracer diffusive fluxes [22].. A large number of binary solutions, both electrolytic and non-electrolytic, have been studied. In the past few years a number of multi-component systems have also been investigated. Although much work has been done, when one considers the large number of simple systems available it becomes apparent that the surface has barely been scratched. Much more work needs to be done before there can be a precise understanding of molecular diffusion. Theoretical efforts have centered on determining the relationship between diffusive fluxes and the physical and chemical properties of the system, such as molecular weight, molecular size and shape, viscosity, state variables and solution thermodynamics. There has also been considerable interest in the relationship of ordinary diffusion to tracer diffusion, both from a predictive and a correlative standpoint. In multicomponent l systems, much emphasis has also been placed upon verify- ing the theory of Onsager based upon the principles of irreversible thermodynamics, particularly the Onsager reciprocal relationships. There have been two basic theoretical approaches to the description of diffusion processes. One is based upon modifications of the absolute reaction rate theory of Eyring [16], and the other upon modifications of the hydrodynamic flow model of Stokes [29]. This work will follow the hydrodynamic approach. According to hydrodynamic theory, transport of a species through a solution in which there is a concentra- tion gradient of that species takes place by means of two processes. The first process is the flow of individual molecules through the surrounding medium as a result of a force acting upon those molecules. This has been termed by Hartley and Crank 'intrinsic diffusion' [20]. The second process is the transport of molecules due to flow of the medium. This flow occurs because of hydrostatic pressure gradients which arise from the differing volumes of the diffusing species. Hartley and Crank termed this process bulk flow. The first process can be characterized by a combina- tion of chemical and physical properties of the diffusing species which Hartley and Crank called the 'intrinsic diffusivity.‘ This 'intrinsic diffusivity' is the product of two terms, one involving the physical properties of the diffusing species and the surrounding medium, and the other involving solution thermodynamics. The first term will be called here the "intrinsic mobility" of the species, as suggested by Carman [6]. Equations have been developed relating diffusivities to the intrinsic mobilities of diffusing species and solution thermodynamics. These have been modified by assuming that in some systems molecular interactions can be characterized by a chemical association. In these systems, a given stoichiometric component may undergo intrinsic diffusion not only as monomers, but as dimers, trimers, and other associated complexes as well. In this work ordinary (binary mutual, and ternary) diffusion has been studied by means of a Mach-Zehnder interferometer [5], and tracer diffusion has been studied by a capillary technique, for several systems of interest. It is shown that the degree of associative behavior in a given system can be determined from the tracer diffusi- vities of the components. Equations are developed relating association (as determined from tracer diffusivities) to the intrinsic diffusion process, and to solution thermodynamics, upon which ordinary diffusion is highly dependent. These equations will be tested by the diffusivity data previously mentioned. It will be shown that hydrodynamic theory predicts that the Onsager reciprocal relations are valid for non— associated systems, and several specific types of associated system. Experimental measurements made in the ternary system acetone - benzene - chloroform agree with Onsager's reciprocal relation within experimental accuracy. Unfortunately, precise experimental verification by this method is quite difficult, for reasons which will be discussed later. An analysis has been made of the probable causes of low experimental precision in the measurement of ternary diffusivities by this method. Possible avenues of investigation will be suggested which might lead to an improvement in the method. It is hoped that future work along these lines will lead to a precise experimental verification of Onsager's reciprocal relation. BACKGROUND Hydrodynamic Flow Equations Hydrodynamic theory states that a diffusing molecule behaves like a particle undergoing viscous flow through a continuous medium. The driving fOrce for diffusion which causes this flow is generally agreed to be the gradient of the chemical potential of the diffusing species, acting in the direction opposite to the gradient of the chemical potential: Fid = - Vui (1) Since there is assumed to be no acceleration, this must be balanced by a drag force upon the molecule, due to the viscosity of the medium. Sutherland [29] and Einstein [10] independently showed that the viscous drag force for a sphere flowing through a continuous medium is given by = - 2 Fsr 61rrsnvSm ( ) where rS is the radius of the sphere, n is the viscosity of the medium, and vsm is the velocity of the sphere with respect to the medium,-and the negative Sign is because the drag force is in the direction opposite to the flow. If the molecule were truly a sphere diffusing through a continuous medium, equations (1) and (2) could be combined to obtain dui .. “dz = enrinvim (3) dui where Vui has been replaced by d§_’ denoting one- dimensional diffusion. Multiplying by the concentra— tion of the diffusing species, and solving for the flux of that species gives C. du. =_ l l i Vim Ci 6nrin dz (4) m is the flux of species i with respect to the where Ji medium. However, most molecules are essentially non— spherical, and unless the diffusing molecules are much larger than the surrounding molecules the medium cannot be considered continuous. Therefore, the radius of the diffusing molecule rS will be replaced by an empirical constant 2%, which will be called the 'friction coefficient‘ of species i. This yields Ci dui Ji =-61-fi-a'E- (5) The defining equation for the chemical potential of species i is given by 0 pi = “i + R1 Ln ai (0) Substituting into equation (5) gives, at constant T and P, Ji = - o.n EE- (7) m Ci [aLnai Sci] 1 L. i T,P T,P This is the expression for the flux of Species i due solely to the random molecular motion of 1 molecules, with respect to a coordinate system fixed in the sur- rounding medium. However, this is not a directly measurable quantity (although experiments can be conceived which could measure this flux, they are beyond the capabilities of current techniques). The flux which is measured in most experiments is the flux of component i with respect to a coordinate plane across which the total volume flux is zero. This flux may be obtained from equation (7) by the following argument. First, the flux of component i with respect to a coordinate system fixed in the laboratory in an N-component system is given by the expression .m ' <8) HMZ <1 c_. + < ll <1 where vmc is the velocity of the medium with respect to laboratory-fixed coordinates, and v is the Vc velocity of the volume-fixed coordinate system with respect to laboratory—fixed coordinates. This can be seen by making a balance of volumes of diffusing species in the laboratory-fixed coordinate system. The quantity which has the units of velocity, represents the flux of volume due to random molecular motion, relative to the medium. To relieve hydrostatic pressure gradients, the medium itself must flow, relative to fixed coordi- nates, and this velocity is Vmc' Unless the partial molar volumes of the diffusing species are constant, the total volume changes as diffusion proceeds. As a result, the volume-fixed coordinate system acquires a velocity relative to fixed coordinates, va. The total volume flux (which is a velocity, and is represented by the left side of equation (8)) with respect to fixed coordinates must equal the velocity of the volume-fixed coordinate system with respect to fixed coordinates. Equation (8) can then be solved for the velocity of the medium: i=1 1 i Vc (9) Wirth [34] gives a detailed derivation of this expression. The measurable flux Jiv is given by i (Vmc - VVc) (10) Combining equations (7), (8) and (10) gives the basic hydrodynamic flow equation: v i J = - ——— i din Q) A I: 3' p. + :3 },.1 ’=1 0 Bz I (11) Solution Thermodynamics Historically, there have been two approaches to describing non-ideal liquid solutions. It is agreed that intermolecular forces lead to deviations from Raoult's law. The differences between the two approaches arise from different interpretations of these inter- molecular interactions. The more widely accepted approach, as originally developed by van Laar [32] and van der Waals and their 10 followers, considers all intermolecular forces to be general in character. They arise from such phenomena as coulombic attraction and repulsion, dipole interaction (both permanent and induced dipoles), and van der Waals forces. This approach is called the "physical approach," and is readily applicable to most types of solutions. The second approach, originally set forth by Dolezalek [9], considers all deviations from Raoult's law to arise from specific intermolecular forces which lead to chemical bonds between molecules. According to this theory, a solution of components A and B consists of A monomers, B monomers, plus various associated complexes such as A2, A3, A4,..., B2, B3, B4,..., AB, A2B, A32, and so forth, depending upon the specific interactions present. These individual species are then assumed to obey Raoult's Law. The proportions of the species present in solution are determined by an equilibrium characterized by an equilibrium constant, such as A + B ++ AB. Because of KAB this equilibrium assumption, this is known as the "chemical model" of solution non-ideality. Dolezalek originally proposed this model before the nature of chemical bonding was well understood. He was led by his model into some rather improbable hypotheses. For example, he tried to describe the ll vapor-liquid equilibrium in the system nitrogen—argon by postulating the dimerization of argon, a most unlikely occurrence. These two approaches are not mutually exclusive, though for years there was rather heated debate between the two schools. As our knowledge of chemical bonding increased, it became apparent that some systems really do associate in liquid solution. This is especially true of molecules which are capable of hydrogen bonding, such as water, alcohols, amines, etc., and of molecules which form charge-transfer complexes. On the other hand, there are many non-ideal solutions in which the formation of associated complexes is rather unlikely. Further, there is no good reason to conclude a priori that the various species present in an associated solution should obey Raoult's Law, as Dolezalek assumed. It would seem logical to try to combine the two approaches. Non—associated Solutions Before considering associated solutions, it would be well to look at non-associated solutions. Perhaps the simplest method of describing activity data in non-associated systems is to assume that the natural logarithm of the activity coefficients can be expressed as a power series expansion of the mole fractions of the stoichiometric components. This is the approach l2 taken by Margules in deriving the equations which bear his name. The constants in the power series expansion are restricted by the Gibbs-Duhem equation. Within these restrictions, the values of the constants are determined by fitting the series to experimental thermodynamic data, generally vapor-liquid equilibria, by means of least—squares analysis. The equation can be made to fit experimental data to whatever degree of accuracy desired by simply taking more and more terms into the series expansion, though at the expense of introducing more arbitrary constants. It can be extended quite easily to multicomponent systems, and is not restricted in its range of application except by the number of terms in the series that one wishes to use [35]. Though this procedure is mathematically rigorous, and useful for describing experimental data for use in design calculations, it sheds very little light on the true nature of interactions in liquid solutions. Van Laar [32] proposed a far more restricted equation, based upon theoretical considerations, for binary systems. This equation, and modifications of it, have been very successful in describing binary systems, especially those for which the activity data are rather symmetrical, and for which the molecular sizes and shapes are not too different. It has several disadvantages, though. It is not easily extended to multicomponent 13 systems, without introducing further assumptions and more arbitrary constants, which tend to decrease the physical meaning of the equation. Also, it cannot be used reliably for those systems in which there is considerable bonding—type molecular interaction, nor for such systems as high-polymer solutions. Both the Margules and van Laar equations, as well as several modifications, are discussed in considerable detail by Wohl [35]. Another approach which is based upon theoretical thermodynamic considerations is that of Hildebrand and Scott for regular solutions [25]. Since it forms the basis for some later conclusions, it will be described in a little more detail. If a solution contains enough thermal energy, the different intermolecular forces of the various com- ponents will not be sufficient to cause any one molecule to tend to aggregate with any particular type of molecule, either like or unlike. The entropy of mixing will then be the same as for an ideal solution. Such a solution is termed 'regular', even though it is non- ideal, and the partial molar entropy of mixing is given by AS = - R Ln x (12) 1A By making three assumptions, Hildebrand and Scott show that the heat of mixing in the binary regular system of components i and j is (13) where oi is the volume fraction of component i (neglecting expansion on mixing) and 61 is defined by where EiV is the internal energy of vaporization. The assumptions leading to this relationship are: (a) the energy of interaction between two molecules depends only upon the distance between them and their orienta- tion, (b) the volume change of mixing at constant pressure is zero, and (c) the mixing of molecules is random. The third assumption is essentially the definition of a regular solution. The first, although not rigorously correct, has been the basis for most successful attempts at modeling the liquid state. The second can be eliminated by extensive modification, as shown by Hildebrand and Scott [21], but will not be done here. For regular solutions, where the entropy of mixing is ideal, the activity coefficients are given by [> El H. 1 Ln yi = = if. 4). (c3 - s.)2 (114) :30 *3 P C4 L1. P This can be extended to multicomponent systems quite easily. Under the same assumptions as before, for a ternary system we find that where with 6 and ¢i as defined before. Detailed derivations i of equations (12) through (15) are given by Hildebrand and Scott [21]. Associated Solutions The intermolecular forces which define an associated solution would seem to be precisely those forces which disqualify that solution from being considered a regular solution. In a regular solution, the molecules mix as though they had no preference as to the nature of their nearest neighbors. In associated solution, on the other hand, any given molecule has a distinct preference for another molecule as its nearest neighbor, as expressed by the polymerization equilibrium. For example, a molecule with a hydroxyl group will prefer 16 to have another molecule with a hydroxyl group as nearest neighbor (rather than a saturated hydrocarbon, say) due to its ability to form the hydrogen bond. It might then be a good assumption that all those forces which lead to a solution not being regular are due to complex formation by chemical bonding. Certainly chemical bonding would be the major contributor to non— regularity in associated solutions. It would seem logical then that the true species present mix to form a regular solution. The mole fractions of the true species are determined by the equilibrium equation and the stoichio- metric mole fractions. Consider for example a binary system in which one component dimerizes as a regular ternary system, consisting of monomers of each component plus dimers. Equations (15) can then be used to predict the activity data from knowledge of the equilibrium constant K. Alternatively, the equilibrium constant can be determined from activity data by adjusting it until equations (15) give the best fit. This procedure requires knowledge of the partial molar volume and the molar energy of vaporization of the dimer. This can be handled in either of two ways. These quantities may be treated as adjustable parameters, in which case equations (15) will be a three—parameter set of equations for the binary system [13]. Otherwise, l7 assumptions can be made about the values of these parameters, or their relation to those parameters for the monomers. Equations (15) will then be a one-parameter set of equations for the activity coefficients in the binary system. The latter method will be used later for fitting activity data in both binary and ternary associated systems. THEORY Solution Thermodynamics—- Nonassociated Solutions The Gibbs free energy of a system of N components is given by where n1 is the number of moles of i, G10 is the molar free energy of pure component i, Xi the mole fraction of i, and AG represents the difference in free energy between one mole of real solution and one mole of an ideal solution with the same composition. The partial molar free energy of component i, the chemical potential of i, is given by :1 LG = “10 + RT Ln Y1 XI (17) where “10 is a function of T and P only. Carrying out the indicated differentiation on equation (16), and equating to equation (17) gives Ln yi - filr‘r‘ —°—— [AG 2 mil (18) 19 If a power series expansion is written for AG, and the coefficients constrained by the Gibbs-Duhem equation, after drOpping terms of higher order than X3 there results + X1X32al3 + X2X3 2a23 2 2 . 2 1 2 33112 1 2 3&122 + X1 X3 3a113 (l9) 2 2 2 113 2 3 3a223 + X2X3 3a233 X X + X1 2 3 6a123 This is the three-suffix Margules Equation for a ternary system [35]. Binary systems may be treated as Special cases, and the corresponding Margules Equation obtained from equation (19) by simply setting X3 equal to zero. Carrying out the differentiations in equation (18) gives the activity coefficients: 2 = [- .. _. Ln Yi 2X1X2 LA21 X1A21 X2A12J + X2 A12 . 2 _ _ ] + 2xlx3 [A31 x1113l X3A13] + x3 A13 (20) + (x2x3 - 2xlx2x3) [A21 + A13 + A32 - t] 2O 2 u — r _ _ n Y2 2X2X1 ~A12 X2A12 X1A21J + X1 A21 2 + A _ _. 2x2xj [A32 X2A32 ‘ X3A23] + x3 A23 (21) + (xlx3 - 2xl x3) [A2l + Al3 + A32 0] Ln Y = 2X X [A - X A - X ] + x 2 A 3 3 l ‘13 3 l3 1 31 l 31 2 + 2X2X3 [A13 - X3A13 x1 31] + x1 A31 (22) + (xlx2 - 2xlx2x3) [A21 + A13 + A32 - C] where the constants are defined by 2a12 + 3a122 = Ai2 2a12 + 3°112 ‘ A21 2&13 + 3a133 ‘ A13 (23) 2al3 + 3a113 = A31 2a23 + 3a233 = A23 2a23 + 3a223 = A32 3all2 + 3°i33 + 3a223 ' °°i23 = C The binary analogues to equations (20), (21) and (22) are 2 Ln yl = 2x lx2[A21 - xlA21 - X2Al2] + x2 Al2 (2A) 2 Ln Y2 = 2X2 x ltA12 — X2Al2 - xl A21] + x1 A21 (25) 21 Wohl [35] gives detailed derivations of equations formally similar, and algebraically identical to equations (20) through (25). If isothermal ternary activity data are available equations (20), (21) and (22) can be fit to the data by a le st-squares technique, with seven adjustable m oarameters. This is an unreasonably large number, but some may be specified by other means. If the ternary data are available, then binary data for each of the three pairs of components are almost sure to be available. The binary data may be fit by a least-squares technique, thus fixing six of the seven constants in equations (20) through (23). The constant C may then be determined from the ternary data. In most cases of interest, ternary isothermal data is not available, but these equations are still useful. Wohl interprets the various constants in equation (19) in terms of physical interactions between molecules. Thus a12 represents the energy of interaction of molecules 1 and 2, a the interaction between two molecules of type 113 l and one of type 3, etc. From equation (23) we see that under this interpretation C is a ternary interaction parameter. Since in non-electrolyte solutions three- molecule interactions are not so strong as two-molecule interactions, C can probably be taken as zero. As long as this holds, ternary activity data can be predicted 22 from data for the subsidiary binaries. This has been done for the system acetone-benzene-chloroform. Data for the system acetone—chloroform have been taken from Hildebrand and Scott [21], for benzene-chloroform and acetone—benzene from Timmermans [30]. These data, and their approximations by equations (24) and (25), as well A as the least-squares parameters A 21,... are given 12’ in Tables F—l through F-A of Appendix F. Solution Thermodynamics-- Associated Solutions As discussed previously, an associated solution may be thought of as a regular solution of the true species present. By making approximations as to the values of‘V and 6 for the associated complexes, the equilibrium constants may be treated as adjustable parameters in fitting the activity data. This will be done here for a specific case, a binary system of components A and B, where the association reaction A + B ++ AB occurs. To avoid later confusion, let us refer to the stoichiometric components by letters A and B, and the true species present by numbers 1, 2, and 12, where 1 refers to the monomer of component A, 2 to the monomer of B and 12 to the dimer. The true equilibrium constant 23 for this reaction is given by Ka’ the ratio of activities of species: a waif l 2 Defining K and KY by K = Xi2 X1X2 Y: fi=_:‘_a. Y Y1Y2 we see that Ka = KK Nikol'skii [27] has shown that the chemical potential of a component in solution is equal to the chemical potential of its monomer: = X AYA 1Y1 Hence, the activity coefficient of component A is given by The activity coefficient YA is a directly measurable quantity, but Y1 is not. (27) (28) (29) (30) (31) (32) 24 For a regular solution, equation (15) holds: V . Ln Y1 = —l (51 —a>2 (15) RT ' where 61 and 8 are defined by ~ s h 31" 51 = :__ (33) V. l _ 3 5 = 2 n1 Si . (3“) i=1 xii/’1. 4’1‘xV +xi7‘ +xV (35) 1 1 2 2 12 12 Similar equations give the values of Lny2 and LnY12° Now, for a given X and XB’ the mole fractions of A the true Species depend only upon the value of K, and can be determined from equations (28) and the stoichiometric relationships (36) Let us make the following approximations: 25 a) V12 = V1 + V5 (37) b) E¥2 = EX + E; (38) Assumption a) is required for consistency of constant molar volumes which will be assumed later, and which Kett [23] determined to be a good assumption for the associated system ether—chloroform-carbon tetrachloride. Assumption b) is reasonable if the energy of the dimer— ization bond is approximately the same in the vapor state as in the liquid state, that is, if AE for the equilibrium reaction is the same in both states. Note that AH will probably be different, as there is a change in PV in the vapor state. Now equations (15) and (32) through (39) can be combined to give ln YA in terms of measurable quantities (V's, EV 'S, stoichiometric mole fractions, temperature) and one adjustable parameter K. It is now possible to determine the value of K from experimental isothermal activity data, by a least-squares technique. This has been done for the system ether- chloroform from total-pressure data at 25°C from Kohnstamm and van Dalfson [2A]. The results agree reasonably well with the data of Guglielmo [30] at an apparently higher unreported temperature for the vapor-liquid equilibrium. 26 This procedure can be easily extended to a ternary system of A, B, and C where A and B dimerize as before, and C is inert. The value of K found for the binary dimerization equilibrium Should not change in the ternary solution, provided the solvent C does not change the mechanism of the reaction but merely dilutes the reactive components. By making this reasonable assumption, it is possible to predict the ternary activity data for such an associating system from physical properties and activity data for the associating binary pair of components. These principles are easily carried over to other forms of association. Tracer Diffusivities-— Nonassociated Systems The tracer diffusivity is defined by a modified version of Fick‘s Law: (39) where the superscript * designates the tagged molecules.- Since the tagged molecules are considered to be identical to the untagged molecules physically and chemically, if there are no external pressure gradients there will be no bulk flow.. For every molecule diffusing in one direction, there will be another molecule diffusing back in the other direction. Since these molecules have 27 the same volume, there will be no hydrostatic pressure gradients, and therefore no bulk flow. In this case, the velocity of the medium-fixed coordinate system and the volume-fixed coordinate system will be zero (relative to laboratory-fixed coordinates). This means that for tracer diffusion equation (11) becomes C * d * —'. ll- J.V* = I.m* :: _ —_"'__. _I' ((40) l l o.n dz 1 For tracer diffusion equation (17) becomes * o. x x = 7“ "7‘ hi hi + R1 Ln Yi Xi (A1) * Applying the chain rule gives the derivative of “i : x x * Y x * du du. dC d Ln . d Ln x. d C = 1? d:=RT[‘—T“—‘l+ *lein) °Z dci dci d0; Z Since the solution is chemically uniform, it can be Shown that x d Ln Yi - 0 _____¥__ dCi (43) x d Ln Xi —"——-'¥' d Ln Ci 28 Combining equations (A0), (A2) and (A3) gives v, RT i (AA) i din , (A5) Carman [5] suggested that the combination of physical properties on the right hand side of this equation be termed the ‘intrinsic mobility' of species 1. This designation will be adopted here. The important result is that the tracer diffusivity of a non-associated com- ponent is equal to the intrinsic mobility of that component. Tracer Diffusion-—Associated [Systems The most easily analyzed case of tracer diffusion in an associated system is the binary system of A and B where there is an association to form an AB dimer. Denoting the species present as l, 2 and 12, the total concentration of tagged component A is given by C = C + C (A6) The flux of A molecules in tracer diffusion is 29 V*-_ ‘ _TV* V* JA ’ DA dz ‘ ”1 + J12 (“7) Substituting equation (A0) for the fluxes gives * *- dC dC l + 1 A2 ] (A8) dz dz °12 V* = _ ET _1 JA n Ecl * Differentiating equation (A6) with respect to CA gives 8C * “C * 1 ° 2 ’ A 8 CA Substituting into equation (A8) gives * ii- * it so 1 acA RT 3012 A *“ dz - o * dz JV*=__Rr£.[l- (50) A °1n ac 2 A If the physical and chemical properties of the tagged molecules are the same as the untagged molecules, we may assume that the distribution of tagged molecules is the same in the two species as in the component: c ____ = ___ = ___ (51) This ratio is a constant, since it depends only upon the * proportion of CA and CA when the solution was made up. Therefore, .30 'X' 8C12 _ 12 - C (52) acA A Now for convenience, define the pseudo-mole fraction of component 12 by X12 = E“:‘C" (53) Combining this with equation (52) and substituting into equation (50) gives, after rearrangement of terms 0 * V, _ RT - 1 1 1 X12 aCA _ J1 "—n'-3‘+<5‘-“3— 7—3? (5“) 1 12 1 A Comparing this with the defining equation, Fick's Law, leads to the desired result: 0 X * r't DA=5—;-£-C,—l-+(;L-3-l-) -——,l{21 (55) 2 12 1 A The derivation for component B only requires renumbering the species and components in the previous equation, so that A becomes B, 1 becomes 2, and vice versa. Equation (55) then becomes 0 x * _ RT . 1 1 _g 12 DB-—HL3—+<5—-o -———X1 (56) 2 12 2 B 31 Wirth [3A] derived equivalent equations, and carried the derivations out for two other simple systems, a binary system where one component is inert and the other forms a self-dimer, and a ternary system where two components form a cross-dimer and the third component is inert. These equations, along with those to be derived in this work are given in Table l, on page . Theoretically, equations corresponding to (5A) and (55) can be developed for any associating system, provided an equation can be written for each association equilibrium. Practically, such equations become very difficult to handle if there are more than 2 or 3 such equilibria. Furthermore, Since a considerable part of the value of these equations lies in their ability to model the measured tracer diffusivity data, the number of associations must be small, or there will be a large number of adjustable parameters available to fit the data. Also, whenever the equations become too complex, they lose much of their physical meaning in the algebra. We will now consider another simple system, and develop equations predicting the tracer diffusivities. Let components A, B and C be ternary system in which there are two competing equilibria A + B ++ AB A+C++AC 32 Let the true species present be designated 1, 2, 3, l2 and 13 representing A, B, C monomers, AB and AC dimers respectively. The relative concentrations at any compo- sition can be determined from the equilibrium constants K and K . The fluxes and concentrations are given by l 2 the following: 8C V * . V JA x = _ DA "'5“: = J1 * + J1 V* + .113V* (57) NC * V * o B V V _ JB * = ' DB 32 = J2 * + J12 * (98) 8C V * C V V JC * = — DC -a-Z— = JB * + J13 * (59) if 'X' * 9(- CA - C1 + 012 + 013 (60) * x * , C2 = C2 + C12 (01) if * 9(- CC = C3 + C13 (62) Equations (58), (59), (61), and (62) are equivalent to those for the binary system just considered. The tracer diffusivities are analogous to those in equations (5A) and (55): X o l 4) i2] 1(63) 2 °12 °2 B 33 a: :I) *3 1 l3 ) i: 1 (6A) + ( _ 1 D - [3; o O :5 _I C73 The equations for component A are slightly more complex. From equation (57): * x * V* _ RT 3C1 RT 8C12 RT ac13 , JA — - [c n 82 + o n 32 + o n 32 J (05) 1 12 13 * Differentiating equation (60) with respect to CA and substituting into equation (65) gives C n * x * x V, _ RT 1 3 12 3"13 1 8C12 1 ac13 CA JA - -7 5—(l-——-.xr - -———T)+C 0’ f] (66) 1 BOA BOA 12 BCA 13 BCA * * 8C12 3C13 AS before, the derivatives ————¥ and ———;— can be written 8C BC A A 0 C12 in terms of pseudo-mole fractions X12 = C + C “+ C and A C C ° 13 . X = to give 13 CA + CB + CC 0 o . * v, RT 1 1 1 X12 1 1 X13 °CA JA - - —H E 3— ( o - 3—) X + (o - 3—0- X J 82 1 12 1 A 13 1 A (67) 3A From this it can easily be seen that O O x 1 12 1 1 13 2 + (--—- -) + (— --) -——-] (08) 1 XA 013 O1 XA which is the desired result. The generalization to a larger number of competing equilibria is obvious. If component A dimerizes with N other components, the tracer diffusivity of A will be given by . 0 DA = --]-:- C— + Z (5__ - a") 31(1 1 (69) ' 1 i=2 11 l A This equation is general, and though it provides physical insight into the effect of several association reactions on tracer diffusion, it is probably not too useful in fitting experimental data, since it allows N adjustable parameters (the friction factors Gli) if the equilibrium constants can be determined independently, or N2 if they are also considered free. The physical meaning of this equation is that the tracer diffusivity of an associating component is equal to the intrinsic mobility of that species, decreased by the difference between the mobilities of the monomer and the associated complex (corrected for the amount of association) for each of the association reactions. 35 This same type of result can be obtained for the slightly more complicated case in which one component associates with itself. Consider the binary system in which component A reacts to form a dimer, according to * and component B is inert. In this case, the flux of A is given by ** 3011 11 RT ,2. 1 (70) 11n where the dimer can carry either one or two tags, as denoted by the number of asterisks. Stoichiometry Shows * * 'X' ** = + CA C1 + Cll 2Cll (71) Proceeding as before, we write 3" * 3C * 3" ** BC * v, RT ”1 RT 11 ”11 . A . JA 1 { ___¥ [._——¥-+ 2-———1—u} -———(72 nol n no 32 3 11 3C GO A A A Differentiating (71) gives 3C * 3C * "C ** 0 BOA BCA 30A 36 which, when substituted into (72) yields, after rearrangement, 5(- 96* V RT 1 1 1 ac11 1 1)°Cii 3 A .1A * + - -—l;— + (3—— - g— h + 2(g- - g— x 1 «z n 1 11 1 ac 11 1 ac ° A A (7A) Equation (7A) can be written ac * ac ** V m, .. JA * = - 5% 13$ + (g—i - 5;) (———£% + 2 -——i%—)J (75) 1 11 1 acA acA The derivatives again depend only upon the equilibrium constant K, although calculation of their values is some- what involved. The tracer diffusivity, from equation (77) is then ac * ac ‘* 3% DA =Eg-[5-l-+(5—l—-3-l-) <—1—}+2-—-i%—)1 <76) 1 11 l BCA 30A Wirth [3A] gives a Slightly different derivation, which leads to a formula which is less difficult to evaluate: * RT 1 -—-C 1 37 The form given by Wirth is more useful for trying to fit data for which there is only one dimerization equilibrium. It is diffucult to generalize, however, while the derivation of generalized forms of equation (78) is rather simple. Consider the system where there are two components, one of which undergoes two self—polymerization reactions, A + AA3 and the second Component B is inert. The flux of tagged A molecules is * C * ** v, RT 1 3C1 1 3 11 8C11 JA = - fi—'£3_' 82 + o E 82 + 2 3z 3 1 11 3C * “C ** a" *** -\ d U + O l (__%11_.. 2 __%ll__.+ 3 $11 . :)1 <78) 111 Z 02 Z and the stoichiometric formula is + *** <79) By the same process as before, this time leaving out the intermediate steps, “A * 3C *3!- * 1 l 1 OK“ A ‘B'fla‘l' <.——-.—> .4...” 11> 1 11 1 3C GO A A 8C * 3C ** 3" *** 1 1 'V + (O 1 _ 5-) ( 111 + 2 11 + 111 3 (80) 111 1 BOA BOA BOA The generalization of this equation by this method to a system where one molecule undergoes repeated Simple self— polymerization reactions is straight-forward, but notationally very difficult. The effect of repeated polymerization on tracer diffusivity is easily seen from equation (80), however. The tracer diffusivity in this case is given by the intrinsic mobility of the monomer, decreased by a correction factor for each polymer. These correction factors involve the intrinsic mobilities of each polymer, the number of tags carried by each polymer, and the amount of each polymer present (determined by the equilibrium constants). Once again this model is not too useful in fitting tracer diffusion data unless the number of polymerizations is small, and there is some information indicating that there is no further polymerization beyond a certain point. Of course, assumptions could be made to reduce the number of adjustable parameters. For instance, it would seem reasonable that for hydrogen bonding molecules, after the TABLE l.-—Summary 39 of Hydrodynamic Theory Predictions of Tracer Diffusivities. System and Association Predicted Tracer Diffusivities A, a, c,....u no association L)i noi i = A’ B’ C”"' N O n X «a DA *_.[_1.(_l_-_1)}1(23 “’ n 01 012 01 A A + B ++AB ( O I." ‘ a b, —; [—l + ( l - —l) A“ 1 -J O) 0.13 02 [LB 7 .1. O l" O A, a, c L, 1% [EA + ’01 - EA) :- + (El - El) 33 j “ 1 12 1 “A 13 1 ‘A A + a ++AB , .. O A + C +*AC -A H ~_i;_1.1_1-1>%23 ”a ” °2 °12 OT *5 1( O 1.11. . __l_ _ .1613; VG n ‘03 \013 01/ 2X“ 4 n c A, a, L, A w ET . _1 + E+l ( 1 _ 1. X11 7 A + 5-6»: “ n 01 i=2 O11 °1 ”A 4 A + 21+» All .... . . * ,. other components as in LB above A b , * C ** . , ’ D‘- F‘T {—3. + \__l_. _. A) 21.11.}. 2(_1 _. __1_) 34%.] A + A++:A, “ ” °1 °11 O1 3:, O1 011 acA B inert D2 RT 5 no; A, B 3 fi * C ** b a A + A+-A DA 5% [6; + (EL‘ ‘ El) “"li+ ““l%—' 2 1 11 1 8 C, BC A A A + A2++ A3 * ** **5 ac 3 c 3c - .1. _ .1 111 111 111 B inert (0111 01) (ma—E—1F + 2 -.—;E_¥ + 3 a )1 A ' A CA RT DB n02 no first one or two polymerizations, all the equilibrium constants might be assumed equal. It might also reasonably be assumed that no polymer carries more than a certain number of tags, say three or four at most. This would make computation much easier. The physical insight of equations of this type is considerable, as will be seen in later discussion. Binary Mutual Diffusion-- Nonassociated Systems For a binary nonassociated system with components A and B, equation (11) gives the flux of component A: v'- CA BuA C CAVA auA + CBVB BuB A 82 GB az (81) Substituting the definition of chemical potential gives J v'= BE.[_ SA a Ln aA + C (CAVA 3 Ln aA + A n 0A az A 0A 82 C V 8 Ln a + 8 B 82 B)] (82) B By the chain rule a Ln aA = _£ 3 Ln aA 8 CA (83) 32 C 8 Ln CA 8z “2:8 = .3: : :2 :3 B B Substituting (85) and (86) into (84) gives J V = -D 3C = 52 [_ _l 3 Ln aA + CAVA 3 Ln aB A AB 02 n GA 8 Ln Cn CB 3 Ln bB + CAVA : in :A] ~ (85) CA n A From the definition of mole fraction and partial molar volume the following hold: 8 Ln X X, a'Ln CA = c; (86) A B B 8 Ln XB = XA (87) 3 L C“ C V n D A A Substituting (86) and (87) into equation (85) and making use of the fact that CAVA + CBVB = 1 one obtains the well known Hartley-Crank Equation [20]: RT‘ B A A D = — E— —] --.——-— (88> AB n CA GB 3 Ln XA This equation has been used with some success in predicting mutual diffusivities in non-associating solutions. It was originally derived under the assumption that the molar volumes were constant, but this is not a necessary condition. Taking the limits as comparing to equation (#5) A2 XA + O and as XB + O, and we see that Lim . D _ = ——| = D (89) XA+O AB noA A Lim _ RT _ * XB+O DAB ‘ noB ‘ DB (90) Binary Mutual Diffusion-— Associated Systems Consider a system of two components A and B, in which component A undergoes the simple dimerization A + A ++ A2 and the second component B is inert. The flux of A is V , Cl aul Cll Bull Clvl 3“1 JA - ' o n 82 - 20 n 32 + CA( o 32 1 11 l Cllvll a“11 C2V2 3“2 + o 82 + o 32) (91) 11 2 where the true species have been numbered as previously. Proceeding as in the derivation of the Hartley-Crank Equation, using the definition of the chemical potential, the chain rule, the relation CAVA the mutual diffusivity is found + CBVé = l, and the 2V assumption V11 = A’ to be 43 x x O X 0 RT A l 1 ll 1 3 Ln a D =—-—[——+X..(-—-——+ ——-)l————— (92) AB n 02 D XA 01 XA all 8 Ln X Detailed derivation of this result is given by Anderson, <3r Wirth [34], and will not be reproduced here. Consider a system of two components A and B, in vfluich the dimerization reaction A + B ++ AB The flux of A is given here by occurs. v _ C1 3“1 C12 8“12 Clvl 3“1 JA - ' o n 32 ' o n 32 + CA< o 82 1 l2 1 + C12V12 01112 + C2V2,3“2) 012 32 02 32 By’ the same process, with the assumption that VBJB = VA + V5 this leads to the diffusivity: o o m X X DAB=E%[FLJ—XB+E££~XA l A 2 B o 2 + 1 X12 (XA ‘ XB) 3 Ln a (94) X X 3 Ln X °12 A B Detailed derivation is again given by Anderson [1] or Wirth [3n]. (93) 44 Note that equations (92) and (94) differ from the Hartley-Crank equation only in the addition of an extra term. In the case of self-association, this term is positive, and predicts that the mutual diffusivity is more than that predicted by the Hartley-Crank equation. In the case of cross-association, this term predicts a smaller diffusivity. It is generally true that the activity term in the Hartley-Crank Equation over-corrects. That is, when the system shows positive deviations from Raoult's Law, the thermodynamic correction predicts the diffusivity to be less than it would be if the solution were ideal. In many of these cases, the experimentally measured diffusi- vities are greater than those predicted by the Hartley- Crank Equation, though still less than for an ideal solution. When the system shows negative deviations from Raoult‘s Law, the thermodynamic correction predicts diffusivities higher than for an ideal solution. In these cases, the measured mutual diffusivities are found to be less than those predicted by the Hartley-Crank Equation, but still larger than for an ideal solution. Equations (94) and (96) reduce the magnitude of the deviation of the Hartley-Crank Equation from ideality. In a system where association takes place, they should be better predictors of diffusivity than the Hartley-Crank Equation. This has been found to be so for several 45 systems. However, in some cases there is strong evidence that the components do not associate significantly, and the Hartley-Crank Equation still tends to over—correct. Two such systems will be presented here. Some theoretical explanation must still be found for this discrepancy. Ternary Mutual Diffusivities-- Nonassociated Systems There have been two major approaches to multi- component diffusion. Onsager [28] proposed a set of equations for an N-component system relating the flux of each component to the concentration gradients of all the components, thereby defining N2 diffusion coefficients: J = g D 30' 1 j 13 -—l i = 1, 2,....N (95) Then, based upon the theories of irreversible thermo- dynamics, he showed that only (N - l)2 of these diffusion coefficients were independent. These diffusion coefficients, however, are not easily measured. Baldwin, Dunlop and Gosting [3] therefore proposed a different description, involving only N - 1 independent fluxes, and (N - l)2 diffusivities: Jul J 0 D15 ——l 1 = l, 2,....N - l (96) 32 _l a i 1 46 These diffusivities are not the same as the Onsager diffusivities, but are related by the expressions {71' .7 ' '=l,2,....N-l (97) Since then, Costing and coworkers have presented several methods for experimentally determining the Dij's [13, 14]. It is preferred, therefore, to relate hydro— chynamic theory to the Costing diffusivities. This will bee done for a nonassociated ternary system to demonstrate true method. It is also useful to develop equations puredicting the phenomenological coefficients of the Cuisager theory, to show that hydrodynamic theory predicts tine validity of the Onsager Reciprocal Relations. From equation (ll), the fluxes are V - Cl Bul A CiVl Bul C2Vé 3u2 C3V3 3u3 Jl - - o n 32 + °l( o 32 + o 82 + o 32) l l 2 3 . (98) J V = - C2 Bul + c (Clvl Bul + C2V2 8u2 + C3V3 3H3) 2 o2n 32 2 Ol 82 02 82 03 82 (99) 13 = - 8.1. La - :2. L... 32 C3 32 C3 82 47 Combining equations (98), (99) and (100) to eliminate the gradient of chemical potential of component 3 gives v c- l-V.C VQC, an , c c V V an 51 = " a: “—0; l) + 3‘3 5: ' "—1 2C3; ‘ 53'] ““322 (101) l 3 ° 9 3 2 J v = _ clcztv; _ El] 8ul _ 32 [(l-C2V2) + V3C2] 8H2 (102) 2 n 0 ol 82 n 02 G3 82 Mathematically, the total derivative of the chemical potential can be given in terms of the partials of all the independent variables. For a ternary system at constant temperature and pressure, there are only two independent variables, which may be taken as Cl and C2. Therefore, 8ul = 8ul 8Cl + 8ul 802 (103‘ 82 8 l 2 8C2 82 ) 8u2 = 8u2 8Cl + 8u2 8C2 (10“) 82 8Cl 82 8C2 82 Substituting these into equations (101) and (102) yields 7 V C C 8n v _ Ci (i-Clvl) C1V31 25“1 3 2 l 2 2 Cl J1"{T[T_+8Jac+(§_"a_) na—_}'§T l 3 l 3 2 l C — 1 "n _ {_l [(1 clvl + clv3] all + clc2 v; _‘V2) 8u2} at, n Cl 03 802 n 03 02 8C2 82 (105) II O . v _ 0102 v3 V1 8111 C2I_(l—C2V2) c273 3112 301 J2 - 4-? <5— “ 5—)T + “at o + 0 33c 82 1 1 2 3 1 _{clc2 (v3— _ [haul + oath-02V” + 057313112} 302 0 01 C2 n 02 03 8C2 82 (106) From the definition of chemical potential, these can be written J V = - 52 {C r(l-Clvl) + ClV3] 8 Ln a1. 1 n 1 Cl 03 801 3 2 1 RT r(l-Clvl) ClV3 8 Ln al ' _fi {GIL o + o 3 8C 1 3 l V V 8 Ln a 8C 3 2 2 2 + C C (—— ——) } (107) l 2 03 02 802 82 JV=-3?-{CC(Y-§-—Y33Lnal+ 2 n l 2 03 02 8Cl + C [(l'c2v2) + C2V'2j 801 8 Ln a2 } 801 2 02 03 82 801 82 “9 V 'V 3 Ln a RT 3 l l - ——-{c c <—— - -— n l 2 03 Cl 8Cl + 02[(l'C2V2) + C2V3] 8 Lralfla2 8:2 (108) 02 03 C2 2 Comparing these to the defining equations (96) gives the FT diffusivities predicted by hydrodynamic theory: ll AA n l Cl 03. 801 V ‘V 8 Ln a 3 2 2 + 0.0 — - ) } (109) l 2 O3 02 3Cl 12 AB n 1 Cl 03 802 V V 8 Ln a 3 2 2 - + C C —— — —— } (110) l 2 03 02 801 D =D =BE{CC (E-EBLnal 21 BA n l 2 03 Cl 801 1 c V‘ C'V 3 Ln a + c2 [( 2 2) + 2 3] 2} (111) (112) Kett [23] derived these equations, and generalized them to a system of N components, obtaining D = c; [(1—vici) + civN] 3:; 13 n Oi ON 8C3 N-l cch Vk V? a k k=l ” ON Ok 3 k#i The theory of irreversible thermodynamics states that the rate of entropy production in the ternary diffusing system is V 8“2 T ——-= — J ——— J ——— J- ——— dt 1 82 2 82 3 82 However, equation (100) can be used to eliminate u3, amd the constant volume relationship lel + J2V2 + J3V3 = O can be used to eliminate J3: T 9§ = J V Y + J V Y (11%) 2 8p. 1 = 2 Y = - Z (6 + l 1) —4l i 8:1 13 C353 32 5 ={0 J21 13 1 3:1 c... ll [1“ K + t“ K; c_. II L'" K: + L Substituting the expressions for Y from (115) into 1 equations (116) and (ll?) and rearranging yields V 8“1 3“2 J1 ’ -(aLll + YL12) 82 ‘ (BL11 + 5L12) 82 8p 8p V.. 1. __l _ ._i§ J2 “ ”(8L21 ' YL22) 82 (BL21 + 5L22) 82 where a, B, y, and 5 are defined by l l 2 l a=l+———-—— B=_..___ C3V3 C3V3 C3 3 C3 3 Equating coefficients between (118), (119) and (101), (102) yields four independent equations for the Li J's: (115) (116) (117) (118) (119) l l) l Llla + Ll2Y —E E C + O 1 (120) l 3 c c V V _ __l_2_ _3_ _ .2.- L118 + L12‘S ' n [o o J (121) 3 2 c c ‘V ‘V . _ l 2 ,-_3_ _ _2_ g... L210L + L22Y ' n L03 0 1 (122) g c 1-V c v c ' _ _2 ( 2 2) 3 2 L218 + L226 - n E O + ————O l (123) 2 3 The Onsager Reciprocal Relation states that L12 = L21. To test this, solve equations (120) through (123) to obtain —- - - la - --—E + 18 n o o n o o L = 3 2 l 3 (124) 12 a6 - BY 0102 [X1 V1:lcs C2 [(l-V2C2) + V3C2] n o ' 3— ' n o G Y L = 3 1 2 3 (125) 21 a3"; BY lV1 + C2V2+ C3V3= l, and the defining expressions for a, 8, y, and 6 we may obtain from Using the relation C equations (124) and (125) the desired result: 12 — 21 C l C2V1(1‘C1V1) 0' 1n 53 C C _ 1 2 Vé(1-CZVE) 0'21] 3 Therefore, hydrodynamic theory states that Onsager's Reciprocal Relation is valid for ternary isothermal dif— fusion in a nonassociating system. The only required assumption is that of constant molar volume, used in obtaining equations (101) and (102). Miller [26] has developed equations which allow Onsager's Reciprocal Relation to be tested experimentally: where = a D12 - c Dll ad - bc d D21 - b D22 ad - bc Clvl 8ul (l + 5‘5?) 35— 3 3 1 02V? 8u2 (1+é—V‘)'é'é" 3 3 l C V 8n 1 1 l (l + ) -—- C3V3 8C2 Q) “I: [\J c) O Q) "C: F" a) O Q) I: N co 0 m (127) (128) (129) (130) (131) d =(l+C_2_;,§.)::a+clV2.a_u_]; C3 3 8C2 C3V3 8 2 Ternary Mutual Diffusion-- Associated Systems (132) Kett [23] has developed equations for the ternary diffusivities in a ternary system of components A, B and C subject to the dimerization equilibrium where AA AB A +_B ++ AB component C is inert. Kett's equations are given here without derivation: c c- V'c an F l (1 V c ) + ‘2 (1 v c ) + 3 A] l 1 _ _ cl 1 A 012 12 A o3n 8CA V c c c- _ V’c c an E 202 A + 0 i2 (1’V12CA) + 30A B] 302 2n 12n 3n A — 2 C C V C 8n 1 — 12 — 3 A 1 [o n (l’VlCA) + o n (l V12CA) + o,n 3 8C - 12 3 B v c c c v c c an E“ 8 i A + o 12 (1‘V12CA) + 3oAnBJ 302 2 12 3 B (133) (134) ‘37. .13 BA " 02 _ c VCB E 32—; ”'Vzca) + —— ————n—-J 5—1— (135) + V c c c 'V c c aul D = [- -——l—§ + 12 (l-V12CB) + o O' — 2 C _ C _ V C 8n + [_—_2 (l-V ) + ——-——l2 (l-V C ) + 3 B J 2 c (136) G2” 2 B 012” 12 B By making the assumption that V12 =‘V + VB he obtained A c c V c C‘V .. - _.:.L__B__ -- _. 42.45.... -— LBA ‘ LAB ’ " oln <1 VlCA) 02 (l V208) c C V c A B 3 — — l2 - — + o3n (l'VlCA'V2CB) + 3:;fi(14vl2CA) (l-V12CB) (137) thus verifying Onsager's Reciprocal Relation for this simple associated system. The assumption of constant molar volume leads to the assumption made above, so this verification is exactly as reliable as the previous case. Kett [23] also developed similar equations and verified the Onsager Reciprocal Relation for the self— dimerization system of A, B and C, where A is subject to the equilibrium 56 A + A ++ A and B and C are inert. Equations of this type now will be derived for a slightly more complex type of associated system. sider the ternary system with competing equilibria pre- viously discussed under Tracer Diffusion. same nomenclature, the fluxes in ternary diffusion are Using the given by J V = _ .31 iii _ 012 Bul2 _ C13 3“13 A oln 82 012” 82 01311 82 n 01 82 012 82 013 32 87 a c V 1 2 2 .12 1 .3.1 .13 02 82 o 82 3 J V = _ C2 3“2 _ C12 Bu12 B o2n 82 012” 82 CB C1V1 3“1 + C 2V12 a“12 C13V13 3“13 + '71- E o 32 o 32 + o 32 1 12 13 (138) (139) For this system 8“12 = 8“1 82 82 8u13 = 8ul 82 82 Therefore, equations C V 1 _ UJA - -[EI(l-CAV1) + c 12 — -[-—(1-c V ) - 012 A 12 c V 12 nJ - -[-——(l-c V ) B 012 B 12 c -[5-1—2-(1-cBV2> + 12 c c V _ B 3 3J 3“3 G3 82 57 a“2 . 8u3 1 (138) and (139) become 0 c an IT ———(l-C V 1) + ———(l-C V )]——— 012 A 12 013 A 13 32 C C V Bu C C C V an A 2 2 2 B A 3 3 3 J -[ (l-C V )- J 02 82 5'13 A 13 03 82 (142) _ CBCI 1 _ CB013V1333‘J1 01 013 82 012 B 12 82 013 (143) By combining the stoichiometric relationships A l l2 13 CE = 02 + 012 (144) CC = C3 + Cl3 with the Gibbs-Duhem equation, we obtain ? 211-313.111-323“? (1145) ‘1 82 CC 82 CC 82 Since the chemical potentials are functions of CA and CB only, the expression for the total derivative gives B (146) (147) Substituting (145), (146) and (147) into equations (142) and (143) yields, after considerable algebraic manipula- tion an c c c nJA = -[,Cl{O l(1- -cAVl ) + 012(1— —cAV12 ) + (1-55)(1-CAV #) A 112 C13 c503 V3 23112 012 CB (:13 __ + .C o } + ac {o (l-CAV12) ' 5" (l-CAV13) c 3 A 12 c 13 _ CA02V2 + CACBC3V3 Joe, _ [3“1 {El(1-C'V ) o2 Cco3 oz BOB 01 A l 2C _. c 0 CA 3V 12 A 3 + (1-c V ) + (1———)(1— c )— } 012 A 12 cC AV 3 oi 13 Acc3 3 8“2 C12 1 "B C1 — CAC2V2 + at {o (l'tAV12) "5‘ 8""(1‘CAVB> ' o B 12 c 13 2 c c c V ac + ‘éfié—l'i} 1 .3 (1A8) CO3 02 6O an _ r‘JB "' ' [’a‘t‘l {FEM-CBV12) ' CB( <2: 1 + lg 13) A 12 1 13 c c c. V. c V Bu C C A B 3 13 3 3 2 12 V 2 '- + ( 1 ),+M{ (1CV)+-—(l—CV> CC 013 O3 °VA 0‘12 A 12 °2 B 2 c 2 c V c V ac a c + 2(13 13 + 3 3)}3 A A _ {£1 {lg—(141v ) C 013 03 OZ A 012 D 12 VlCl V13C13 CAGE Cl3VlB C3V3 ”an, + o >+c - C-\ + ——————> BB n 012 B 12 B Cl Gl3 cAcB 013V13 C3V3 Bul 1 012 _ + C ( o + o )1 BC + F [a (l'CBVl2) c 13 3 B 12 c _ c - c V V c an + 53(1—CBV2) + g ( 13 13 + g 3)] —53 (151) 2 C 13 3 B Toe cross-coefficients are given by the same expressions, except that the differentiation of the chemical potential is with respect to the other component. The validity of the Onsager Reciprocal Relation will now be demonstrated fcr this system. As before, the rate of entropy production is given by Bun 8n Bu” 8% 8n g§ _ _ 1 _ 1 2 _ T 3 _ 12 _ 13 . T dt J1 32 ”2 éz U3 32 J12 82 J13 32 (152) Applying equations (1“1 ) and (142) gives d8 8“1 3“2 T E? ‘ ’(J1 + J12 + J13) az ‘ (J2 + J12) az 3u3 8p? Equation (146) allows —§§ to be eliminated: C C 8p m ds- r V .1: ~ .1 -___1 1 5E — -LJl + 412 - d3 + (1-C ) J13J 82 C C CD 3u2 - LJ2 + J12 - q (.43 + J13)] T-Z— (1514) The assumption of constant volume lel + J2V2 + U3V3 + J12V12 + J13Vl3 = O (193) allows J3 to be eliminated from equation (154): T d8 - P* “J R' NJ (1") a? ’ ‘ d1 ‘ W 2 ‘ J12 ’ 13 9° where the expressions P, Q, R, W are defined by Applying equations (141) and (142) gives d5 341 8“2 * a? = ’(Ji + J12 + J13) az ‘ (J2 + J12) az 3u3 _ (03 + 013) BZ (193) an Equation (146) allows —53 to be eliminated: C C 3p m d8 _ F T i "’ ‘ i l 1 at ’ -LJ1 + “12 ‘ on 03 + (*'c ) J131 82 C C C: an? _ ' _ .z T 1 — \ The assumption of constant volume ’ - _ J _ — _ = R'— alvl + J2V2 + 3V3 + J12V12 + J13V13 0 (1,3) allows J3 to be eliminated from equation (154): l - QJ2 - Rdl2 — WJl3 where the expressions P, Q, R, w are defined by (156) 63 Bul C V aul CBVl 3u2 ——-—-——— + ————'-r“. CV3 82 ‘ CCVB oz 8L‘2 C3V2 aue CAV2 3“i Q‘ W+W7§+QV§TE 82 32 CC 3 oZ CCVé 32 w = 3“1 + °“2 + CAV12 “1 + (l_ig aul + CBV13 °“2 Bz 32 CCV3 2 CC 32 CC‘V’3 Bz _ .Cji 3.31.2 CC 2 JA = Jl + J12 + Jl3 JB = J2 + Jl2 yields m 9§ = — J P - J Q - J (R-P-Q) - J (w-P> (157) ‘ dt A B l2 13 Now if we assume that Vl2 = Vl + V2 and Vl3 = Vl + V3, leaving T g_s_=_J (3141 + CAVl 3“1 + CBVl 3112 dt A 32 CCVB az CCVS az -J(31‘2+E£.Y_2_3_‘11+C_BY32‘_2_) (158) B 32 GOV: az COVE az 64 Irreversible thermodynamics also states that the fluxes are related to the phenomenological coefficients: J v= _ L (3“1 + CAVl 3“1 + CBVl 3“2 A AA 82 C V 32 C V 82 C 3 C 3 — L (ill-a + CAV2 :3; + CBV2 :33 (159) AB 82 Cnv 82 C V 82 C 3 C 3 J v= _ L (3“1 + CAV1 aul + CBVl 3“2) B BA 82 C W 82 C v 82 C 3 C 3 - L (8u2 + CAV2 aul + CBV2 8u2) (160 BB 32 C V 82 C V 32 C 3 C 3 Carrying out the multiplications in equations (159) and (160), equating coefficients to equations (142) and (143) and solving for LAB and LBA gives c c V c C‘V .. —_1_1.3__- 2A : LAB ‘ LBA ‘ ' oln (l‘VlCA) ' 02n (1’ 2°B) c c V A B 3 — — _ O3” (VlcA + V2CB 1) C 12 — — + g——fi(l-V12CA)(l-V12CB) 12 C13 — — + 8——fi(l'V13CA)(l'V13CB) (161) 13 [’Y‘a‘ 1nus, hydrodynamic theory predicts the validity of the Onsager Reciprocal Relation for this associated system, under the assumptions of constant volume and that the volume of the dimer is equal to the sum of the volumes of the component monomers. EXPERIMENTAL Tracer Diffusivities Tracer diffusivities for this work were measured by means of the capillary technique, as modified by Wirth [8]. In the basic capillary technique, a capillary of known length, with one end closed, is filled with a solution containing tagged molecules of one component. This capillary is then immersed in a relatively large volume of a solution with the same chemical composition, but containing no tagged molecules. Diffusion is then allowed to proceed for a period of time, after which the capillary is removed from the bulk solution and emptied. The relative amounts of tagged material before and after the experiment are determined. The boundary value problem for diffusive transfer from the capillary is then solved to give the change in the concentration of tagged molecules as a function of time, capillary dimensions and tracer diffusivity. Since the time and capillary dimensions are known, the tracer diffusivity can then be found from the change in concentration of tagged material. Ordinarily, the molecules are tagged with a radio- active isotope. In this case it is easiest to measure 66' 67 the total amount of radioactivity present rather than the concentration. This presents no difficulty, however. There are four main sources of error in this basic technique: a. Inaccuracies in determining precisely the amount of radioactivity before and after diffusion. This is particularly important in determining the initial count rate. b. Proper maintoinance of the conditions of the boundary value problem during the experiment. This means that there must be no convective mixing within the capillary during the experiment, and no material may be transferred by any means other than diffusion. c. Immersion effects. Material must not be washed out the end of the capillary by the turbulence created in the bulk solution when immersing the capillary or removing it at the end of the experiment. d. Convective transfer during the experiment. Convection near the end of the capillary must be strong enough to maintain the boundary condition of zero con- centration at the end of the capillary. Yet it must not bee so strong that it washes material from a segment of the capillary, thus effectively shortening the length of the capillary during the experiment. The two latter problems are due to the open end of the capillary, and are difficult to correct as long as the 68 end is open. The other two problems are generally less serious. To correct for (c) and (d), Wirth covered the open end of the capillary with a very thin (0.007 in.) porous glass disc. This changed the boundary value problem, introducing a resistance term at the end instead of a constant concentration. Since material could diffuse through the disc, but could not flow through, this pro- cedure effectively eliminated convection from the capillary. However, since the resistance of the frit had to be calibrated by measuring a known diffusivity, a new possible source of error was introduced. The error due to (b) can be largely eliminated by making the bulk solution slightly less dense than that in the capillary. Then, as diffusion proceeds, a density gradient is established, which tends to eliminate convec- tion within the capillary. This unfortunately introduces the possibility of some mutual diffusion occurring along with the tracer diffusion. It has been shown by Van Geet and Adamson [31] that if the concentration difference between bulk and capillary solutions is greater for tracer diffusion than for ordinary diffusion, the tracer flux will be much greater than the ordinary diffusive flux. Since the difference in the concentration gradients was quite large in these experiments, the author feels that any error 69 introduced from simultaneous mutual diffusion will be covered by the calibration of the resistance of the glass frit. Inaccuracies introduced by (a) can be decreased only by careful experimental technique. Wirth's [34] original technique was modified in this study only slightly to improve the accuracy of the initial count. The procedure for an experimental run is given in some detail in Appendix A. Calculation of Tracer Diffusivities Let the closed end of the capillary be designated z = O, and the open end be designated 2 = L, the length of the capillary. If transfer of material within the capillary is only due to diffusion, the well-known diffusion equation holds. The initial concentration is constant throughout the capillary. At the closed end, a material balance will show that the concentration gradient must be zero. If, at the Open end there is a constant resistance to flow, the following boundary value problem holds: 1 (162) * 0C, 5.C. 1; 2‘ - O for 2 = C t > 0 OZ 3 — x r— 2 D * as. N * DOG. : - l _. .L r‘ 9" ') _ - qu v n * A* v- 1.C. ~C, = C, for t - 0, o 5.2 5.2 1,0 x a , wnere C is Cu concentration 01 tagged component i, n * . t lS the time, 2 is the distance coordinate, C,O is the *3 o o a - * a a n 4- . initial value of C. and n is_the constant resistance to n -‘ O A ‘r‘ ’- ‘ 'r I"; '3‘. «.‘f‘ A r‘ ‘2 q - .Y transfer from the open end 01 Cue Capillary. This boundary value problem can be easily solved by separatioh of variables, to give the concentration as * a v- C, m sin 1,2 2 * .L r" L; \ ‘1 ———— = 22 ' - . ~- - ex —A D 2 “OS A 2 * _,‘A C + Sin A 2 cos 1 2 p( n i ’° n ‘ C. n-i n n n 1,0 ' (164) where Ar is given by the solution of A 'A " "“ * A CO: K 4." - :4 n I L n A detailed solution of the boundary-value problem is given in Appendix III of reference [343 F.) If equation (164) is integrated over the length 0 "3 the capi lary, a very use-ul ratio results: 71 0* m sin2 8 D *t 2‘ i _£%EKE = 22 [ . n exp(-B -—§-J] (165) Ci 0 n=1 Bn‘an + Sin 8n cos 8n) n L where an is given by the solution of * RD. _ l cot 8n - Sn L * where C is the average concentration in the i,ave * capillary, as defined by the expression I: Ci dV = * Ci V. This ratio is the ratio of the final count to ,ave the initial count measured for the capillary in the experiment. Since transfer through the frit is diffusive, as soon as the process reaches steady-state the resistance of the frit becomes inversely prOportional to the dif- fusivity: R-—d—'g’ Di The constant of proportionality a depends only upon*the RD i is pore geometry of the frit. Therefore the group dependent only upon geometry of the experimental apparatus, and can be determined by some calibration technique. 72 ‘ From equation (165) a plot can be made of * * * Ci av Di t RDi ——f——E vs 2 for several values of L . Then i,o L from an experiment with a chemical whose self—diffusivity is known, the value of R may be determined. Wirth did * RD. this, using carbon tetrachloride. He determined that ~31— ’1':- had a value of 0.012, with a variation from capillary to capillary which would lead to a .7% maximum variation in measured tracer diffusivity. Consequently, the value of * RD L was taken as 0.012. * 3|: * C, D t RD Using a plot of ‘ avg vs 12 for L = .012, L the tracer diffusivity is determined from the count rates resulting from each experiment. The ratio of final count * C rate to initial count rate is equal to -£¢3KE. The value C it * 1,0 Ci av Di t of -—:¥—5 fixes the value of L2 . Since t and L are i o 3 * known, the diffusivity Di is easily calculated. Mutual Diffusivities Mutual diffusivities were measured in this labora- tory by means of a Mach-Zehnder [4,5] interferometer. This instrument is shown schematically in Figure B-l of 73 Appendix B. A collimated monochromatic light beam is split into two beams by a half—silvered mirror. One beam is passed through a solution in which diffusion is occurring. The other beam is passed through a reference solution in which there is no concentration gradient. When the two beams are recombined, if the optical path lengths are only Slightly different, interference pro- duces h) fringe pattern. Since the optical path length is dependent upon the refractive index of the medium, it can be shown that the fringe pattern formed by the recombination of the beams ve index vs position in the *Jn represents a plot of refraCt diffusing system. The refractive index is in turn'related to the composition of the system. By photographically H, be Fl ’53 d "f recording the changes ringe pattern with time, 1 the changes in compositior (and hence the diffusivities) .. can be determined. a he diffusion cell was constructed so that a step- change initial condition could be approximated, and so that diffusion would be one—dimensional along the vertical axis. The cell was filled from the bottom with the denser of two solutions varying slightly in composition. The less dense solution was slowly introduced down the wall of the cell, forming a layer above the denser solution. When the cell was full, and had reached e uilibrium temperature in the interferOJeter thermostat b , 71; solution was removed slowly through slits on opposite sides of the cell, and replaced in the cell from top and bottom; After some time, a steady state was reached, in which the solution above the slit was of one composition, and that below the slit of another. The thickness of the boundary determined how closely the step-change approxi- mation was obeyed. In practice, the boundary could be made small enough that it closely approximated the con- centration distribution after a few seconds of diffusion from a true step—change initial condition. The flow into and out of the cell was stopped, and free diffusion from this initial condition occurred, which was followed photographically. Details of an experimental run are given in Appendix B. Analysis of_Results of Binary Mutual Diffusion Experiment The problem of one-dimensional free diffusion in an infinite medium is an old one. It was solved by Wiener [33] in 1893. If the initial position of the boundary is designated as 2 = O, and the initial distribu- tion of concentration gradients is Gaussian, the solution for the gradient in terms of position.and time is given by AC dC o ' 2 _ g —- exp(_z /u D t) (166) dz 2’?D_—t' AB AB - 75 provided the diffusivity is constant. If the initial concentration difference is small enough, both these conditions will be met. Furthermore, the refractive index difference may be considered a linear function of the concentration difference for small AC. This solution may then be written 3% = _.___A_r_1___ exp {—22/4 DABt) (167) 2/NDABt This may be integrated to give the refractive index difference between the points z = 0 and z = 2. n2 - no 1 Ar = 5 erf (2/ V4DABt ) (168) Solving this for 2 gives n z = VED, t erf.l (Z—E————9) (169) The photographic image can be considered a plot of refractive index vs position in the cell. The total refractive index difference between any two points is proportional to the number of fringes crossed by a vertical line between the two points. The fringe number can then be used as a measure of refractive index. Call the fringe number of a reference point in the straight line portion of the photograph fringe number zero. The 76 .............. a J 77 total number of fringes crossed by the vertical line is J. The fringe number of the point z = O is J/2, from the choice of coordinates in the boundary value problem. The difference between the refractive indexes of the point 2 = 23 and 2 = 0 is given by 0 I ’O = . .-%5_— “J (170) The distance between any two fringes numbered 3 and k is n1 — n n.-n z - 2 = /4D Lt [erf-1(2—JEE——9) - erf 1(2‘éfifi"'9)3 (171) The actual distance in the cell is not the same as that measured on the photograph, but differs by the magnifica- tion factor of the camera: 2, = M2. ‘ (172) d J t where aj is the measured distance, 23 the true distance and M the magnification factor of the camera. Com- bining equations (170), (171) and (172) leads to the desired result: 4M2D t = E 23 - 2k 12 (173) AB err‘l(—1——2 'J) - err'l(———2k‘J> 78 The true time of diffusion is not the measured time, since the initial boundary was not a perfect step-change: where tm is the measured time, t is the true time, and t the true initial time. If the right side of equation 0 (173) is plotted vs tm, the slope of the line will give the diffusivity, and the intercept the true initial time: 2 slope 4M DAB (174) intercept - 4M2 D t AB 0 Calculation of Binary Mutual Diffusivities The photographic plate was measured by a microsc0pe with a traveling eyepiece, capable of measuring down to 0.0001 cm. The total number of fringes was counted and recorded. Then a set of ten fringe numbers was chosen, five of which were higher than J/2 and five of which were lower. These were chosen so as not to extend into the region of curved fringes near the edge of the diffusion boundary. These were paired, and the right side of equation (173) calculated for each pair. For each exposure, 23' and Zk’ were measured for each pair of fringes. From the measurements and the previous calculations, five values of the right side of equation (173) were calculated and averaged. The average 79 values were then plotted against tm as described above. A straight line was fit to these points by the method of least-squares. If the correlation coefficient was less than .995, the run was rejected (although generally it was above .999). Otherwise, the value of D was AB calculated from the least-squares slope and intercept. Analysis of Results of Ternary Mutual Diffusion Experiment Fujita and Costing [13] have shown that the ternary diffusivities defined by equations (98) can be determined experimentally from knowledge of the behavior of the refractive index gradient curves as diffusion proceeds. Their method involves measuring the second moment and the height-area ratio of graphs of %% vs 2 at several times during the experiment, and from these determining the reduced second moment and the reduced height-area ratio. This is done for several different initial composition differences, and the graphs of reduced second moment and reduced height-area ratio vs refractive index fraction are then used to calculate the diffusivities. A typical plot of refractive index gradient vs position is shown on the following page. Here 20 is the centroid of the curve, and 22 is the maximum value of d2 max dn 32’ which is at the centroid for Gaussian curves. 80 The centroid is defined by if” 2(33) d2 (175) ° f°° (%)d2 -00 It can easily be seen th .at the denominator is equal to the total refractive index change across the boundary, which is proportional to the number of fringes 0. Therefore 2 = .l___ foo 2(92) C12 (176) 1 d2 where A is the proportionality constant. The second moment is defined by M(Z-Z >2 (g 2) dz 1 m a = x3 I cog—h z)dz (ca—20>2 <3 3) dz (177) 81 and the total area under the curve is co An = f (331—21) d2 =)(J (178) area .CD The refractive index gradient can be determined from the photographic plate as follows. Near the center of the boundary, where the fringes are almost straight, %% can be approximated from the distance between two fringes: 8n (a ) A 02 = (179) m 2 = ZJ+1 + J ZJ+l - J In the curved portions of the pattern near the edges of the boundary, this approximation does not hold, and the value of g?“ must be determined by measuring the tangent to the curve: d2 dy dy QB_=9£a%—Y.=9£tane (180) m m The value of gg-can be determined by measuring the distance between two fringes in the y direction: n - n A(j+1 - J) A (.12. = 1+1 41 = = (181) dy yj+l-yj yJ+l-yj yJ+l-yj Equation (180) can now be written 82 tan 6 (182) Q: N % L1 + H I V C.» €HV where A and w are proportionality constants representing the change in refractive index per fringe in the z direction, and the distance between fringes in the y direction respectively. The distances measured on the photograph are not true distances, so equations (176) and (177) must be corrected for the magnification factor of the camera: <—) dzm (183) dzm (184) Note that in substituting for %%— the constant A cancels. m The reduced second moment is defined by E? l\) D2m = 2? . (185) and the reduced height-area ratio is defined by - 2 2 ' le), (AJ) D = = (186) A 1m: {3-3 12 41rtM2 [z A __ z 32 max » 3+1 3 max 83 The measured time is not the true time, but t = tm + tO so these can be written 2 _ —2 ‘ D2m m + D2m t0 (187) J2 = D t + D t (188) “WMZ E l - z 32 A m A 0 3+1 J max m D2m and DA can be calculated by plotting 5; and the left side of equation (188) vs tm. The slopes will be D2m and DA respectively, and the intercepts will give the true initial time. The left side of equation (188) and m _3 can be calculated from measurements of the photographic 2 plate. This would ordinarily be rather difficult, but a computer program has been written for this. This program, the data deck structure, and the procedure for measuring the plate are given in Appendix C. Sample refractive index gradient curves and plots of IEE-Ivs tm are given in Appendix D, with the experimental results. In a ternary system, the refractive index can be expressed as a function of the concentrations of any two components, for example n = n(CA,CB). This in turn) can be expressed as a Taylor series expansion in terms of CA and CB. For small enough concentration differences, the higher order terms of the expansion can be dropped, and 84 An = RAACA + RBACB Defining the refractive index fraction of component A by. R AC A A a = (189) A RAACA + RBACB we see that d + d = l (190) The values of RA and RB can be determined from measurements of An at several different AC's by a least- squares technique. Since accurate direct measurements of An require relatively large concentration differences, the preferred method is to determine the values of RA' and RB. defined by the equation ! l J = RA ACA + RB ACB (191) I I where RA = ARA and RB = ARB. This will allow smaller composition differences to be used, since J can be measured more precisely than An. . -AC By equation (191), we see that if J A is plotted vs. AC ' ‘ -7f3 , the result should be a straight line, with slope 85 R! fiTE-and intercept §%-. As can be seen in figure D-10, A A this is true for the ternary system chloroform acetone- benzene which will be investigated in this work. Note R R that .75.: fig and that only this ratio is needed in R B B defining the refractive index fraction. Fujita and Gosting [13] have shown that plots of reduced second moments and the reciprocal of the square- root of the reduced height-area ratio vs. refractive index fraction of one component should be straight lines: Dzm = 82m GA + 12m (192) l — WA ‘ SA 0‘A * IA (193) where 32m and SA are the slopes and 12m and IA are the intercepts at dA = 0. Their proof is based on the same assumptions which have been made here, and which hold whenever the concentration differences are small. For convenience in notation, define the intercepts at dA = l by the expressions L2m = I + 32m ~ (194) LA = 1A + sA (195) where reference [23]. diffusivities is given in Appendix C. 86 Their expressions for the diffusivities are L L2m S2m IA 2 D = _ JDljj + L2m J D131 + SA AA 32 m t I2m S2m LA D = - DlJ + 12m DiJ + sA BB 32 m AB 31 i2m ”33’ R1 DBA 2; (”2m ' DAA) [DH]15 is the root of the cubic equation 3/2 lDljl + (12m ‘ IA ‘EZQ [D A detailed derivation of these equations is given by Fujita and Gosting, or may be found in Appendix II of 1 (200) for [Dul’2 and then calculates the four ternary the plots of equations (192) and (193) for the system studied in this work may be seen in Appendix D. (196) (197) (198) (199) (200) A computer program which solves equation The linearity of RESULTS AND DISCUSSION The following systems were studied experimentally in this work: 8.. Tracer diffusivity of 2—butanone in the system 2—butanone - carbon tetrachloride Tracer diffusivity of p-benzoquinone in the system p-benzoquinone - benzene Mutual diffusivity in the system p-benzoquinone — benzene Tracer diffusivity of diethyl ehter in the system ether - carbon tetrachloride Mutual diffusivity in the system diethyl ether — carbon tetrachloride Mutual diffusivity in the system carbon tetrachloride - chloroform Mutual diffusivity in the system benzene - chloroform Ternary mutual diffusion at equimolar composition in the system acetone - benzene - chloroform Experimental results, and intermediate determinations for the ternary system, are given in Appendixes D through F. The discussion of results will be organized by type of diffusivity studied, rather than by composition of the systems studied. 87 88 Tracer Diffusivities Equation (82) predicts that for a self—associating component in a binary system, the product of the tracer diffusivity and the viscosity should have its highest value when that component is very dilute. As the con- centration increases, the ratio of polymers to monomers present in the system will increase, and the tracer diffusivity will decrease. In binary system with cross-association the situa- tion is slightly different. As can be seen from the equilibrium constant expression, ‘the ratio of dimers to monomers of component A is pro- portional to the mole fraction of component B. Hence ‘the percentage of A molecules which are tied up in the (dimers is highest when component A is very dilute, i.e. XMhen xB + 1. Therefore, the tracer diffusivity- ‘Viscosity product of a component is lowest when that Ccnnponent is extremely dilute, and increases as the Concentrat ion increases . * Figure 1 shows the variation of the DAn product for associating components in three systems. In the Systems ethanol - carbon tetrachloride and acetic acid carbon tetrachloride hydrogen bonding is quite strong 7, dynes D n x 10 89 Mole Fraction Component A A Figure l.-—Din vs Mole Fraction for Associating Components. | | O l | | 8 \\ 2 o-P°\ \ \\ \ \\ \\‘ \0\ O\ \ \ C O E . \\ \ \\ \\\ \\U\. .~"" ‘::o——"O————. 1.0... .M A. -ir—- Diethyl Ether (A) ——E+—- Chloroform (B) -—O-- Ethanol (A) - CClu (B) -.—o—- Acetic Acid (A) - CClu (B) 0 1r 1 1 1 4 o .2 .4 .73 .8 l. 90 between ethanol molecules and acetic acid molecules. Carbon tetrachloride, on the other hand, is probably quite inert. The curves for these two systems have * exactly the shape predicted by equation (80). The DAn product is highest when x + 0, and decreases as A x + l. A In the system ether - chloroform, spectroscopic evidence [15] suggests that hydrogen bonding occurs between ether and chloroform to form dimers with the form Presumably, steric hindrance prevents the formation of larger ploymers in this system. Ether—chloroform is therefore a cross-associating system. The curves in Figure 1 agree with the predictions of equation (55). s in component is extremely dilute, and increases as the The D product for each component is lowest when that concentration increases. I Equation (45) predicts that for a system where neither component associates appreciably, the tracer diffusivity - viscosity product will be a constant independent of composition. Figure 2 shows the tracer diffusivity - viscosity product for several systems in 91 Mole Fraction Component A * Figure 2.-—Din vs Mole Fraction for Nonassociating Components. :-—n———n-- D--n———U— —O--———g-9—-- —--§ (.1 V' 0 V A Q a _ f ’ C p L 4. .I I; I I- .01 ——€+—— Ethyl Iodide (A) [7] __.__. Butyl Iodide (B) [7] ——4+—— Benzene (A) [34] -—0—- Carbon Tetrachloride (B) [34] ——a——- Chloroform (A) [34] ——1F—— Carbon Tetrachloride (B) [34] t i ‘T : a 0 0.2 0.4 0.6 0.8 1.0 91 * Figure 2.--Din vs Mole Fraction for Nonassociating Mole Fraction Component A Components. UJ _ T’ x: ngL. 11.7:21. ii: at IL’E 1p m“ ! E3 1.0+ >4 C * c: -—£+—- Ethyl Iodide (A) [7] -—4F—— Butyl Iodide (B) [7] ——<+—— Benzene (A) [3“] -—0—-— Carbon Tetrachloride (B) [34] .——a—-— Chloroform (A) [34] ——1F—— Carbon Tetrachloride (B) [34] O i f t : a 0 0.2 0.4 0.6 0.8 1.0 92 which there is good reason to believe that there is no association. The normal iodides for instance are essentially non—polar, saturated, and contain no groups active enough to form hydrogen bonds. Carbon tetrachloride is also non-polar, and the electron clouds of the chlorines are quite inert to hydrogen bonding, even in an electron-donor capacity. As predicted by equation (45) the D:n products for these systems are straight lines. Furthermore, the D:n product for 0C1“ is the same for all the systems given here. This can be taken as supporting evidence for the assumption that diffusing species behave like particles flowing through a continuous medium. The diffusion process is influenced by the viscosity of the medium, but not by the character of the molecules which comprise the medium. Spectroscopic studies have suggested that ketones, 'being polar molecules, undergo some dipole-dipole inter- actions which lead to the formation of self-polymers in ssolution. Anderson [1] successfully applied the self- (dimerization model to explain the positive deviation :from Raoult's Law in the system 2-butanone - carbon ‘tetrachloride. He then used equation (92) to fit experimentally measured mutual diffusivity data with excellent results. Wirth [34] later measured the tracer 93 diffusivities of carbon tetrachloride in this system, confirming the fact that carbon tetrachloride did not associate. In this work, the tracer diffusivities of 2-butanone were measured, hoping to verify the self-association model. ‘Experimental results for this system are given in Appendix E, and shown in Figure 3. The tracer diffusivity - viscosity products for s in products are constant throughout the entire concentra- this system are shown in Figure 4. Since the D tion range, it must be concluded that there is no associa- tion in this system, at least with respect to diffusion. The dipole-dipole interactions observed spectroscopically apparently are not strong enough to hold the dimers together against the shear forces they presumably undergo while diffusing. This would indicate that the inability of equation (88) to predict mutual diffusivities in this system is not due to the formation of polymers. A possible cause would be inaccuracies in the vapor-liquid data in the literature. The system clearly warrants further study. Spectroscopic studies have shown that highly con- jugated molecules with electron withdrawing groups adjacent to the conjugation, such as cm2/sec 5 or DAB x 10 , D l\) O l 94 3.0- 1.0m . D ex erimentall AB, p —--- Hartley—Crank Equation * D DA 2—butanone * 0 DB carbon tetrachloride [34] O : : : : 4 O 0.2 0.4 0.6 0.8 1.0 Mole Fraction 2—Butanone Figure 3.—-Mutual and Tracer Diffusivities for the System 2—Butanone - Carbon Tetrachloride at 25°C. 95 .0" t 0 a O- O .0“ * -——€F——- DAn 2-Butanone * ———O———- DB” Carbon Tetrachloride [34] 0 i i f, i as 0 0.2 0.4 0.6 0.8 1.0 Mole Fraction 2—Butanone Figure 4.-—Tracer Diffusivity — Viscosity Products for the System 2-Butanone — Carbon Tetrachloride at 25°C. "flu. O i\‘02 x/ \\\., ¥ / , ~. g E p-benzoquinone * \_, 1,3,5-trinitro 4,// Nd;\// N02 benzene '0 can undergo charge-transfer interactions with donor molecules, usually aromatics, which stabilize the structure ’::\C 0:," ,eeo It has been established that p-benzoquinone will associate to form dimers in solution with aromatics [2, 11] and equilibrium constants have been measured for several of these systems.- In an effort to experimentally verify equations (150) through (151) for a ternary system with competing equilibria, it was decided to study the system p-benzoquinone — benzene — p-xylene. It was expected that the quinone would form dimers with benzene and xylene, and that no other associations would occur. Mutual and tracer diffusivities were studied for the component binary system quinone-benzene. Since quinone is only slightly soluble in benzene at 25°C, the results cover only the solubility range. Experimental results are given in Appendix E and shown in Figure 5. 2 r DAB x 105, cm /sec l-' 97 .0+ 0 .0" ._49__ D AB . * DA * O DAn 0 L l l A l l 1 l I I 0 .01 .02 .03 .04 .05 Mole Fraction p—benzoquinone Figure 5.--Mutual and Tracer Diffusivities in the System p-benzoquinone (A) — benzene (B) at 25°C. 98 The tracer diffusivity-viscosity product for quinone is also given in Figure 5. The variation of this DZn product should be that predicted by equation (55), since this system is considered to have only cross— association. That is, the DZn product should increase as the concentration of quinone increases. As can be seen in the figure, it does not increase, but decreases instead. This was interpreted as some sort of interaction leading to self-association of quinone which masked the effect of the cross-association. The change in the D:n product for a small change in concentration is much greater for self-association than for cross-association. It is possible that if the concentration of quinone could be increased, the cross- association effect would again become predominant. In any event, the associations present are too complex to be treated by the equations developed here, and work in the ternary system was not carried further. Tracer diffusivities were measured for ether in the system diethyl ether — carbon tetrachloride across the entire composition range, and for carbon tetrachloride at the two endpoints. Results for this system are given in Appendix E, and shown in Figure 6. There were con- siderable experimental difficulties in working with this system, due to the high volatility, the low viscosity, and the surface-wetting characteristics of solutions with 4.; IA 99 Figure 6.-—Mutual and Tracer Diffusivities in the System Mole Fraction Diethyl Ether Diethyl Ether — Carbon Tetrachloride at 25°C. 8.0.- 7.0._ 6.0__ O + (D U) \. (\l S 5.0,, Lhc)fl H + x B Q “.0‘- + 2.. 0 2|: Q 3.0-r + 2.0., g 7 DAB 1.0-L , ¢DA 0 : : : : 4 0 .2 .4 .6 .8 1.0 100 high ether content. As a result, the uncertainty of this data is higher than for the other systems studied, as indicated by the bars on Figure 6. It was assumed that the ether — carbon tetra- chloride system would be a simple nonassociated system, and tracer diffusivities would be as predicted by equation (45). In fact, the shape of the DZn product curve for ether is more like what one would expect for a cross-associated system. Since the carbon tetra- chloride is non-polar, and its chlorines do not form hydrogen bonds, this phenomenon is rather difficult to explain. Being somewhat unfamiliar with the mechanics of charge-transfer complexing, the author hesitates to eliminate this possibility, but it does seem unlikely. Furthermore, over a period of time the bulk solution discolored, indicating a reaction of some sort pro- ceeding. It is possible that the reaction (though not extensive and rather slow) indicates that some inter— molecular interactions were occuring beyond the usual attractive and repulsive forces. Another alternative is that the assumption of a continuous medium breaks down here. This is supported by the fact that the Dgn product of CClu changes only about 10% over the concen- tration range, while that for ether changes about 30%. Again, this system warrants further study. Investigation 101 of other properties besides diffusion might also provide some insight. Mutual Diffusivities As part of a study of the ternary system ether - chloroform - carbon tetrachloride, Wirth [34] measured tracer diffusivities for both chloroform and carbon tetrachloride in the component binary chloroform - carbon tetrachloride. These measurements showed that, as expected, both chloroform and carbon tetrachloride are nonassociated in this system. The Hartley-Crank Equation, equation (88), should describe mutual diffusion in this system. Wirth [34] measured mutual diffusivities in this system to experimentally verify this equation. He encountered some experimental difficulties, and scatter of data cast some doubt on his results. The best data he could obtain from his results, however, showed that the Hartley-Crank Equation is inadequate to describe this system. The shape of the Hartley-Crank curve is wrong when compared to Wirth's data. It was of interest then to attempt to duplicate Wirth's data, to determine whether thel discrepency is truly in the equation, or whether it might be in the experimental results. The author encountered less experimental difficulty in measuring this system. Experimental results are given in Appendix E, and shown (along with Wirth's 102 results) in Figure 7. As can be seen, the author's results agree quite well with the best results obtained by Wirth. The Hartley-Crank Equation is definitely inadequate for describing this system. The discrepancy in this case is very hard to explain. The assumption of a continuous medium is probably a good one, since otherwise the effects would have shown up in the tracer diffusivities as well. The activity data reported in the literature used in calcu- lating the thermodynamic correction factor appear to be quite good. The system is only slightly non—ideal, so the correction factor is not too large in any case. There is definitely no association, as can be seen from the Dzn products. This phenomenon is puzzling, and will probably require further investigation to provide an explanation. The Hartley-Crank Equation also fails in another nonassociated system, 2-butanone - carbon tetrachloride. In this case, however, the predicted mutual diffusivity curve has the correct shape, differing only in the magnitude of the correction from ideality (see Figure 3). The author suggests that this may be due to a slight- error in the activity data, from the experimental vapor- liquid equilibrium measurements of Fowler and Norris [12]. 103 2.0 O (D U) \. (\J E 0 m0 O H x ‘13; Q 1.0 "' a 0 3|: Q --0 -— DAB’ this work - 9-- DAB’ Wirth [3A] "‘a“‘ DA * ——e— DB 0 : : : t 4 0 .2 .4 .6 .8 1.0 Mole Fraction Chloroform Figure 7.--Mutual and Tracer Diffusivities in the System Chloroform (A) — Carbon Tetrachloride (B) at 25°C. 104 The author measured mutual diffusivities in the system benzene - chloroform, to further test the Hartley- Crank Equation. Experimental results are given in Appendix E, and shown in Figure 8.‘ It was expected that this would be a simple nonassociated system. Although tracer diffusivities were not measured in this system, the self—diffusivities of both components are available. When the D:n product of the pure component is compared to the D Bn product when that component is extremely A dilute, the results indicate that both components are nonassociated. As can be seen in Figure 8, the Hartley-Crank Equation again fails for this system. The shape of the curve is qualitatively correct, but the correction is again too much. The error in this case is probably too large to attribute to inaccurate activity data. Further investigation in this system is warranted, particularly measurement of the tracer diffusivities over the entire concentration range, to make certain there is no association. The author has also measured mutual diffusivities in the system diethyl ether - carbon tetrachloride. Activity data are not available for this system, there— fore it cannot be used to test the Hartley-Crank Equation. Furthermore, experimental difficulties (previously described under Tracer Diffusivities) caused 105 3.0-r / / \ \ / \ / O o 7‘ \\ / O \\ (5\\ O \ \ e 1 2.0“ 0 .' (D (I) \C (\J E 0 m0 0 r_| x m < c: 1.01 0 Experimental results ---- Hartley-Crank Equation Ideal system 0 .2 .4 .6 .8 1.0 Mole Fraction Benzene Figure 8.—-Mutual Diffusivities in the System Benzene (A) - Chloroform (B) at 25°C. 106 considerable scatter in the data, especially near the center of the concentration range. Experimental results are given in Appendix E, and shown in Figure 6. As a measure of the uncertainty, the standard deviation of the data are listed in Table E-3, and are indicated by bars on Figure 6. Mutual diffusivities were also measured for the a system p-benzoquinone - benzene, up to the solubility . limit. Results are given in Appendix E, and shown in Figure 5. Again, no activity data are available. The 'fi complex associations present in this system precluded testing the hydrodynamic equations in any case. Error Analysis--Mutual and Tracer Diffusion Bidlack [27] and Kett [16] found that for the instrument used in this study, the experimental precision was i1% for volatile liquids such as used here. This was based upon several runs on aquaeous solutions of sucrose. These runs were compared to determinations made by Gosting [17] on the sucrose - water system, with agreement within i0.5%. They therefore conservatively estimated the precision of the method using this inter- ferometer as 21%. This author accepts the figure of i1% for the precision of the method and the instrument. Since the experimental procedure was not changed from earlier 107 procedures, the only remaining source of error would be that introduced by the experimenter. The author takes the agreement between his data and that collected by Wirth for the system chloroform - carbon tetrachloride as evidence that no systematic error has been introduced which would give consistently high or low experimental diffusivities. At several compositions in the systems chloroform - carbon tetrachloride and benzene - chloroform mutual diffusivities were measured two or more times. At all these compositions, the values obtained agreed within 12%, and in most cases within 21%. The author takes this as evidence that random error introduced by the experi- menter is within the precision specified for the method by Bidlack and Kett. The experimental precision for the studies in this work will therefore be taken as 21%. This figure does not apply to the system ether - carbon tetrachloride, because of experimental diffi- culties felt to be inherent in this system, which have been discussed previously. In this system four or more, determinations were made at each composition, and averaged. The averages are reported, along with the standard deviation of the data, in Appendix E. Wirth [34] has shown that the modified capillary technique used in this study has an experimental pre- cision of 22%. This was shown.by comparing tracer diffusivities at extreme dilution with mutual diffusivi- ties extrapolated to zero concentration (which must be 108 identical according to equations (91) and (92)), and by repeated runs for the same composition and comparing the reproducibility. The author accepts this as the experimental precision of the method using this experimental apparatus. To determine the amount of error introduced by changing the experimenter, the author reproduced the self diffusivity of carbon tetrachloride (which Wirth used for calibrating the cells), with a deviation of about 21%. Further evidence is the comparison between mutual and tracer experimental diffusivities at extreme dilution in the systems 2-butanone - carbon tetrachloride, and p-benzoquinone - benzene. The author concludes that the experimental error introduced into the method by changing the experi- menter is within the experimental precision reported by Wirth. Tracer diffusivities reported here are therefore assumed to be accurate to within 22%. Again, this does not apply to the system ether - carbon tetrachloride. Experimental difficulties here make the results somewhat more uncertain. Determination of precision is rather difficult. The values at extreme dilution are within 25% of mutual diffusivities (which are themselves uncertain). The author estimates tracer diffusivities in this system are accurate within 25%, and are so reported in Appendix E. ‘4 109 TernaryADiffusivities Ternary diffusivities were measured in the system acetone — benzene - chloroform, at an average composition of x = x = x = .333. Typical curves showing the A B C change of the refractive index gradient throughout the run are given in Figure D-l. Values of D2m and DA and. the time correction factors are given in Appendix D. The linearity of equations (187) and (188), which are used to evaluate the reduced quantities, can be seen in Figures D-2 and D-3. and V DA vs refractive index fractions for all three independent choices of components Plots of D2m are given in Figures D—4 through D-9. (Ternary Dif- fusivities can be expressed in three different ways, depending on which components are considered, i.e., D D D D or D D D D or D D BA’ BB AA’ AC’ CA’ CC BB’ As will be pointed out later one set of AA’ AB’ BC’ D D CB’ CC' diffusivities may be more advantageous in testing the hydrodynamic model and Onsager's Reciprocal Relation than the other two sets.) The slopes and intercepts of these lines were determined by a least-squares analysis, and are given in Appendix D. These slopes and intercepts were then used with the computer program given in Appendix C to determine the diffusivities, which are given in Table 2 for the optimal choice of components for this system. 110 TABLE 2.-—Predicted and Experimental Ternary Diffusivities in the System Chloroform (C) - Acetone (A) - Benzene (B) at 25°C. Predicted EXperimental 95% Confidence O A ‘- n / >30) DCC 5-674 .5070 11.02 P U) fi\\ D "'4 ‘ + 1 >04. A — oh I. — o H E CA 99 1 55 7 5 U) 0 A a , D - .942 — .80 _ .815 AC (Hm e40 -, 3,3 DM 2.515 1.74 21.52 x I‘ll“. H m' C) U) ”€98 L12 5L3 ET -3.561 2.337 29.88 oz) 8:: L o . E94 21 2 . C: O l l on) n D... characteristics. As can be see. in Figure 9 and Figure 0 O 10, the scatter in the data is large enough to mask any curvature due to association. It is probable that there may be some cross—association in the binary acetone - chloroform, but the other two bi.aries are felt to be In any event, the curvature due to association is not likely to be extreme, since the end points vary by only 25% in the binary systems. It was therefore assumed that the ternary system could be considered a dynes 3 D n x 10 111 2.0» o D ' U o [3 O D O E: ’ ' ' ' . l o o E Q U 0 o 3 O o O o 1.0. * DAn, Acetone, McCall and Douglass [25] * . DAn, Acetone, Anderson and Babb [l8] * D DBn, Chloroform, McCall and Douglass [25] * O DBn, Chloroform, Anderson and Babb [18] O s i ; : 1 O 0.2 0.4 0.6 0.8 1.0 Mole Fraction Acetone * Figure 9.——Din vs Mole Fraction in the System Acetone - Chloroform at 25°C. 112 2.0» <9 0 o 8 o o o o 4 : o o o o ‘ . U) (1) 5:: >5 '6 n 1.0l [x O H N 3k .C Q 9% 0 DAn, acetone, McCall and Douglass [25] 4(- ‘ [th benzene, McCall and Douglass [25] 0 ; : s + : o 0.2 0.14 0.6 0.8 1.0 Mole Fraction Acetone * Figure lO.——Din vs Mole Fraction in the System Acetone — Benzene at 25°C. 113 non-associated system, at least as an approximation. Equations (lll) through (114) can'then be used to predict the ternary diffusivities. These equations depend on the assumption that the molar volumes are constant. Since the composition differences within the diffusion cell were kept very small, the author feels that this assumption has been met experimentally. Activity data for the three binaries were fit to Margules equations and then combined to give ternary isothermal activity data (as discussed on pages 18-21). The friction factors were taken to be the weighted averages of the friction factors at the end-points in the various binaries. Viscosity was measured with a Canon-Fenske viscometer. These quantities were then used with the computer program in Appendix C to predict the ternary diffusivities, which are given in Table 2 along with the experimentally measured values. The 95% confidence levels of the measured data, which will be determined in the next section are also listed with the experimental data. It can be seen that the predicted values of the diffusivities fall well within the 95% confidence limits. The author therefore feels that the experimental determinations support hydrodynamic theory. The predicted values of the phenomenological coefficients L and L2]- are also 12 within the 95% confidence limits of the measured values. A...— ILA... _ . __. FJ F" .t‘: Within experimenta' precision, this can be taken as empirical verification of the Onsager Reciprocal Relations. The 95% confidence levels are quite large for this set of experimental data. This will be discussed in the next section in detail. It will be shown that the values of the cross—coefficients are extremely sensitive to the experimentally measured intercepts, and that a very slight error in determining the intercepts can lead to an extreme error in the cross-coefficients, as well as a significant error in the main coefficient. From this sensitivity analysis, and a consideration of theexperimental data, suggestions will be made for modifying this procedure. The author believes that through a thorough investigation of certain factors leading to experimental uncertainties in the present method, techniques can be developed which will allow this method to give 95% confidence levels within 20% or so for the cross-coefficients. This would then give a rigorous test of the hydrodynamic model and the Onsager Reciprocal Relation. Error Analysis—~Ternary Diffusion Since ternary diffusion has been studied in so few systems, and since the time involved in making a complete determination at any one composition is so long, 0 determination of experimental preCision by comparison to 115 published data is virtually impossible. The number of reference systems is also quite small. In the past, the usual procedure was simply to determine the experimental uncertainties in the measurement of the various slopes and intercepts used in calculating the experimental ternary diffusivities. Kett [23], for instance, reported 95% confidence levels of a percent or so, and concluded (implicitly) that his experimental diffusivities were of the same order of precision, a percent or so. In fact, this confidence level would lead to a much larger confidence level for the cross-diffusivities than he implied. Later evaluation of his data showed that the confidence levels were actually somewhat larger than he reported, which would lead to even more error in the diffusivities. The errors in the cross-coefficient resulting from a 1% error in the value of the intercepts can be as large as 200%. This is because the calculation of the cross-diffusivities involves subtraction of two rather large numbers to obtain a small one, so that uncertainties in the larger numbers are greatly magnified. Furthermore, errors in the main coefficients as large as 20% can result from a 1% error in the intercepts. It would seem quite worthwhile then to look at the sensitivity coefficients of the ternary diffusivities, which relate the change in a calculated diffusivity to a change in a measured parameter. 116 If a dependent variable is a function of several independent variables, the functidnal form usually involves several arbitrary parameters. The values of the parameters for a given physical system are usually determined by measuring the dependent variable at several values of the independent variables. The experimental "best" values of the parameters are then assumed to be those which give the least-square error when fitted to the data of the experimental measurements. If the equation is of the form y = f(al a2 ... a x2 ...x ) (201) where the ai's are the parameters and the xi's are the independent variables, then the sensitivity coefficients are defined by a? iaj = l: 2: 00- n ya = (a; )a x K = l, 2, ... m (202 i i J’ K j # i The sensitivity coefficients measure the change in the dependent variable with a change in the parameters, and are themselves functions of the independent variables. Ternary diffusivities can be treated as dependent variables whose values depend upon independent variables (component mole fractions, temperature and pressure) I and the parameter I SA’ and S2m‘ The functional A’ 2m’ form of this dependence is given by equations (196) ....J ...! N through (200). Since the functional form is rather complex, analytical evaluation of the sensitivity coefficients is rather difficult. They can easily be determined numerically with the aid of a computer, however. A small change is made in the value of one of the slopes or intercepts, keeping the others constant, and the change in the diffusivities is noted. This has been done for the system measured in this study. If the sensitivity coefficient is multiplied by bitrary parameter, a reduced sensi- p) *5 the value of the tivity coefficient may be defined s: s a. y = a. < df) (203) i i i a 2 This gives the chan~e in the diflusivities for a one— percent change in the parameter. This is useful, since if the percentage uncertainty in experimentally measured parameters is known, the uncertainty in the diffusivities can be determined. These same arguments also hold for the phenomenological coefficients used to test the Onsager Reciprocal Relations. If the experimental data are to be used to test a proposed model, it would be best if the sensitivity coefficients with respect to the measured parameters were as small as possible. In the case of ternary dif- fusivities, three independent choices of components may be made. The sensitivity of the main and cross- diffusivities will not necessarily be the same for each choice. It would be best then to choose those components for which the sensiti <; } 1 Cf '. I (D U) m *3 (D lowest. It can be seen from Table 3 that in the s stem acetone (A) - benzene‘ % (B) - chloroform (C), the sensitivity coefficients are generally lower when the set of diffusivities DCC’ DCA’ DAC and DAA is Chosen to describe diffusion. This is the basis for the choice made in ,reparing Table 2. If the 95% confidence levels for the parameters are known, then the 95% confidence levels for the dif- Flo fus vities can be approximated from the sensitivity coefficients. By assuming that the sensitivity coefficients are constant for different values of the parameters, upper a.d lower limits for the diffusivities t may be calcula 95% limits of D = (20A) a, > (.1. H Cf Hwn % H Confidence limits on the parameters I , I , S and S A 2m A 2 51 may be determined from the variances of the parameter (which are determined during the least—squares analyses) by means of a statistical t—test: .. . yF'—————- 95%, limits on a, 2 hi (205) ... e ll l+ CT (I) S1 V 119 TABLE 3.——Reduced Sensitivity Coefficients, sgi in the System Acetone (A) - Benzene (B) - Chloroform (C) for all Choiced of Coordinates. a i I2m IA S2m SA y DAA + 12 + 18 + 03 - Ol DAB — 09 - 10 +.O3 +.03 -5 x10 “ DBA + 12 +.24 0 0 0 0 DBB - 09 - l3 +.O3 + 04 ; DAA - 19 + 34 0 0 + 01 D - 08 + 12 - 02 + 02 AC x10-5 DCA + 38 -.78 0 0 - 02 DCC + 16 - 27 + 03 -.05 DBB - 21 — 36 + 03 - 04 D - ll - 23 0 0 0 0 BC xlo 5 DCB + A2 + 65 - 08 + 07 DCC + 2H +.42 O O 0.0 where t stands for the statistical parameter from the t-test, and s:(a1) is the statistical estimate of variance of the parameter a1, as determined from the least-squares analysis. The 95% confidence limits on 120 the diffusivities were determined by these formulae, and are given with the values of the diffusivities in Table 2. To provide a rigorous test of hydrodynamic theory and of the Onsager Reciprocal Relations, it would be necessary to reduce the 95% confidence levels on the main coefficients to about 10% or so of the values of those coefficients, and the 95% confidence levels on the cross-coefficients to at least 50% or so of the values of the cross-coefficients. Since the sensitivity coefficients for the intercepts are so high, it would be necessary to reduce the variances of those intercepts to within a few tenths of a percent. It would also be necessary to reduce the value of t from the t-test. Since t decreases with the number of degrees of freedom (i.e. experimental measurements made) at a given con- fidence level, a statistically large number of measure- ments should be made for every set of diffusivities desired. Since t approaches a constant value as the number of measurements increases, it becomes apparent that the sample variance must also be reduced. Concisely, this means that more precise measurements must be made, as well as more of them. This means that the spread in the data seen in l the plots of D and MDA vs a must be reduced. The 2m 121 reasons for the spread in the data are difficult to determine. One probably important cause of this spread is the quality of the initial boundary. It can be seen from Table A th t the true initial time determined from equation (187) and the 321 vs t curves is not the same e as that determined from quation (188) and the 1 - w-fi ~ 9 I 0 1 ~ vs t durves. ii the initial boundary nad been m VD A a true step change, and the timer started the instant that flow from the cell was stopped, the true initial W l—~l c1 time would have been t” = 0. 1.; s assumed that the m initial boundary is such as would have been formed by ‘ diffusion from a step—change for a short period of time. This would have resulted in a true initial boundary in which the refractive index gradient curve was Gaussian, and the true initial time would have been the same whether determined from equation (18?) or (188). It can be seen that the measured refractive index gradient curves in some runs are obviously not Gaussian. The curves are slightly skewed to one side or the other. Ternary iffusion from a step change boundary always gives skew curves, except at two times during the run when they are true Gaussian curves. However, in order M f-’ for equations (187) and (188) to apply, the true init a distributions must be Gaussian. (Note that this is the true initial di (1) tribution, not the experimental boundary.) The author believes that the error introduced by a 122 TABLE 4.--True Initial Times for Ternary Runs, as Predicted by Equations (187) and (188). to’ sec. to, sec. Run Number eq. (187) eq. (188) 60 -20 —43.8 62 -23 -37.1 63 -90 -55.2 65 -56 "”9°2 66 -13 ' 7'8 68 _52 -65.7 69 _74 -72.8 70 —58 -“7.9 non—step-change boundary can be related mathematically to the difference in the initial times determined from the two curves, and possibly one measurement of the refractive index gradient during the run. He was not able to derive such a relation, however. It certainly seems reasonable, however, to use the difference between the time corrections for the two curves as a criterion for rejecting a run. If the two initial times varied by more than a certain amount, the run would be rejected. Determination of what difference should lead to rejection will probably take considerable study and experimentation. It might also possible methods of this boundary could be worthwhile to investigate other forming the original boundary. If be improved, the better approximation to a step-change would undoubtedly lead to better .1 .- results, and less spread in the D2fi and /3_ vs a H .‘ A curves. In summary, t'e experimental results for the system acetone — benzene - chloroform support the predictions of hydrodynamic theory, within the experi— H) sion o F—‘. mental prec extreme sensitivity method presently is ’5 C) ' 5 u. m test of the ge is as valid as many The author believes enough, through fur rigorous test of th C :3 (D '5 Fl ti < (D U) Cf ' 1 0'4 the present method. Due to the of the cross-coefficients, the rigorous Reciprocal Relation, although it previously published veriiications. that the method can be improved ation, to give a .- . ‘eCiprocal Relation. C) :3 m n) 0‘} (I) *3 SUMMARY Hydrodynamic theory has been used to derive equa- tions describing the effects of composition on mutual and tracer diffusion in certain associated and non- associated liquid systems. These equations have been tested by experimental measurements of binary mutual, ternary mutual and tracer diffusivities. Tracer diffusivities have generally verified the predictions of hydrodynamic theory quite well (ether - carbon tetrachloride being an exception). As the per- centage of a component which is associated into complexes increases, the tracer diffusivity - viscosity product decreases, and vice—versa. Although many systems have been found for which hydrodynamic theory does apply quite well, three non- associated systems are presented here which seem to be exceptions. In non-associated systems, the Hartley- Crank equation should describe mutual diffusion. In benzene - chloroform and 2-butanone - carbon tetrachloride, the Hartley-Crank equation qualtitatively predicts the shape of the mutual diffusivity curve, but fails quantitatively. In chloroform - carbon tetrachloride, it fails qualitatively as well. These failures may well 124 125 be due to inaccurate vapor — liquid equilibrium data, however. In the ternary system acetone - benzene - chloroform (which is here considered to be non- associated) hydrodynamic theory predicts the ternary diffusivities, within experimental precision. Unfor- tunately, experimental uncertainty is rather large, and this may not be considered a rigorous test of the theory. Within experimental precision, hydrodynamic theory also predicts the validity of the Onsager Reciprocal Relation. The author feels that this experimental uncertainty is due primarily to the difficulty of forming a good boundary within the diffusion cell, which is critical in measuring ternary diffusivities. Suggestions are given for future investigations to reduce the experimental error and provide a more rigorous test of hydrodynamic theory. The author suggests that future work in liquid non—electrolyte di“lusion is needed in four particular areas: (a) improvement of experimental methods and techniques, so that difquivities may be measured more precisely, reliably, and hopefully more easily than is now possible; (b) more systems need to be studied to support conclusions which have been previously arrived at on the basis of a small number of studies; (c) those systems which seem to offer contradictions to hydro- dynamic theory need to be studied more carefully; (d) further theoretical work needs to be done, *0 O U) U) ...-I. 0' }__1 <<: +4. {1) I extending the principles used here to continuous assoc tion, or simultaneous self-a sociation and.cross— U) association for example. Experimental problems which lead to low precision in measurements were discussed considerably under the Ternary Diffusion error analysis. The author feels that a thorough study of the effects of different boundary conditions, different methods of forming the initial boundary, and possibly a new mathematical treatment for obtaining diffusivities from refractive index gradient curves would be a self—contained and quite worthwhile research program. 127 Besides this work on ternary diffusivities, however, the author feels that much improvement could still be made in the techniques used for measuring binary mutual and tracer diffusivities. For example, a technique which would totally eliminate convection in the capillary, perhaps by using a porous capillary instead of an open one, would be well worth developing. Or an interferometer which used a laser light source and a better set of lenses, or for which a better boundary could be established, would be worth investigating. Although the systems so far studied have generally supported hydrodynamic theory, there is not enough evidence to conclusively say that it is correct. This is a general problem in liquid diffusion--there simply has not been enough raw data generated in the past to thoroughly test any new theory prOposed except for a comparatively small number of cases. For instance, to the author's knowledge, there has never been published a complete study of any ternary system (mutual and tracer isotherms of all three component binaries, ternary mutual diffusion isotherms, and ternary tracer diffusion, to say nothing of the effect of temperature on all of these). Systems such as 2-butanone — carbon tetrachloride, and chloroform - carbon tetrachloride, which appear to contradict hydrodynamic theory (or at least present ambiguities to be resolved) should be studied more carefully. It is quite possible that studies of phenomena other than difiusion would be very useful here. It has been recently proposed that the chlorine atoms of carbon tetrachloride engage in a limited form of hydrogen bonding with alcohols, and therefore they might cause some very weak bonding effects in these to the breakdown of the Hartley- Q. systems which lea Crank equation. This seems unlikely, but perhaps spectroscopic studies directed at this particular ’0 henomenon might provi‘e some useful information. 1 Theoretical work based upon lydrodynamic theory J could be directed at finding a simplification bf equation (80) and its generalization which could be applied to tracer diffusivity in a ~ystem with extensive association, such as ethanol - hexane, r aniline - benzene. Or equations could be developed for application to systems like aniline — toluol where there is both self- association and cross-association. Equations for con- tinuous self-association as pplied to tracer diffusion pa could be related to imilar equations for mutual U) . - e H, Fl 0.. diffusion. This is again open for much investigation. 3 3 my D E Di D l 129 APPENDIX A 130 ...-1 7... x 0 F1”, Fl. :fi _ Figure n-l lS a schematic diagra capillaly. SpeCific details regarding .‘1 C) p) |. J F.J §< Cf 0 prepared gravimetif by means of a CLristian and Becker balance. i al or mole fraction on a large analy c solutions were degassed just beiore out #0 C to remove CY minutes at a U) F.)- X O Q) 'U [.1- |._J *4 (l) }_J (D U) 2' (D ’5 (D * b t I H l._J (D Q. 0' % cr f.) .L. in the screw cap. Torbal balance. an experiment from the of The capill fiI-r'fi'fiv n} 1 cirruSIoh ~ m of the diffusion dimensions, C). (I) U J }_J FS (D a) O O B ’0 O (D I J. C i’ H. O :3 (I) 2 (D *5 (D fraction, torsion Bulk solutions were prepared to within 1.005 The tracer for 15 solution.* ‘e following procedure: malleable nickel to insure and placed ary was inserted * Air bubbles coming out of solution during the run and drifting up to the glass disc were experimental difficulties. They would the capillary as they drifted upwards, changed the resistance constant in the condition at the capillary end. F“ W ..J one of the principle mix the contents of and probably boundary 132 EV FRIT HOLDER gar-z...— GLASS FRIT l I I I . . M CAPILLARY l l I I I I W i‘ i I l '_ FOlL DISC SCREW CAP Figure A-l.--Schematic Diagram of the Modified Capillary Cell. F.) LA) LA) into the screw cap, and tightened with a pair of pliers to insure a good seal to the foil The capillary was immersed to within one half inch in the bulk solution, and the bulk con- equilibrate at 25°C for one hour. The capillaries were then filled with degassed (h tracer afimnmawn mam Amv ohm AHV whommfiz oomOQEHImonm m>mso xmpsH o>aooosuom sous .Hfioo coamsuuao F Amv Adv b Hfll Pflsvam a .............. _ x uuuuuuuuu en ---- thun..u-..---ndv madame — 1|d IL pwpmoehmhp 15H equd studeasoqoqd 141 glass solution reservoirs made from 50 cc syringes [‘__j filling E] A [[‘% B[£ _ syringe valve 2 valve 1 \ cell cell body window “* $- valve A /;;/ valve 3 boundary ’//4 sharpening slits siphon valve 5 Figure B-2.——Diagram of Diffusion Cell. i'i2 reservoir was covered with aluminum foil to retard evaporation. Valve 5 was Opened, and solution allowed to fill the cell to just below the level of the slit. The filling syringe was then used to draw fluid back and forth through valve 5 to remove air bubbles trapped near the valve stem. Solution was then allowed to fill the cell to about one hal inch above the slit. Valve 5 was then closed. Valve A was then opened. Solution was forced through valve A by the filling syringe, until the solution level in the cell was just above the slit. Care was taken not to force any air from the cell into the siphon line. Valve A was then closed, and valve 5 opened. Solution was allowed to flow into the cell until the level was again on half inch above the slit. Step (A) was repeated until fluid flowed from the siphon line, to insure that the line was 0 O '3 FC) |._J (D C I‘ (D ly filled up to the tee in the line. The process of steps (A) and (5) was then repeated for valve 3, to fill the other side of the siphon line. 10. 1A3 Valve 1 was opened, and the plunger removed from the filling syringe. At this point, the siphon was checked by slightly opening valves 3 and A consecutively to make sure fluid would flow freely from the cell. Valves 3 and A were left closed after checking the siphon. Valve 2 was closed, and the filling syringe filled with the less dense solution. The plunger was then replaced. Valve 2 was then opened very slightly, and solution was allowed to flow very slowly down the wall of the cell. The flow rate was kept very slow until the level was an inch or so above the slit, to avoid turbulence and mixing at the boundary. After this time, valve 2 was opened a little to allow solution to flow in more freely. To stabilize the boundary, valves 3 and A were opened so that solution flowed through the siphon at a rate of one drop every two or three seconds. When solution began to appear in reservoir A, valves 3 and A were again closed. Solution was forced back and forth through valve 1 to remove any air bubbles from the valve stem. With liquid above the bottom of reservoir A, valve 2 was then closed. The less dense solution was then added to FJ Fl reservoir A until the level was even with that in reservoir B. Reservoir A was then covered with aluminum foil to retard evaporation. The diffusion cell was now ready to be placed in the water bath for the experiment. Before the cell was placed in the water bath, the fringe pattern was checked to make sure the fringes were straight, vertical and in focus. It was usually found that they had drifted slightly away from the vertical since the last experiment. This could almost always be corrected by making a fine adjustment of mirror 3. The cell was then placed in the water bath. Valves 1 and 5 were opened several turns each. Valve three was then opened until the flow rate from the siphon was approximately one drop every six seconds. Valve A was then opened until the flow rate was one drop every three seconds. It was important that the flow rate be the same _from each side of the cell maintain flat boundary. It ()1 was also important that the flow rate into the top and 1 bottom of the cell be the same so that the initial (‘f distribution of concentra ion gradients would be symmetric about the boundary. It usually took about 20 to 30 minutes for the cell to reach the equilibrium temperature, and for a good 77' boundary to form. wne i . the boundary had formed, valves p 15‘ ' rted. A series w I“ U m :5 Q U E (D ..S (D 0 k- 0 (I) (b p 91 D c) . d :3 (D d F) :§ (D *3 U) c1 :1) of seven pictures, at intervals of 2 minutes were taken. (In the faster diffusing systems, the intervals were somewhat shorter.) In some cases, valves 3, 4 and 5 were again opened, anOther boundary formed, and a second set of pictures taken. The photographic plate, a Kodak Type M plate, was developed by the following procedure: 1. The plate was developed for 5 minutes in 2. The devel op me.t was then stopped by a one- (1') minute oak, with continual a in tap 0‘! water. 3. The image was then fixed by a 5-minute soak, with ir termittent agitation, in Ko da k Rapidfix 4. The plate was then removed from the fixer, and washed for about one minute under running F4) water, and then allowed to dry or at least The photographic plates were extremely sensitive to light, 1 solute darkness (no safe 0‘ and had to be handled in a light) throu gho out the entire procedure, until the fixing step had been co.pl eted. A new plate was then inserted into the film holder, making sure that the er*u lsion side of the plate faced outward. This was irport ant because the thickness of the ..J :i 0\ plate was enough to throw the image out of focus, and perhaps change the magnification factor of the camera. The developer and fixer were replaced after every ten runs, in order to maintain a consistently high imag quality in develonment. 1147 COFRUTLR PROGRAMS FOR ATA ANALYSIS U The following computer program uses measured values of the slopes and intercepts to calculate experi- mental values of the ternary diffusivities. The program language is FORTRAN IV, with specific deck structure .1: H 0:) C) O O O ’ I U {:1 (‘1‘ (D *5 O * b c 1‘ 1’3 (D O 0 FJ forl the IB lege of Engineering, . .. n . . 2 L hiChigan State niversity. One constant, R /R1, mUSb be specified within the program at the designated point. -‘ Other data (the lepes and intercepts) is read by the computer. // J$B // RflR TDIRF *IECS (CARD, 14A *EXT END RD RRECIS *NCNRRZC ESS RRflG *ZNE NERD l\iLoL AM -REAL 12M, IA, L2x, LA CEMMZN x READ (2,10 FERMAT (A E LA = IA + SA L2M = 12M + 3 IF (12M) 3, u, WRITE (3, 20) o FERMAT lHl, 6 2% IA, 82M, SA HFJ o 3 :2N, IA, s2M, SA, L2M, LA 3 15.5) 2 R /R1 MUST BE SPECIFIED LL U3 Q? 3 2 C C TI R AL C AT THIS REINTR C e.g. R = .A2999 1A8 CALL PRTS (I2M, II = -(X**2 + L 322 (X**2 + I 31 = (I2M — 32 D2l = (L21 — o; N.:T3 (3,AC) 31; LRIT: (3, AI) 31 WHITE (3, 42) D2 WRIIE (3, 43) 32 AO F RMAT (83 3AA AI FZRMAT (83 3A3 = A2 T RMAT (83 23A = A3 FCRMAT (8H DEB = GE TC 1 A A CENT '33 CALL EXIT END // FTR PRTS *EXTENDED PRECZSZEN "CNL WORD INTEGERS *NCNPRCCTSS TRLCRAM *LIST SEURCE PRLTRAM UBRZUTZNE PRTS IN ms TN A, EM N X pgawqp> 1\ _ - ~\. J./ -' 23 J-/ 2 2 \ \ ) I [11.0 /'\ &/‘\/‘\ 8 H L) r] m w L) *U O U C) LU U [u L 5‘ XEQ TD The a non a ar volum cinary m1 8.. ine ll [1) ( I \ 72 \J- \ J ) (1 H LU 1/. ()A ll vities H following program predicts ternary diffu U} "sociated system from friction coefficients, , viscosities and the Nargules constants of U) :1 C U) c I U‘ (D *3 (D Q) Q. [1) U) } J :3 p xfi “V“; fl r—T‘ ~ ‘V systems. inese constart program language is FORTRAN IV, with specific deck structure for th IBM 1800 computer of the College of" 203 *2 4 _- .- L.“ V. fi..,\ , h .- ' angineering, Micaigau state univerSity. // J23 // F23 :Izcs :NQN;D *EXTE:T3 *UNE FRED JED 1J1“ ..UJ. ¢\‘ WEED ImeGERS REAL oni, LNG2, LNA, LNB READ (2, 200) 412, A21, 413, A31, A23 A32, 0 FzRALT (7? 10.5) READ (2, 201) ETA f \) C) {\J OOCDO NH Mk4 LLLM4T (F10 5) IL (ETA) 1, 2, 1 LL40 (2, 202)V1, V2, V3 LL43 (2, 202) 3:01, :02, s:c3 LL40 (2, 202) X4, X3, 10 32344? (3L10.5) 31c1 RLPLLLLNTS LT/(LFICT :ZN CfiEFF) VZLJ = 1m 4 01 + XL * V2 + X 4 V3 C4 = X4/VZLL 03 = 1:3/VZLL 00 = XC/VZLD LXG1 = 2*X4*XL*(421-X4‘421-XL*412) + XL*XL*412 + 2*X4*X0*(431—X4*431—Xc*413) + XC*XC*A13 + (X3*Xc—24X4*XL*X0) * (421 + 413 + 432 - 0) X4? = X4 LP = XL X0? = c LNG2 = 2 *XL*X X4* (412— 1:L*412— X4*421) + X4*X4*421 + 2.*XB*XKC* (432-1:L*432-Xc*423) + XC*X0*423 + (X4*Xc-2. *X4*XL*XC)*(421 + 413 + 432 - C) LNA = LNGl + 4L00 (X4) LNB = LNG2 + 4L00 (XB) VCXLX ; VoLL - (X4 + .003) * V1 - XL * V2 cmst = ONE N/V3 TMZLs = X4 + XB + .003 + CMOLs X4 = (X4 + .OO3)/TM6LS X8 = XB/TMZLS X0 = c1quS/T40L s LNGl = 2. *X4*XL*(421-X4*421—XB*412) + XL*X3*412 + 2.*XA*XC*(ABl—XA*A31—XC*Al3) + XC*XC*413 + (XL*Xc-2.*X4*XL*XC) * (421 + 413 + 432 - C) LNG2 = 2.*XL*X4*(412—XB*412—X4*421) + X4*X4*421 + 2.*XL*XC*(432-XB*432-Xc*423) + XC*XC*A23 + (X4*Xc - 2.*X4*XL*XC) * (421 + 413 + 432 - 0) DADCA = (LNGl + ALOGCXA) - LNA) * VOLD/.003 DBDCA = LNG2 + ALOG< '11 :1:- XB+ +||+Il + 1) L13 :17 f‘J 2.: j.— 9 O qjqunjm 00000 152 1 = 2. 4X44XL4(421— X44421-XL4412) + XL4XL4412 + 2.4144X04(431-X44431— X04413) + X04X04413 + (XL4X0-2.4X44 L4X0) 4 (421+413+432-0) 2 = 2.4tL4t44(412—XL4412- X4+421) + 44X44421 + 2.4XL4XC4’432-XL40 32— XC*'A 23) + X04X04423 + (X44X0—2. X44XL4X0) 4 (421*413+432—0) CL = (LNGl + 4L00 - LNA) 4 V0L0/.003 CL = (LNGQ + 4L00(XL) — LNB) 4 VOLD/.003 (3,100) (3,101) 04004, 0L304, L400L, DLDCL (1' L4204 33004 343031 DLDCL T) 4P 3? CP 04004 ET4*CA*(SIGl*(l.—Vl*SIGl) + SIG3*V3*CA) LLLc4/LT442 40L4(V340103-V243102) L4LCL LTA40L4(Q*14(1.-V143101) + S1034V340. LLLLL/LL'4 H40 4(V343103-V243102) DJDCA/ZTA*"“*CS* V3” G3—Vl*SIGl) LLLC4/Lt442:4(s*22 (1,-V240L) + SIG3*V3*CB‘ LLL B/F”‘*“A*C *(V3*SIG3—Vl*SIGl) DBD L/L1.40L4(s 024(1.—V240L) + SIG3*V3*CB) (3,105) (3,106) 044 (3,107) 34L (3,108) :34 (3,109) DEB (13 , // 'TLL LLL010TLD DIFFVSIVITIES ARE (1L , 1L44 = 1 ,.L 15 5) (1L , 104L - 1 , L 15.5) (1H , 'DBA ' , F 15.5) (1L , 'DBB = 1 , L 15.5) 1 ‘XIT .2“ n U APPEI‘QDIX 153 APPENDIX D oncentrations in the system * (B) - Benzene(C) C') :D m C) C) EM (muu.:nm maswflm .o.m .10.: U120 x SOT 159 .on EQOMOLOHQQ o.m : Amv mcmNcmm u A We. H au.mna magmfim +ooa .oom IL“ u-i Q 160 l BETfiii. .Amv mflmNflmm I A ENQII.wIQ mhsmfim oa H+ o.o HI mI mI :I w n L 0 w 4 o Io.H Io.m O .-o.m O IrOo: . o.m mag SOT X 161 ii“ 3> o 200‘~ g: 1001f O + ., + . ‘ I . + I 1, _u -3 -2 -1 0 0 +1 0‘0 Figure D-7.-- —£— vs do in the System Chloroform (C) - Acetone “DA (A) — Benzene (B). 162 .A Em mun.muo wasmfim m6 m+ m+ 0.0 as m: m: a: 1-H + fi— lb #b O 163 m We. .23 onumo< I ADV ELOMOQOHQU I Amv mcmwcmm Emumzm map 5” a w> IHI II.mIQ mhswfim ma m+ m+ H+ o.o HI NI mI :I .r i i c i u w u o ..00H 0 o O ..oom IL“ r—i Q l6u .mm new Mm meMpmcoo mecH m>wpommmmm mo coaumcHEhmmeII.oaIQ mhswflm IhI o< :oo.+ moo.+ moo.+ Hoo.+ 0.0 Hoo.I moo.I moo.I :Hoo. :moo. Imoo. IrJOO . L.moo. :2 x0 .... C. a» e. .C I .Q 3 A mi system 8 5 I Q. m 0 r ..I \./ nu ( e n e Z n e B _ \/ ppm“ I..\ e n O t e C «H—u _ ) .0 Av 7. /-\ _ m D T O Q «I n O a r 0 r0 .1 . .3 D nv S e e h P t u do n .l T; 3K *I. orm(C) r‘- ~ gs 051101701 E. J 5 ,4 .1 71 x0 «1. Ann Qu APPENDIX E 166 Téikalle E—1.--Experimenta1 B' £33ystem.Chloroform(A) — C b {oiokIlIbM wtwvmwm flattOtt ‘TAEBILIS E-2.--Experimental Binary Mutual Diffusivities in the X APPENDIX E EXPERIMENTAL RESULTS nary 1 Mutual Diffusivities in the arbon Tetrachloride(B) at 25°C. D AB x 105, cm2/sec' 1.557 1.572 1.680 1.757 1.779 1.976 2.007 System Benzene(A) - Chloroform(B) at 25°C. D AB x 105, cm2/sec IX \ O.C)C)ES 0.C)C)ES 0.2?(3'7 0.207 0-31453 0-3‘H-53 0.4-$314 0.7'8355 0.785 0.9‘5953 _\ 2.345 2.359 .1414 .uua .u95 .u19 .396 .3ua .344 .265 NNNNNNNN 167 ...J O\ 00 TABLE E-3.-—Experimenta1 Binary Mutual Diffusivities in the System Ether(A) - Carbon Tetrachloride(B) at 25°C. XA DAB x 105, cm2/sec 0.005 1.50 i .05 0.200 1.95 i .06 0.1400 2.38 i .114 0.600 2.99 i .20 0-800 3.76 2 .3‘4 0. 995 u 2: .59 .07 ”“3 TABLE E—A.-—Experimenta1 Binary Mutual Diffusivities in the System p-benzoquinone(A) - Benzene(B) at 25°C. 5 2 i steink DAB x 10 , cm /sec 0.0041 2.20 0.0057 2.16 0.0132 2.09 0.0239 1.98 0.0266 1.95 0.0315 1.96 0.01430 2L9u TAE31613 E-S.-—Experimental Tracer Diffusivities in the System 2-Butanone(A) — Carbon Tetrachloride(B) at 25°C. J<1X D: x 105, cm2/sec D; x 105, cm2/sec \ “-000 1.320 8-837 1.611 0.518 2.260 0' '78 2.720 0.325 3.183 ' 82 3.300 0995 2.973 / FJ 0\ KC) P3 ‘AES ‘E-6.--Experimenta1 Tracer Diffusivities in the T17 fSysytenIp-benzoquinone(A) - Benzene(B) at 25°C. XA DA x 10 , cm /sec 08x 10 , cm /sec 0.0000 2037 <0.0010 2.%2 0.0103 201 0.0464 201 'TAEHJS E—7.--Experimenta1 Tracer Diffusivities in the Ehystem.Ether (A) - Carbon Tetrachloride(B) at 25°C. * * xA DA x 105, cm2/sec Df3x 105, cm2/sec 0.000 1.32 0.024 1.629 t .082 0.024 1.674 t .084 0.332 2.636 t .132 0.530 3.796 t .190 0.850 5.383 t .269 0.999 4.395 i .220 1.000 7.91 r .396 TABLE E-8.--Experimenta1 Density and Viscosity in the System p-benzoquinone(A) - Benzene(B) at 25°C. XA density, g/ml viscosity, cp 0.0011 0.8738 0.600 0.0103 0.8767 0.608 0.0283 0.8826 0.617 0.0387 0.8861 0.629 0.0464 0.8896 0.631 170 TABLEIL9.-—Experimenta1 Density and Viscosity in the EWstem Ether(A) - Carbon Tetrachloride(B) at 25°C. XA density, g/ml viscosity, cp 0.000 1.585 0.913 0.1851 1.4177 0.677 0.3119 1.3047 0.445 0.4808 1.1536 0.583 I 1.000 0.7074 0.225 TABLE E-10.-—Experimenta1 Density and Viscosity in the _ II System Acetone(A) — Benzene(B) - Chloroform(C) at 25°C. xA xB XC density, g/ml viscosity, cp 0.328 0.339 0.333 1.0457 0.4869 0.339 0.332 0.329 1.0427 0.4870 APPENDIX E 171 APPENDIX F THERMODYNAMIC DATA TABLE F-1.--Activity Data for Acetone (1) — Benzene (2) at 25°C (from Timmermans [30]). X lny2, lnyz, lnyl, lnyl, ' 2 exp. eq. (24) exp. eq. (25) 0.1251 .5002 .4827 .0011 .0060 0.2500 .3937 .3896 .0265 .0278 0.3652 .3122 .3022 .0662 .0669 0.5550 .1829 .1672 .1802 .1835 0.7150 .0898 .0750 .3400 .3452 0.8249 .0659 .0300 .4846 .4966 0.8862 .0387 .0131 .5793 .5971 0.9500 .0301 .0026 .7169 .7148 Margules constants: A12 = .8169 A21 = .5685 TABLE F-2.--Activity Data for Acetone (1) — Chloroform (3) at 25°C (from Hildebrand and Scott [21]). X3 1ny3, lny3, 1nyl, lnyl, exp. eq. (24) exp. eq. (25) .0600 -.6733 -.6771 -.0101 -.0019 .1840 -.5276 -.5502 -.0202 -.0198 .2630 -.4308 -.4700 -.0513 -.0431 .3610 -.3711 ~.3723 -.0943 -.0873 .4240 -.3285 —.3126 -.1278 -.1259 .5080 -.2614 -.2379 —.l985 -.1913 .5810 —.l985 -.1788 -.2877 -.2621 .6620 -.1287 -.1208 -.3857 -.3574 .8020 -.0513 -.0441 -.5798 -.5682 .9180 —.0101 -.0079 -.7765 -.7917 Margules constants: A13 = -.9791 A31 - -.7372 172 173 TABLE F-3.-—Activity Data for Benzene (2) - Chloroform (3) at 25°C (from Timmermans [30]). --15 x2 1nY2 1nY2 1nY3 1ny3 exp. eq. (24) exp. eq. (25) 0.1340 -.3439 -.2248 +.0119 —.0066 0.2600 —.2837 -.1548 +.0109 -.0237 0.3180 -.1767 -.1278 -.Ol31 -.0347 0.6400 -.0598 —.0299 -.0845 -.l225 0.7160 -.0284 -.0178 -.1301 -.1480 0.8660 -.0202 -.0036 -.1532 -.2012 Margu1es Constants: A23 - -.3180 A32 = -.2500 TABLE F-4.-—Constants for use in equations (20) through (24) for the system Acetone (1) - Benzene (2) - Chloroform (3) at 25°C. Al2 = .8169 A21 = .5685 A13 = -.9791 A31 = -.7372 A23 = -.3180 A32 = -.2500 C = 0.0 APPENDIX G Capitals A C APPENDIX G NOMENCLATURE interaction perameter a] concentration, interaction parameter diffusivity Onsager diffusivity reduced second moment reduced height-area ratio energy of vaporization force Gibbs free energy enthalpy of mixing intercept of second—moment curve at a = 0. intercept of height-area ratio curve at a = 0. flux, fringe number equilibrium constant phenomenological coefficient; length of capillary intercept of second-moment curve at a = 1.0 intercept of height-area ratio curve at a = 1.0 magnification factor of camera number of species in solution simplifying constants 175 +