TRACER AND MUTUAL DIFFUSION IN ISOTHERMAL LIQUID SYSTEMS WITH SPECIFIC ASSOCIATION Thesis for the Degree of Ph. D. MICHIGAN STATE UNIVERSITY GUY B. WIRTH 1968 LIBRARY Michigan 3 if: (F: University “—fi‘ .—. all... _..? This is to certify that the thesis entitled TRACER AND MUTUAL DIFFUSION IN ISOTHERMAL LIQUID SYSTEMS WITH SPECIFIC ASSOCIATION presented bg Guy B. Wirth has been accepted towards fulfillment of the requirements for KIM—D;— degree in an inee ring WM Major professor Date June 14, 1968 0-169 n‘ .4.- " ‘ ‘ fine.“_ I Y IIN‘BING BY I HUM: & SUNS' I 800K nmnrnv mp nnnnnnnn olnu‘nu "IIIGM". IICUIGII t . a .n’ ABSTRACT TRACER AND MUTUAL DIFFUSION IN ISOTHERMAL LIQUID SYSTEMS WITH SPECIFIC ASSOCIATION by Guy B. Wirth A theoretical and eXperimental study was made to investigate the effects of specific association on mutual and tracer diffusion. The Hartley and Crank (1) intrinsic diffusion approach was used to describe the diffusion mechanism from which equations were derived which relate the mutual and tracer diffusivities to the thermodynamic and molecular prOperties of the systems of interest. For associated systems, it is necessary to con- sider diffusion of clusters as well as the individual molecules. A general chemical model, suggested by Nikolskii (2), was used to calculate the concentration of these clusters and the solution activities. Three types of associated systems were studied: (a) binary mixtures of A and B where A associates to form a dimer cluster and B remains inert (b) binary mixtures of A and B where A associates with B to form a bimolecular cluster (C) ternary mixtures of A, B and C where A aSSO" ciates with B as in (b) and C remains inert. Guy B. Wirth A Mach-Zehnder diffusiometer was used to measure the mutual diffusion coefficients. Tracer diffusion co- efficients were determined by a modified capillary method, develOped to minimize errors inherent in previous capillary studies using Carbon—l4 tracers. The accuracy of the capillary method was found to be 12%. Data were collected for six systems at 25°C: (I) benzene-carbon tetrachloride--an example of a nonassociating system (2) chloroform - carbon tetrachloride-~an example of a nonassociating system (3) methyl ethyl ketone-carbon tetrachloride-—a type (a) associated system (4) acetic acid - carbon tetrachloride-~a type (a) associated system (5) ether—chloroform--a type (b) associated system (6) ether-chloroform - carbon tetrachloride--a type (c) associated system. These data agree well with values predicted by the equations derived in this work. Deviations can be attributed to the inability of the models to take into account nonideal behavior as well as inaccuracy in the eXperimental values of diffusion coefficients and act1v1ties. 1G. S. Hartley and J. Crank, Trans. Faraday Soc., 32. 801 (1949). 2S. S. Nikol'skii, Theoreticheskaya i EkSperimental'naya Khimiya, g (3), 343 (1966). TRACER AND MUTUAL DIFFUSION IN ISOTHERMAL LIQUID SYSTEMS WITH SPECIFIC ASSOCIATION By 1‘1 \ -2 Guy Bi Wirth A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemical Engineering 1968 ACKNOWLEDGMENT The author wishes to eXpress his appreciation to Dr. Donald K. Anderson for his guidance and encouragement during the course of this work. The author is indebted to the Division of Engineering Research of the College of Engineering at Michigan State University and to the Dow Chemical Company for providing financial support. Appreciation is extended to the author's sister, Carol, for her help in the preparation of this manuscript. The understanding and patience of the author's family and expecially his wife, Darlene, is sincerely appreciated. ii TABLE OF CONTENTS ACKNOWLEDGMENT . . . . . . . . . . . . . . . LIST OF FIGURES . . . . . . . . . . . . . . LIST OF TABLES . . . . . . . . . . . . . . . INTRODUCTION . . . . . . . . . . . . . . . . BACKGROUND . . . . . . . . . . . . . . . . . Intrinsic Diffusion . . . . . . . . . . . Planes of Reference . . . . . . . . . . . Mutual Diffusion in Binary Systems . . . . Tracer Diffusion . . . . . . . . . . . . Diffusion in Multicomponent Systems . . . Thermodynamics of Solutions . . . . . . . THEORY I O O O O O O O O O O I O O O O O O 0 Concentration Dependence of Diffusivities --Associated Systems . . . . . . . . . . Selection of Systems . . . . . . . . . . . EXPERIMENTAL METHOD 0 O O O O O O O O O O 0 Modified Capillary Cell . . . . . . . . . Procedure . . . . . . . . . . . . . . . . Calculation of Diffusion Coefficients . . Calibration of the Cells . . . . . . . . . Analysis of Radioactive Samples . . . . . Materials . . . . . . . . . . . . . . . . Discussion of Error . . . . . . . . . . . RESULTS AND DISCUSSION . . . . . . . . . . . Theoretical . . . . . . . . . . . . . . . Experimental 0 O . O C O O C O O O O O O 0 CONCLUSIONS . g o o o o o o o o o o 0 0 0 0 iii ii vii 10 21 24 3O 32 46 46 63 74 75 78 8O 82 85 86 89 91 91 96 114 FUTURE WORK . Theoretical Experimental APPENDIX I APPENDIX II APPENDIX III APPENDIX IV APPENDIX V APPENDIX VI APPENDIX VII APPENDIX VIII NOMENCLATURE BIBLIOGRAPHY Fortran Compute determination o Constants . Experimental Procedure Solution of Equation (226) Fortran Computer Program for Diffusion Coefficient Calculations Determination of Friction Coefficients From Tracer Diffusivities Experimental Tr at 25°C . . . Experimental Mutual Diffusion Data at 25°C . . Experimental De Data at 25°C iv 1: acer Diffusion Data nsity and Viscosity Program for the f Equilibrium Page 116 116 117 120 121 123 128 129 132 135 136 137 141 LIST OF FIGURES Figure Page 1. Activity as a Function of Mole Fraction for the Methyl Ethyl Ketone - Carbon Tetra— Chloride system C O O O I O O O O O O O O 0 66 2. Total Pressure as a Function of Mole Fraction Ether for the Ether - Chloroform System . . 69 3. Logarithm K as a Function of l/T for the Ether - Chloroform System . . . . . . . . . 71 4. Heat of Mixing as a Function of Mole Fraction for the Ether-Chloroform System . 72 5. Schematic Diagram of the Modified Capillary cell 0 O O O O I O I O O O O O O O O O O O 77 6. Diffusion Cell Coordinates . . . . . . . . . 81 2 * * ' * L and 7. C iavg/Ci0 as a Function of Dle/ RD*/L O O O O O O O I O O O O 0 O O O O O O 83 1 8. Count Rate as a Function of Mole Fraction for the Methyl Ethyl Ketone - Carbon Tetrachloride System . . . . . . . . . . . 87 9. Count Rate as a Function of Mole Fraction for the Chloroform - Carbon Tetrachloride 88 SYStem O O O O O O O O 0 O O O I O O 0 O G 10. Density and Viscosity of the Benzene - 97 Carbon Tetrachloride System at 25°C. . . . 11. Density and Viscosity of the Chloroform - 98 Carbon Tetrachloride System at 25 C. . . . 12. Tracer and Mutual Diffusion Coefficients for the Benzene-Carbon Tetrachloride System 99 at 25°C. 0 O O O O O C C O O C O O O c O O Figure Page 13. Tracer and Mutual Diffusion Coefficients for the Chloroform - Carbon Tetra- chloride System at 25°C. . . . . . . . . 100 14. Tracer and Mutual Diffusion Coefficients for the Methyl Ethyl Ketone - Carbon Tetrachloride System at 25°C. . . . . . . 103 15. Tracer and Mutual Diffusion Coefficients for the Acetic Acid - Carbon Tetra- chloride System at 25°C. . . . . . . . . 106 16. Tracer and Mutual Diffusion Coefficients for the Ether - Chloroform System at 25°C. 0 C O O O O O O O O O O O O O O O O 108 17. Density and Viscosity of the Ether — Chloroform — Carbon Tetrachloride System at 25°C. I O I O I O O O O 0 O O O O O O 111 18. Tracer Diffusion Coefficients for the Ether - Chloroform - Carbon Tetrachloride System at 25°C. . . . . . . . . . . . . . 112 vi LIST OF TABLES Table page I Capillary Cell Calibration Results . . . . 84 II Summary of Diffusion Equations . . . . . . 92 III Tracer Diffusion Data for the Benzene-Carbon Tetrachloride System . . . . . . . . . . . 132 IV Tracer Diffusion Data for the Chloroform - Carbon Tetrachloride System . . . . . . . 132 V Tracer Diffusion Data for the Methyl Ethyl Ketone - Carbon Tetrachloride System . . . 133 VI Tracer Diffusion Data for the Acetic Acid - Carbon Tetrachloride System . . . . . . . 133 VII Tracer Diffusion Data for the Ether - 133 Chloroform System . . . . . . . . . . . . VIII Tracer Diffusion Data for the Ether - Chloroform - Carbon Tetrachloride 134 System . . . . . . . . . . . . . . . . . . IX Mutual Diffusion Data for the Benzene - 135 Carbon Tetrachloride System . . . . . . . X Mutual Diffusion Data for the Chloroform - 135 Carbon Tetrachloride System . . . . . . . XI Density and Viscosity Data for the Benzene - 136 Carbon Tetrachloride System . . . . . . . XII Density and Viscosity Data for the Chloro- 136 form - Carbon Tetrachloride System . . . . XIII Density and Viscosity Data for the Ether - Chloroform - Carbon Tetrachloride 136 SYStem o o o o o o o o o o c 0 vii INTRODUCTION Interest in liquid diffusion had been maintained for over a century. Since the diffusion process is one of the fundamental aspects of liquid behavior which must be described by a liquid state theory, diffusion rates have been useful in testing such theories. The importance of understanding diffusion has also increased as more basic consideration is being given to various mass trans- fer processes where liquid diffusion is one of the important factors. In the usual sense, diffusion of mass refers to the dissipation of a chemical concentration gradient by molecular transfer with no overall mass flow caused by forces external to the diffusion system. Thus, if two miscible liquids, containing the same components but having different concentrations, are placed together, the molecules of the components will diffuse from the higher to lower concentration until the solution is uniform. ThlS diffusion process results from the thermal energy of the molecules which gives rise to their Brownian movement. Fick (28) was one of the first to mathematically describe diffusion. By analogy to conductive heat trans- fer, he deduced that the rate of transfer of a material .,_Jn__ _ _ _ ‘-P'“*"“fiwaur J. was proportional to its concentration gradient at constant temperature and pressure. That is: '3 = -Dvc (1) where 3 is the flux (i.e., the rate of transfer) of material across a reference plane of unit area and VC is the three dimensional concentration gradient. The constant of proportionality, D, is called the diffusion coefficient and was first thought to be constant for each system at constant temperature and pressure. But, it is now known to be a function of concentration. Equation (1) is usually simplified to the one-dimensional form given by equation (2) because of the difficulty in obtaining measurements for diffusion in more than one direction. — .. 39. Either equation (1) or (2) is used to define the diffusion coefficient. Tracer diffusion is different than ordinary diffusion in that it is observed when there are no chem- ical concentration gradients present. To better visualize this, consider a solution of uniform chemical composition in which part of the molecules of one component is labelled. If there is a higher concentration of labelled molecules in one region of this solution than in another, measurable diffusion of these labelled molecules will take place even though the solution is chemically uniform. The flux of “1““ “‘" “-‘=¥£2‘.F.J labelled molecules is measured in the laboratory. Frequently this labelling is accomplished with a radio— active tracer, from which the name tracer diffusion is derived. As in ordinary diffusion, the tracer diffusion coefficient is defined by equation (2). For example, denoting the labelled component by *, the defining rela- tion is: J* = -D* —- (3) where D* is the tracer diffusion coefficient. Like the ordinary diffusion coefficient, it also is a function of concentration. Studies have been published recently concerning the concentration dependence of the diffusion coefficients in nonideal solutions in which nonideality is caused by association of the molecules. In an associated mixture, two or more molecules may cluster together and diffuse through solution as a unit instead of as single molecules. In this study, relationships between tracer and ordinary diffusion coefficients are derived for systems where the molecules associate to form simple clusters of two molecules. Experimental measurements of the diffu— sion coefficients were used to verify these relationships. 4-. ‘uua up. ‘-‘.- BACKGROUND Intrinsic Diffusion For a system to be at equilibrium at constant temperature and pressure, the chemical potential must be uniform throughout the system. For example, a system having two phases will be in equilibrium only if the chemical potential of each species is the same in both phases. If the chemical potentials of a component are not equal in both phases, there will be a spontaneous migration or, prOperly, diffusion of that component from the phase of high potential to the phase of lower poetn- tial. This diffusion will continue until the chemical potentials, but not necessarily the concentrations, are equal in both phases. The diffusion process will also occur within a single phase. The molecules will diffuse from a region of high concentration and high potential to a region of low concentration and low potential until the concentra- tion and the chemical potential are uniform. The rate 0f approach to equilibrium will depend upon the differ- ence in chemical potential gradient. Therefore, it is reasonable to state that the chemical potential gradient 4 is the quantity which drives the system to equilibrium, that is, it is the driving force for diffusion. [driving force for diffusion] = - 3%; (4) A nonuniform chemical potential can be thought to my ‘ :- exert a force on a molecule which causes the molecule of the i—th component to move with a velocity, Vim’ with respect to the surrounding medium. A molecule moving in I this fashion is said to be undergoing intrinsic diffusion. As this molecule diffuses, momentum transfer to the surrounding medium results in a resisting drag force which is exerted on the molecule in a direction, Opposite to that of the velocity Vim' If the molecule is Spherical and if the surrounding medium is assumed to be a continuum, Stokes' law applies and [resisting force for dissusion] = ~6nrirnfim (S) where ri is the molecular radius and n is the viscosity of the surrounding medium. When the molecule diffuses with a constant velocity (i.e., it experiences no accel— eration), the sum of the resisting force and the driving force for diffusion must be zero.1 From equations (4) 1For a more exact description of nonsteady-state diffusion, a term associated with the acceleration of the molecule should be included. Although such terms may be important in some cases, they become negligible for the small chemical potential gradients encountered in experi- ments designed to measure diffusion coefficients (42). and (5) this sum is Sui - —§; - N67rrinvim = 0 (6) where N is Avrogadro's number. Solving this equation for vim and then multiplying by Ci’ the concentration of the i-th component in moles/volume, yields: Ci aui ivim = m 73' ‘7’ Note the Civim is the molar flux of 1. And, Since vim IS the velocity relative to the medium, Civim will be the molar flux relative to the medium, Ji . Hence, m J = “Ci LL11 (8) im 6nrinN 32 From thermodynamics, _ = , 9 pi “io NlenAl ( ) where Ai is the activity of i and “i0 is the chemical pOtential of i in the standard state-—a function of temp— erature and pressure only. Using equation (9), one can rewrite equation (8) as __ kT alnAi 33$ . (10) Jim I 6nrlfi alnCi 82 Comparing this equation to equation (2), one can see that the bracketed term corresponds to a diffusion coefficient. Since the flux, Jim’ is with respect to the medium, the diffusion coefficient is defined with respect to the medium. alnA. kT 1 (ll) Dim = 6nr.n alnC. 1 1 In dilute solution, the activity of the solute is proportional to its mole fraction. Further, its mole fraction is Ci Xi = Esolvent (12) so that alnAi = 1. (13) alnCi Thus, equation (11) can be written as kT (14) Dim = 6wr n ° ' i This equation, known as the Stokes - Einstein equation, was derived independently by Einstein (24) and Sutherland (64). It has found application in the diffusion of colloidal particles and macromolecules in low molecular weight solvents. However, if the radius of the diffusing molecule is smaller than the molecules of the solvent, the molecule can no longer be considered to diffuse through a continuum Rather, the medium must be viewed as containing particles Also, most molecules are not spherical of finite size. Thus, the and their shape will effect their mobility. resisting force for diffusion may not be given by Stokes' law. The idea that the diffusion coefficient is inversely proportional to viscosity is an old one. In 1858 Wiedemann (68) observed that for a given solute in a series of solvents, the product Drivaried less than D alone. In view of this observation, it may be assumed that, even where Stokes' law is not obeyed, the resistance to diffusion can be separated into a product of viscosity and a parameter, f, which depends very little on viscos- the resisting force for diffusion is: ity (33). Now, (15) [resisting force for diffusion] This friction co- where fi is a friction coefficient. efficient will, in general, be concentration dependent and also dependent upon the size and shape of the molecule By calculating fi for several solutes in one solvent, Anderson (1) found that it is approximately proportional to the square of molecular radius. Bidlack (7) confirmed Anderson's findings and also concluded that fi depends upon molecular shape, but only for very unsymmetrical molecules. In any case, equation (15) will be accepted as the resisting force. Again, one can sum the driving and resisting forces for diffusion to obtain: aui ... ...—32 — finv1m = 0. (16) From equation (16) it follows, as before, that: alnA l - (17) kT D . :: 1m fir131nCi This diffusion coefficient, which Hartley and Crank (33) called the intrinsic diffusion coefficient, will be used later in this work. Intrinsic diffusion in liquids and solids may be clarified by identifying it with the widely known kinetic theory of diffusion developed by Eyring and co-workers According to this theory, the (25, 26, 27, 39, 54, 62). diffusing molecule spends most of its time oscillating within a "cage" of other molecules. At relatively infre— quent intervals, such a molecule receives sufficient energy to break through the potential barrier which limits its motion and moves in a random direction into another "cage." This diffusion process thus consists of a series of activated, random jumps relative to the loose quasi— crystalline lattice of the liquid or the better-defined lattice of the solid. The theoretical treatments of the above authors give the flux of the diffusing molecules with respect to the surrounding molecules. Diffusion coefficients calculated from this theory should be identi- fied with intrinsic diffusion coefficients. From the foregoing discussion, it can be said that the diffusion of molecules relative to their surroundings, i.e., relative to the surrounding medium, is one of the basic mechanisms for liquid and solid diffusion. Further, intrinsic diffusion is seen to occur because of the ran- dom motion of the molecules. 10 Planes of Reference While Fick (28) extended the laws of heat conduction to the problem of diffusion in a simple manner, a complication arises in connection with diffusion. In a diffusing mixture, the velocity of the individual 2 are different and as a result, there is a This bulk velocity itself To speCies, Vic, bulk velocity of the fluid. contributes to the overall flux of each Species. separate pure diffusive flux from the overall flux, it follows that the flux of a species should be measured relative to a plane which moves with the bulk velocity. There are several ways to define the bulk velocity so it is necessary to carefully specify the reference plane across which the flux is measured. Since the concept of the reference plane is basic to the definition of the diffusion coefficient, it is appropriate to discuss some In this discussion, the following relations will C p= Z oi (18) i=1 of them. be required: 2The velocity, viC does not refer to the velocity of a particular molecule of the i~th component; rather it refers to the average velocity of the molecules in a small volume of the liquid. This velocity is a function of position and time. and c 1 = Z CV. (19.) i 1 i=1 where p is the density of the mixture (g/cc), p is the is mass concentration of the i-th component (g/cc), Vl the partial molal volume of the i-th component (cc/mole) and c is the total number components in the mixture. Mass Fixed Reference Plane The mass fixed refer— ence plane is defined aS a plane through which there is no net transfer of mass. From this word definition, it follows that (20) C Z MiJiM = 0 i=1 where JiM is the molar flux of i relative to the mass fixed plane and Mi is the molecular weight of i. The meaning of this reference plane may be clarified from the standpoint of experiment by determin— ing its velocity and the flux JiM in terms of fluxes relative to laboratory coordinates. The mass flux of the i-th component through the mass fixed plane is determined by the velocity of the component, ViM’ relative to that plane. That is, MiJiM = iniM° (21’ But, ViM is the difference between the velocity of the i-th component relative to fixed coordinates and the 12 velocity of the mass fixed plane relative to fixed coordinates Vim = Vic ’ VMC (22’ the velocity of the mass fixed plane, is called where VMC, Combining the last two equa— the mass average velocity. tions yields: V ...)- (23) MiJiM = pi(vic - Summing equation (23) over all Species and combining the result with equation (20) gives: C i=1 Or, after solving for GMC’ c E: p.v. _ ._ 1 1C . YMC — 1—1 p (25) Identifying pivic as the mass flux of 1 relative to fixed coordinates, one can see that the mass average velocity is the net flux of mass across a coordinate fixed plane divided by the solution density. C 2: OLCL _ _ 0— 1 1C < Since both p and Jic are functions of position and time, will also be so dependent. VMc 13 To demonstrate the contribution of bulk flow to the overall flux of a component, eXpand equation (23) to (27) or M.J 1 1c = i 1M (28) From this last result, one can see that the mass flux the bulk motion. Number Fixed Reference Plane The number fixed reference plane is defined as a plane through which there is no net flux of molecules; or, equivalently, a plane through which there is no net flux of moles. Therefore, reference plane. From the definition of this plane, it follows that C E: JEN = o (29) i=1 Where JiN is the molar flux of the i-th component rela- tive to the number fixed plane. The relationship of this reference plane to experiment can be seen by noting that the flux _ = . - ‘ (30) JiN ‘ CiviN Ci(VlC VNC) 14 where VNC, the velocity of the number fixed reference plane, is the number average velocity. Summing equation (30) over all components and combining the resulting equation with equation (29) yields: C VNC = Z Crivic i=1 -c—-—— (31) :c. 1 i=1 Since Jic = Civic’ c VNC = :Jic i=1 . (32) c :c. 1 i=1 Thus, the number average velocity, at a position in the fluid, is the net molar flux through a coordinate fixed plane at that position divided by the total molar concen- tration of the solution. The effect of bulk motion can be seen by eXpanding equation (30) to J. =J. +c.§-’ . (33) Again, the overall flux relative to a coordinate fixed Plane is the flux relative to the number fixed plane plus 15 a term resulting from the bulk motion of the solution. (However, this bulk motion of the fluid is defined dif- ferently than in equation [28].) Component Fixed Reference Plane The component fixed reference plane is defined as a plane through which there is no flux of one particular component. Usually, the reference component is taken as the solvent so this reference plane is often called the solvent fixed refer- ence plane. From the definition, J = O (34) 00 where the first subscript, o, designates the reference component and the second the component fixed reference plane. The flux of any other component, i, with reSpect to this reference plane is given by: J. = C.v. = C- V. - v 10 i 10 1 1C oC ‘ ) (35) where VOC is the component average velocity. Writing this equation for the reference component, one obtains: J = C v - v 00 0 0C oC ‘ ). (36) Substituting equation (34) into equation (36) and solving f — ' : or v0C yields = 0 0C . (37) CC C 1 lead to values which A 46 47 are too large. Positive deviations can be correlated with a tendency to form clusters of one type of molecule and negative deviations suggest clusters consisting of different types of molecules. According to the chemical model, these clusters are in equilibrium and the concen- tration of each Species can be calculated. Therefore, it is possible, in principle, to modify the derivation of equation (90) to take the association into account (1). In reality, the equations of the chemical model are not analytically soluble when clusters of five or more molecules exist in solution. The only readily soluble cases are for Simple clusters of two molecules. Three of these relatively Simple types of associa- tion will be discussed below. The tracer diffusivity expressions for these three cases have been independently developed over the last three years by the author. However, Stokes (63) derived similar equations for type (a) and more recently, Carman (10, 11) published diffusion expressions for both types (a) and (b). Type (a) A binary mixture of A and B where compo- nent A associates to form a dimer cluster and component B is inert. Thermodynamic Relations——For this case, the following quantities can be specified by inspection: c = 2, (125) r = 1, (126) n = 3 (127) v11 = -2 (128) Oil = l (129) and v12 = O. (130) Specification of these quantities permits equations (116), (120), (121) and (124) to be solved for the mole fractions of the Species and the activities. The following equa— tions result: x K = A; (131) 25.1 X . 33:4 1 1+Y’ (132) 2 2XA 4KXA X231-m-W, (133) (1+Y) 2 - 4KxA _ xll = —————3, (134) ‘ (1+Y) ZXA (1+Y) x0 = , (135) l (1+Y)2 + 4KX§ 49 2xA (1+Y) (1+Y)2 + 4KXA 4KXA i1? = , (137) (1+Y)2 + 4KX§ Ac = XA (1 + [1+4K]l/2) (138) A 1+)! and 2 AS = 1 - ;:% - —:Ef§§ (139) (1+Y) where (140) K: II 1/2 I1 ’ 4KXA IXA ZII It is now possible to differentiate equation (138) to obtain: (141) 4KX alnAC l + A ____A = ____1:1_ alnxA Y and Differentiation of equation (139) with reSpect to XB comparison of the result with equation (141) demonstrates SlnAg BlnAg that alnX - as required by the Gibbs—Duhem relation. A aian Diffusion Relations--Recall that the total flux of a component is calculated by adding the effects of intrinsic diffusion and bulk motion. For component A, 50 the total flux is the sum of the contributions resulting from diffusion of monomers, dimers and bulk motion. That is, JAv = J1m + 2J11m + CA IVVC ' Vme) (142) where in this case J and Jllm are the intrinsic fluxes lm of monomers and dimers, respectively, and ch ‘ va = ‘ (Jlmvl I J2mV2 + J11mV11 ° (143’ As before, aci Jim = ‘ Dim W (144) and ac A = - ——— - 145 JAV DAB 82 ( ) Now, by substituting equation (143), (144) and (145) into equation (142) and solving for the mutual diffusion coefficient, one obtains: ' ~ BC 3C1 + 2 D Scll _ C V l DAB ‘ Dlm TEX" 11m acA A 1 1m acA c: acll “ 3C2 ° vllbllm ‘36; + v2D2m To; 51 The molar volume of the dimer, V11, is assumed to be: <22 11 = 27A. (147) Substitution of equation (147) into equation (146) followed by rearrangement of terms yields: 8cl ~ 8511 D = _ — AB Dlm ScA + 2D11m acA l CAVA I (148) 8c2 _ D2m SCA CAVB' N ' — _ = = ow, the two relations, CAVA + CBVB l and dCA/dCB -Vé/V , are used to reduce equation (148) to: acl ~ 8511 _ DAB = Dlm Fe; + 2D11m "FEX’ CBVB + (149) 3c2 _ Dzm ac CAVA A Equation (17) is now combined with (149) to give: D _ kT cl Blnxl + 2cll alnxll X + AB 7 77 f c alnx ~ BlnX B 1 A A fllCA A (150) Blnx2 ”I NH EInXB xA 52 By differentiating equation (131) to get: 81nx alnx ll _ 2 l BTXA * mg (151) and by using equation (120) in conjunction with the Gibbs- Duhem relation to give Blnxl BlnA _._.___ = = = ——-— , (152) BlnXA BInXA 31an 81nXB equation (150) may be written as: 4~ I 31 AC D =£I___Cl+a—_Cllx+—lxI-———nA (153) AB n fch {.1ch B f2 AJ ainxA For ease of computation, it is useful to define psuedo mole fractions by equation (123). Thus, the final result is: kT AB I“: (154) U II I I'hl X + as first derived by Anderson (1). Tracer diffusion is concerned with the flux of labelled molecules of one particular component. There— fore, there will be a mixture of dimer clusters containing one and two labels in addition to the labelled monomers. The total concentration of labelled A molecules is given by: 53 f- (155) Since there is no bulk flow in the case of tracer diffu- sion, an eXpression for the flux of labelled A molecules is obtained by adding the fluxes of the various labelled species: * = * ~ * ~ * JA Jlm + Jlm + 2J1:m (156) and as before, 3c* ~ 8611 ~ 851i = * * ** . DA Dlm Fe: + Dllm "To; + 2Dllm ‘66; (157) Assuming that a dimer containing two labelled molecules has the same friction coefficient as one containing one labelled molecule, D11; = D11; and equation (157) becomes: 3c* 85 * 85** 1 ~ 11 11 D* = D * + D ** + 2 * - (158) A 1m Sc; 11m BCA BCA From equation (155), ” * ~** 8c* 3311.23311=1-3Ci (159) * so that: ” sci ” (160) . it DA = 91* DlIm it; + Dum Ill 5 l (I) '14. 54 It is now assumed that the labelled A molecules are distributed such that the ratio of members of labelled monomers to the number of labelled A molecules will equal the ratio of the total number of monomers to the total number of A molecules. Of course, this is not an assump- tion if the labelled molecules are identical to the un- labelled molecules. In terms of concentrations, this equality is: (161) 1.9.8 I (3 o a. Is Since the concentration is uniform throughout the solution, it is apparent that the latter ratio is constant. So, 8c* c CA A Equation (160) can now be written as: x‘l’ ~ — - ~ * — * 163 DA" 01;: D11m XA+Dllm ( I When xi is defined by equation (123). Recalling that ~kT (79) * = Dim 7E1 ’ the final result is: kT l 1 X? 1 DA _ 7T (f; ‘ §_—) XA + :—- ° (164) 11 f11 The inert component, B, diffuses only as monomers so one can write simply: * = * JB J2m (165) from which kT 1 * = __ __ . DB n f2 (166) A relatively simple relation between the tracer and mutual diffusion coefficients can be obtained by combining equations (154), (164), and (166). This rela- tion is: kT X11 XB alnAA ._ 'k . _ — _ 0 DAB - XADE 'I' XBDA 'I' 2 n ~ X W (167) fll A A Stokes (63) and later Carman (10, 11) derived equations very similar to this equation. Type (b) A binary mixture of A and B where A associates with B to form bimolecular clusters. Thermodynamic Relations--The principles involved in TYPE (b) are the same as in Type (a). To avoid DGGd‘ less repetition, many equations will be presented without 56 detailed eXplanation. The following quantities are the same as Type (a): c = 2, (168) r = l, (169) and n = 3. (170) But, the stoichiometric coefficients for Type (b) are: v11 = —1, (171) V11 = 1, (172) and 012 = -l. (173) The equations of the chemical model can now be solved to get: K = 12 , (174) X1X2 x - x + z X = A B , (175) l 1 + z _ + z X _ XB XA , (176) 2 _ l + Z s 1 — z (177) X "' I 12 ‘ 1 z 0; firm: 57 X0 = X _ 1 - z 1 A 2 ’ (178) o 1 - z X = _ 2 XB 2 ' (179) X o = l - z 12 2 ’ (180) AC = xA - xB + z A l + z (181) and AC = xB - xA + z B l + z (182) where K + l (183) Also, the following equation can be derived from equation (134): C alnAA _ l SInXA Z ' (184) Equations (181) and (182) can be shown, by direct differ- entiation, to satisfy the Gibbs-Duhem relation. Diffusion Relations—-The total flux of component A is given by: vVC - vmC (185) JAv = J1m I J12m + CA 58 where Jlm and J12m are the intrinsic fluxes of the monomer and cluster, respectively, and ch ' Vmc = VlJ V2 J12m + V2J2m. (186) The partial molar volume of the cluster has been approxi- mated by: (187) Proceeding by the same steps presented in Type (2), the (188) Again, Anderson (1) was the first to derive this equation. Tracer diffusion in this case is less complicated than in Type (a) because the dimer can carry only one label. Thus the total concentration of labelled A molecules is: ~ * (189) + €12 59 and the flux of these molecules is simply: J: = Jlfi + J13. (190) Again, the final result follows immediately by the process discussed for the Type (a) associated system: 0 kT X . D*=__l_:l_._l.+.:l_ - (191) A 7‘ f1 f XA f 12 12 The same process applied to component B yields: P I X D§=§1[?1_~AX_2.~1. (192, 2 f12 B f12 Combination of equations (188), (191) and (192) gives: C x O BlnA kT 12 A DAB XADB + XBDA 2 r) E BlnXA (193) as also derived by Carman (10, 11). Type (c) A ternary mixture of A, B and C where A associates with B and C is inert. Thermodynamic Relations--Here there are three components, four Species and one reaction. So, C = 3, (194) n = 4’ (195) (196) 60 V11 = '1' (197) 611 = l (198) v12 = -l (199) and 013 = 0. (200) These conditions combined with equations (116), (120), (121) and (124) of the chemical model yield: K = I (201) X = , (202) x - x - -———I + U X = B A K + , (203) - - U l 2xc I K + I (204) 12 KX c (205) 1 _ KXC _ U o _ K + 1 X1 ‘ XA ‘ 2 . (206) 1 _ KxC _ U X0 = x _ K + 1 2 B 2 ' (207) 1 KxC _ U 3" = K+1 ' 12 2 (208) C _ x3 _ xC (209) x x - x - C + U AC _ A B K + 1 A — KxC ' (210) l I K + 1 I U Xc c _ xB ' xA ' R‘I‘I + U AB ‘ Kx (211) c 1 + K + 1 + U a nd C 2XC AC = KxC (212) 1 + K + 1 + U where 2KxC szg 4KXAXB 1/2 U = l ‘ K‘I‘I + IE-I_IF§ ‘ 'R‘T‘I (213) Diffusion Relations--Recall that Simple flow equations, such as equation (2), are not applicable to mutual diffusion in ternary systems. But, since there 62 are no concentration gradients of components in tracer diffusion, a single tracer diffusion coefficient still suffices to describe the process. Then, the intrinsic diffusion model can be applied to describe tracer diffu- sion as before. In fact, the addition of the third inert compo- nent to a Type (b) associating system does not change the diffusion equations from the form of those in the previous case (Type [b]). The flux of labelled A molecules is still: .. ~ * J3 - J1; + J12m (214) and the flux of labelled B and C molecules is: - ~ * Jg — J2; + J12m (215) and , = * 216 J5 J3m ( ) The counterpart of equation (76) for this ternary system is: (217) Ci": 1 -— XAVA+XBB cc . . . * where the denominator 18 again independent of X1' Hence, BlnA? i -———— - ' 79 is valid as before. BlnCI l and equation ( ) 63 Equations (2l4), (215) and (216) can now be simplified to: 1" X0 ._ M 1 1 1 1 DIE—7 f—--~—-—)-,-+:-—- (218) 1 £12 A £12 .0 1 kT l l 2 l 1 05:7]— f—-~———-3g—+-J (219) ~ 2 £12 B fl2 and kT 0* = --— ° (220) C nf3 Although the form of these equations is the same as their counterparts in Type (b) [equations (191) and (192)], the composition variables are different because of the pres— ence of the third component. Selection of Systems Usually, there are several systems which one might gues to associate in a given way, but few whose activity and spectroscopic data have been reported in the literature. Since these data are required to make firm decisions as to the exact nature of the association, the number of systems suitable for study is greatly reduced. Another factor to be considered is the availability of labelled material. Suprisingly, many common chemicals 64 are not available from commercial tracer suppliers. Thus, it is frequently a problem to find a system which can be studied. In this section the six systems that were in- vestigated are presented along with evidence of their association behavior. Chloroform (A) - Carbon Tetrachloride (B) The isothermal vapor-liquid equilibrium data by McGlashan, Prue and Sainsbury (49) coupled with the absence of a temperature effect on the heat of mixing of the system (13) indicates that chloroform — carbon tetrachloride mixtures are nonideal and nonassociative. Krishnaiah, g3 31. (44) concluded further that the system is regular. By fitting heat of mixing data (13) to equation (101) a value of B was determined with which the thermodynamic factor can be calculated from equation (100) as: alnAg (221) ————= - . x. . alnX 1. O 74XA B A Benzene (A) - Carbon Tetrachloride (B) Staveley and co-workers (61) report that this system behaves as a regular solution and as such is nonassociative. Lewis and Randall (45) fit the thermodynamic data by Staveley (61) to obtain the B factor in equation (101). Using their value: 65 BlnA glfii— = 1. ~ 0.23XAX 3’0 B (222) 3’ Methyl Ethyl Ketone (A) - Carbon Tetrachloride (B) Affsprung and co-workers (46) investigated self-association 3": A. of ketones in carbon tetrachloride. Their results indicate ‘I that methyl ethyl ketone dimerizes in carbon tetrachloride. 1 I They proposed that the dimers are formed by dipole—dipole L interactions of the type shown below. E x? = (.3 R': :/R O = C \R. Such antiparalled arrangements are feasible as long as the groups R and R' are small enough to permit the close approach of the two carbonyl groups. Such is the case for methyl ethyl ketone. This system is an example of a Type (a) associated system. Equations (138) and (139) were fitted to the experimental activity data of Fowler and Norris (29) by finding the value of K which gave the best fit. A non- linear least square computer program was used to accom— plish this task. The best value of K was found to be 5.62 which agrees closely with the value of 5.60 found by Anderson (1). Figure l is a plot which compares the results of equations (138) and (139) to the data. It can be seen that the equations predict the data very well. 66 L0 0.8" 0.6 r- 0.4 (- ACTIVITY 00.0 Figure l. l 1 ‘1 l 0.2 0.4 _ 0.6 0.8 1.0 MOLE FRACTION KETONE Activity as a Function of Mole Fraction for the Methyl Ethyl Ketone - Carbon Tetrachloride Sys- tem. 0 Data, Reference (29); Calculated from Equations (138) and (139). 67 Acetic Acid (A) - Carbon Tetrachloride (B) There is a great deal of evidence which indicates that acetic acid strongly associates in solution. In fact, the asso— ciation is so pronounced that in solutions concentrated in acetic acid, it “polymerizes” to very large clusters , . (53). Spectrosc0pic data of Barrow and Yerger (6) suggest __-__.—-—=— .31., WW... ‘ . I that acetic acid associates to form only dimers when the concentration is below 0.01 mole fraction acetic acid. For these low concentrations, this binary mixture is an example of a Type (a) associated system. A schematic of an acid dimer is shown below. ¢O---HO\ CH3C CCH \OH-—-oo 3 Barrow (6), Davies (17) and Wenograd (67) deter- mined the dimerization equilibrium constant for acetic acid in dilute solutions. Their values varied from 930 (mole fraction)-1 by Barrow to 55,000 (mole fraction)”1 by Davies. Hence, it is not possible to arrive at a meaningful value for K from the literature. Therefore, the equilibrium constant will be determined later from consideration of the diffusion data. Ether (A) - Chloroform (B) Lord and co-workers (47) and Gerbier and Gerbier (30) studied this system using I. R. and Raman spectrosc0pic techniques, 68 respectively. Both investigations showed that the two components associate to form a bimolecular cluster as shown below. C H / 2 5 C2H5 Cl3 This system is an example of a Type (b) associated system. Given this specific association, other physical phenomena can be predicted. Since only total pressure versus composition data are available for this system, it is not possible to determine the equilibrium constant, K, from the activities of each component. But, it is possible to calculate the total pressure of the system by combining equations (99), (181) and (182) to obtain an eXpression for the total pressure in terms of mole fractions of the components and the equilibrium constant. = C C C C W PAAA + PBAA (223) The best value of K was determined by fitting this equa- tion to the total pressure data with the computor program shown in Appendix I. It can be seen from Figure 2 that equation (223) predicts the total pressure very well. III‘EI Icclr‘c E 9 SN 0 I“ o “02.6 Figure 3. 71 I L 2.8 3.0 3.2 3.4 3.6 3.8 RECIPROCAL OF ABSOLUTE TEMPERATURE, I / T. CK") Logarithm K as a Function of l/T for the Ether - Chloroform System. 0 Calculated from Data, References (21), (22), (43), (50), (58); Least Square fit. T 700 r 600 500) 400 t T 300 cc) / mole - 200 AH“, I I00 Figure 4. 72 ‘fio‘. .31 ‘ l J _l g 0.2 0.4 0.6 0.8 )0 MOLE FRACTION ETHER Heat of Mixing as a Function of Mole Fraction for the Ether - Chloroform System. 0 Data, Reference (38): Calculated from Equation (224). 73 Ether (A) — Chloroform (B) - Carbon Tetrachloride (E) To the author's knowledge, none of the physico- chemical prOperties of this system have been studied. However, one can Speculate as to the association that should be observed. Notice that this ternary system can be formed by adding carbon tetrachloride to the binary ether—chloroform system. Assuming that the carbon tetra- chloride is completely inert, its addition to a binary system should have no effect other than dilution. Then the nature of the ether-chloroform association should be the same as a Type (b) associating system, and the value of the equilibrium constant should not change from the value observed for the ether—chloroform mixtures, 2.86 (mole fraction)_l. rm 2:: f: ‘- EXPERIMENTAL METHOD The Open-end capillary method for measurement of tracer diffusion coefficients is one of the earliest and most widely used methOds (4, 56, 66). The principle of this method is as follows. A capillary of known length, L, closed at one end, is filled with a solution of known concentration. Some of the molecules of one component are labelled with an isotopic form of one of the constitu- ent atoms. The capillary is immersed in a solution of the same chemical concentration but which contains no labelled molecules. The volume of the unlabelled bulk solution is much larger than the labelled capillary solution. There- fore, the concentration of labelled molecules remains essentially zero in the bulk solution as they diffuse from the capillary. Natural convection, caused by slight temperature variations or mechanical Vibrations, sweeps the labelled molecules from the neighborhood of the capillary mouth and maintains zero concentration at the mouth. This process is described mathematically by the following differential equation and boundary conditions: 74 '. 8‘10: ..l ,‘myw’p j, 75 ac; 820; ...—...: 'k ()0 Di a2 (a) 2 ac: = O at z = O, t>0 (b) 02 Tn (225) 2. C; = at z = L, t>0 (c) E I C? = C? at 0 = O, 00 (b) (226) ‘k -D* 8C1 l C* at z = L 0>0 (c) i 02 R i ' where R is the resistance of the frit. The length variable, 2, is defined in Figure 6. 81 Figure 6. Diffusion Cell Coordinates The solution of equation (226) follows by separa— tion of variables as shown in Appendix III. The result is: 00 2 C? SIN 8 D? lavg _ 2 n ex _ 2 10 - . p B (a) Cio n=1 8n (8n + SINBnCOTBn) n L2 where 8n is defined by: (227) RD: COTBn = 8n L . (b) The term Ciavg is the average concentration of labelled molecules in the capillary after a time 0. 82 At first inspection, it seems very difficult to obtain a value of D: from experimentally determined 0, L and Ciavg/Cio' However, the problem can be greatly simplified by examination of diffusion in the frit. Trans- port of material across the frit occurs by ordinary molecular diffusion through the liquid-filled pores. Hence, the resistance of the frit, R, is inversely prOpor- tional to the diffusivity, DE. The prOportionality con- stant will be a function of the frit porosity and thickness. As a result, the group RDz/L is independent of D: and is a function of cell geometry only. Now, only the exponential portion of equation (227) (a) contains DE. A computer program, shown in Appendix IV, was written to calculate C? /Cio for various values of iavg 2 * * * 1 tted versus Die/L and RDi/L. Then, Ciavg/Cio was p o DEB/L2 for particular RDg/L as shown in Figure 7. Using a value of RDg/L, determined by calibration of the cells, and eXperimentally determined values of Ciavg/Cio' 0 and L, D: was obtained from this plot. Calibration of the Cells Rathbun and Babb (57) analyzed the results of several investigations and concluded that the tracer diffusion coefficient of pure carbon tetrachloride was -5 The six cells employed in this 1.32 x 10 (cm.2/sec.). study were calibrated using this as a standard. CTAve /C- 0.8) 0.7- 0.6 0.5 0.3 0.!- 0.0O ' * * Figure 7. Ciavg/Cio a 83 R0?/ L =O.I I 805; L=0.0I2 RDi / L-0.o I I I I I 0.2 0.4 0.6 0.8 )0 0:" G/L‘ s a Function of D*i'/L2 and RDE/L. 84 The cells were calibrated as follows. Values of 2 . 1k t * Ciavg/Cio and Die/L were calculated for three experimental observations of each cell. Then three values of the cell constant, RDi/L were determined from Figure 7 and averaged for each cell. Table I is a listing of these average cell constants for the six cells. Table I. Capillary Cell Calibration Results (a) Cell RDE/L 1 0.012 2 0.014 3 0.011 4 0.014 5 0.013 6 0.010 0.012 average (a) average of three determinations . ' i * ’ It is apparent from Figure 7 that Ciavg/Cio is nearly independent of RDg/L when RDE/L is as small as 0.012. In fact, the t 9% deviation of the cell constants listed in Table I corresponds to only a i 0.7% deviation in D? when C? /C# = 0.5. Since this deviation in the i iavg io diffusion coefficient is not significant, the cell con- stants of the six cells were averaged and one value, 0.012, 85 used for all cells. Later, to determine diffusion coefficients, three cells were run simultaneously and the three resulting values of D: were averaged. This process greatly reduces any error caused by using one call con- stant for all cells. Analysis of Radioactive Samples At the end of a diffusion run, the contents of each capillary were analyzed by a Packard Tri-Carb liquid scintillation spectro—photometer. The scintillation solution consisted of 4 grams of PPO (disphenyl oxazole) and 50 milligrams of POPOP (1, 4 of di 2, 5 phenyl-oxasole benzene) dissolved in 1 liter of toluene as suggested by Price (55). It is generally assumed that the count rate measured by a counter is prOportional to the amount of labelled material present in the sample. However, this assumption is not always a good one. Changes in the chemical composition of the sample can cause varied de- grees of absorption (quenching) of the light pulses emitted by the scintillation liquid. Also, some of the older counters give results which are dependent upon the count rate. To study the effects of quenching, experiments were made on mixtures containing various mole fractions. The count rate from a given quantity of solution (0.15 cc.) 86 was determined in each case. The results are shown in Figures 8 and 9 for two systems: methyl ethyl ketone - carbon tetrachloride and chloroform - carbon tetrachloride. Carbon tetrachloride was the labelled component in each case. The linear graphs indicate that there is no change in quenching with composition of the capillary solution. The eXperiments were then repeated for solutions with 1/2 the activity in the first experiments. At each mole fraction, the ratio of counts obtained from samples with two different activities remained constant. This demonstrated that the counts from mixtures having differ— ent amounts of radioactivity but the same composition were quenched in the same proportion. Hence, the observed count rates were proportional to the concentration of labelled molecules and could be substituted directly into equation (227) to obtain the diffusing coefficients. Based on the results for the above two systems, this con- clusion is assumed valid for all the systems of interest. Except for samples very dilute in tracer, a total of one million counts was recorded for each sample. This reduced the statistical error involved in the count rates to a negligible amount, Materials Spectroquality chemicals were used when calibrat— ing the diffusion cells and the viscometer. Analytical reagent grade chemicals were used elsewhere. The 87 8.0 7.0 S” o 9‘ O :b 9 3.0 t 2.0 COUNTS PER MINUTE x I0“; )0 I I I I 0.2 0.4 0.6 0.8 )0 MOLE FRACTION KETONE Figure 8. Count Rate as a Function of Mole Fraction for the Methyl Ethyl Ketone - Carbon Tetrachloride Svstem. 88 COUNTS PER MINUTE x I0" 0.00.0 (fa 0.4 0L6 0:8 10 MOLE FRACTION CHLOROFORM Figure 9. Count Rate as a Function of Mole Fraction for the Chloroform - Carbon Tetrachloride System. 89 refractive indexes of the materials were measured and compared to literature values (65) as a purity check. Carbon - l4 labelled materials were obtained from the Nuclear Equipment Chemical Corporation and the Nuclear- Chicago Corporation. The manufacturers' statement of purity was accepted with no further purity check. Discussion of Error The accuracy of a given run of three cells is governed by the calibration of the cells and the experi- mental precision of the method. The cell calibration is based on the tracer diffusivity of pure carbon tetrachlo- ride suggested by Rathbun and Babb (57). Since they obtained this value from consideration of several investi- gations, it is probably accurate and the resulting cell calibration is, therefore, reliable. However, further evidence of the accuracy of the cell calibration can be obtained from the tracer diffusion data collected in this work. In dilute binary solutions, theory predicts that tracer and mutual diffusivities are equal. For five binary mixtures studied here, intercepts of the mutual and tracer diffusivity curves were determined by graph- ically extrapolating to zero concentration. The inter- cepts agree to within 0.5% for the five systems. Based on these results, errors in the tracer diffusion coeffi— cients resulting from uncertainties in calibration are estimated to be 0.5%. 90 The accuracy of one given run of three cells is, therefore, affected more by the precision of the method. A measure of the precision was obtained by making several duplicate runs for six different solutions. The repro- ducibility of these runs ranged from i 0.7% for pure carbon tetrachloride to t 2% for pure benzene with an average precision of t 1.25%. Therefore, the accuracy of the method was assumed to be within i 2.0%. RESULTS AND DISCUSSION Theoretical From the application of the intrinsic diffusion concept, the medium fixed reference plane and the chemical model of solution nonideality, equations were derived which express the tracer and mutual diffusion coefficients in terms of concentration dependent quantities and friction coefficients-—one for each species in solution. These equations (see Table II) can be empirically fit to experimental diffusion data by adjusting the values of the friction coefficients to give the best prediction. However, one can justly argue that, given enough arbi- trary constants, any curve can be closely approximated. Therefore, the fact that these equations can be made to fit the data does not necessarily constitute support of the concepts from which they were derived. If, on the other hand, the friction coefficients were determined from independent measurements and the equations still predicted the diffusion coefficients accurately, the bases of the equations are probably valid. It is shown in Appendix V that, for the systems studied in this work, the frictional coefficients can be determined strictly 91 hue ®m>rb WCOWUUSHUHQ CCHmwnmemwflA~ M0 Mlhqw:::~am.u IHIH. OHQH...~. 92 mcoflumsvm COHUSMMHQ mo mumfifism . HH mHQMB mam mx mam me if m aNmHV III? I... ..III '2 I l. l U *n— H mx H H 1 ex 0 NHM fin A NHW HM I C AN 35.: inl+fl IMIIIH' H'MH«Q m...¢ mvcm¢ o L HH < fl m C C Ahmav «Mle mm MMh_Mm m + mamx + madx u mdo m 3.8: H“): n mo 0 Ham - m I— L211 I.50L g C) (0.6 8 S O > L450.0 0.2 0.4 0:6 087 100'5 MOLE FRACTION CHLOROFORM Figure 11. Density and Viscosity of the Chloroform - Carbon Tetrachloride System at 25°C. 0 Density Data: 0 Viscosity nzu-a- 99 2.5 x (0° (cmzl sec) DIFFUSION COEFFICIENT I I Figure-12. I l 0.2 0.4 0.6 0.8 )0 MOLE FRACTION BENZENE Tracer and Mutual Diffusion Coefficients for- the Benzene - Carbon Tetrachloride System at 25°C. 0 Tracer Diffusion Data for Benzene; O Tracer Diffusion Data for Benzene, Reference (40); =1Tracer Diffusion Data for Carbon Tetra- chloride; A Mutual Diffusion Data; Equa- tion (88); _ ___ Equation (89); ----- Rnnafinn (90). 100 x 10" (cm’lsec) DIFFUSION COEFFICIENT I.0 0.0 Figure 13. 1 _l I I 0.2 0.4 0.6 0.8 )0 MOLE FRACTION CHLOROFORM Tracer and Mutual Diffusidn Coefficients for the Chloroform - Carbon Tetrachloride System at 25°C. 0 Tracer Diffusion Data for Chloro- form; 0 Tracer Diffusion Data for Carbon Tetrachloride; A Mutual Diffusion Data; Equation (88); ----- Equation (89); ___ Equation (90). 101 (88) and (89) describe the tracer diffusivities of the benzene - carbon tetrachloride and chloroform—carbon tetrachloride systems very well with constant friction coefficients. The behavior of the mutual diffusion coefficients are predicted by equation (90) is also shown on Figures 12 and 13. D = * * AB XADB + XBDA 5162‘ (90) C [ BlnAA A The activity terms were calculated from equations (221) and (222). Equation (90) predicts the mutual diffusiv- ities of the benzene - carbon tetrachloride system very well. As can be seen from Figure 13, this is not the case for chloroform-carbon tetrachloride mixtures where mutual diffusivities calculated by equation (90) differ by as much as 10% from the eXperimental data. This dis- crepancy can be attributed to an error in the measurement of the mutual diffusivities for this system. Although the data form a smooth curve with intercepts which are con- sistent with the tracer diffusion intercepts, attempts to duplicate these data were not successful. Cordes and Steinmeien (14) also measured mutual diffusivities in this system. Their data differ from the data collected here and are not consistent with the tracer diffusivities 102 at infinite dilution. Therefore, it is not reasonable to attribute the poor agreement to fault of equation (90). EXperimental results for the methyl ethyl keton — carbon tetrachloride system are shown on Figure 14. Anderson's (3) viscosity and mutual diffusion data were used for this system. Equation (166) was applied to the tracer diffusion data of carbon tetrachloride. As can be seen from Figure 14, there is excellent agreement between predicted and experimental results. _ kT 03 _ fif— (166) 2 i0 D. 2 .122 _l. _ ___} .1. + ___} (164) A 9 f1 f XA f 11 11 ~ 0 X alnAC . kT 1 B A = * .. __ ___. __ DAB XADB + XBDA + 2 T) E XA EIHEX (167) Labelled methyl ethyl ketone was not available from com- mercial suppliers so the tracer diffusivities of this component could not be measured. However, it is possible to predict these data from the tracer diffusivities for carbon tetrachloride and the mutual diffusivities. The friction coefficient of the carbon tetrachloride molecules, f2, was determined from the tracer diffusivities of carbon 103 x (0" (cm‘/ sec) DIFFUSION COEFFICIENT 0'00.0 01.2 0.4 0.6 0.8 I.O MOLE FRACTION METHYL ETHYL KETONE Figure 14. Tracer and Mutual Diffusion Coefficients fer the Methyl Ethyl Ketone - Carbon Tetrachloride System at 25°C. 0 Tracer Diffusion Data for Carbon Tetrachloride; 0 Mutual Diffusion Data, Reference (3); Equation (166); ----- Equation (167); _____ ___ Equation (164). 104 tetrachloride. Then, equations (164) and (167), together, were fit to the mutual diffusion data to get the best values of f1 and E11' The activity term was calculated from equation (141). The curve for the tracer diffusion of the ketone, in Figure 14, was then calculated from equation (164). Although this curve has not been quantitatively verified, it is qualitatively correct. At low ketone concentrations, the diffusion coefficient is relatively high because the ketone is diffusing as individual mole— cules. As the ketone concentration increases, the diffu- sion coefficient decreases as the ketone molecules form the larger and more slowly diffusing dimer clusters. At the same time, the viscosity of the solution decreases, causing the diffusion coefficient to increase. In this case, both effects cancel and the diffusion coefficient is relatively constant after an initial decrease. Diffusion data for the acetic acid-carbon tetra— chloride system are shown on Figure 14. The viscosity and mutual diffusion data are by Anderson and Babb (3). A reliable value of the equilibrium constant is not available in the literature. So, it was necessary to ~ fll and K from the diffusion data. To and K, equation (164) and (167) were empiri- find f f ~ f 1’ 2’ find f1' 11 cally fit to the mutual diffusivities and the tracer diffusivities of acetic acid. f2 was calculated from the 105 tracer diffusion coefficient of pure carbon tetrachloride and equation (166). The curves shown on Figure 15 were calculated from equations (164) and (167) using K = 4000 (mole fraction)_l, a value within the range of the litera- ture values. It can be seen that, while the equations predict the qualitative behavior of the data, quantitatively there is disagreement. Also, the disagreement is generally greatest at higher acid concentrations. The disagreement can be attributed to two effects. First the acid molecules may associate to form higher order clusters. Since the concentration of these higher clusters increases with acid concentration, it follows that the disagreement between the dimer model and eXperi- ment should also be greatest at the higher concentrations. Second, the chemical model may not accurately describe the activity of the system. The thermodynamic term, as calcu- lated by equation (141), is 0.45 at XA = 0.00642. While this seems abnormally small, the activity data available in the literature cannot be used as a check because they were not taken at low concentrations (73). Thus, it would appear that little can be concluded as to the validity of equations (164) and (167) from the results of this system. However, three points should be restated. First, the equilibrium constant determined from the data is within the range of literature values. Second, 106 MOLE FRACTION ACID x IO' 0.0 0.2 0.4_ 06 0.8 IO T I F I p—Intercept 3.0" | I ‘6 \ \ \ 2.0)1 \ . \ 3 ‘ \\*’;i g N ‘ \. ~-—--—-——— . \- O '6 \-_9__-__ 2 IO Q3 I. I I . J 0.0 0.2 0.4 0.6 0.8 (.0 00 l MOLE FRACTION ACID Figure 15. Tracer and Mutual Diffusion Coefficients for the Acetic Acid Carbon Tetrachloride System at 25°C. 0 Tracer Diffusion Data for Acetic Acid; 0, Mutual Diffusion Data, Reference (3); -------- Equation (161+); ...... ... ...... Equation (167). 107 the diffusion equations can be fit to the qualitative behavior of the data. And third, this system is a rigid test of any model because the diffusion coefficient more than doubles in a concentration span of less than 1%. Thus, it is felt that this system does indicate the validity of the diffusion equations even though they do not predict the data well. Diffusion data for the ether-chloroform system are shown on Figure 16. Again, the mutual diffusion and viscosity data are those of Anderson and Babb (2). fl and E12 were determined from the intercepts of the ether tracer diffusion data and f2 and E12 were determined from the intercepts of the chloroform tracer diffusion data as described in Appendix V. The two values of flz, determined in this manner, were different by approximately 3%. Therefore, an average E12 was used with fl and f2 in equations (191), (192) and (193) to predict the mutual and tracer diffusivities over the whole concentration range. 1 X0 D.=I$I -11___l-:.1_ (191) A ” E-1_ f XA f 12 12 k l XO 1 Dg=l?l_~___%_:_ (192) n 2 f B f I0.0 9° 0 9’ o DIFFUSION COEFFICIENT x I05 (cmz/sec) .45 O 0.0 Figure 16. 108 I I I l 0.0 0.2 0.4 0.6 0.8 LO MOLE FRACTION ETHER Tracer and Mutual Diffusion Coefficients for the Ether - Chloroform System at 25°C. 0 Tracer Diffusion Data for Ether; O Tracer Diffusion Data for Chloroform; A Mutual Diffu— sion Data, Reference (2): ----- Equation (191); Equation (192); ___ _ ___ Equation (193).‘ 109 ~ 0 C 81 A D = x D* + x 0* - 2 53 X12 n A (193) AB A B B A I) ~ alnX The plots of these equations are shown in Figure 16. The activity term in equation (193) was calculated from equa— tion (184). The agreement between data and the predicted values is very good for the mutual diffusivities and the tracer diffusivities of ether. However, a small but con— sistant error is noted for the tracer diffusion of chloro form. There are two factors which could cause this error. First, the chemical model may not adequately describe association in the system. This is not likely for it can be used to predict the tracer diffusivities of ether very well. The second eXplanation for this error is suggested by comparing the effective molecular sizes as indicated by the molar volumes of the pure components. The molar volumes of ether and chloroform are 104 and 81 cc/mole, respectively. Thus, chloroform is a relatively small molecule surrounded by larger molecules of ether and ether-chloroform clusters. As discussed earlier, the friction coefficients should be constant when the medium surrounding a diffusing molecule can be considered a con— tinuum. Since the chloroform molecules are smaller than the molecules in the surrounding medium, the medium cannot be considered a continuum and, therefore, the friction 110 coefficient of chloroform is a function of concentration. This system provides very strong support for equations (191), (192) and (193). Viscosity and density data for the ether-chloro- form-carbon tetrachloride system are presented in Figure 17 and the tracer diffusivity data in Figure 18. For simplicity, the mole fraction of the carbon tetrachloride was 0.5 for all measurements. Since this is a ternary system, the procedure discussed in Appendix V cannot be used to evaluate the friction coefficients from the data taken for this system. Therefore, the friction coeffi- cients found for the binary ether—chloroform system were used. Tracer diffusivities for ether and chloroform were calculated from equation (218) and (219), using these friction coefficients, and plotted in Figure 18. X0 kT 1 1 __1_ 1 * = __ __ _ ___. _ DA n £1 E XA + E (218) 12 12 O kT 1 1 x2 1 D* = — - :— — + :— (219) B 0 f2— f xB f 12 12 It can be seen from this figure that the agreement between theory and eXperiment is excellent for ether but, as in the ether-chloroform system, it is not as good for chloroform. Again, a comparison of molar volumes (104, 97 and 81 (cc/mole), respectively, for ether, carbon 111 0.7 MOLE FRACTION CARBON TETRACHLORIDE ‘7 .. =05 C I6— 0 I.5— -0.6 g" ,4. ' 7)? g ' .8. E» 8 T 8. )- I.3T 0 I; >- 3 ~05 t: LIJ U) 0 8 I.2- g > U A to ' ‘ ‘ 1 0.4 0.0 01 0.2 0.3 0.4 0.5 MOLE FRACTION ETHER Figure 17. Density and Viscosity of the Ether - Chloro- form - Carbon Tetrachloride System at 25°C. 0 Density Data; 0 Viscosity Data. 112 x I05 (cm‘/sec) TRACER DIFFUSION COEFFICIENT 4.0 t MOLE FRACTION CARBON TETRACHLORID L =0.5 |.C) I I I I 0.0 DJ 0.2 0.3 0.4 0.5 MOLE FRACTION ETHER Figure 18. Tracer Diffusion Coefficients for the Ether- Chloroform - Carbon Tetrachloride System at 25°C. 0 Tracer Diffusion Data for Ether; 0 Tracer Diffusion Data for Chloroform; Equation (218); ----- Equation (219). 113 tetrachloride and chloroform) indicates that chloroform is a smaller molecule diffusing through relatively large molecules. The resulting concentration dependence of the chloroform friction coefficient accounts for the disagree— ment between theory and eXperiment. The tracer diffu- sivities of ether and chloroform in this ternary system can be predicted from the tracer diffusivities of the same two components in binary mixtures. The eXperimental tracer diffusion, mutual diffu- sion, density and viscosity data collected in this study are tabulated in Appendixes VI, VII, and VIII, respectively, for detailed reference. CONCLUSIONS It can be concluded from this study that the intrinsic mechanism of liquid diffusion and the concept of the medium fixed reference can be combined with the chemical model of solution nonideality to derive equations which accurately predict tracer and mutual diffusion coefficients in systems whose components associate in simple, specific ways. To make these predictions, it is necessary to specify friction coefficients-—one for each diffusing species. This study shows that in a binary mixture the friction coefficients can be determined directly from the intercepts of tracer diffusion coeffi- cients alone. Further, it is concluded that the friction coefficient of a component, whose molecules are smaller than those of the other components, is a function of composition. This study provides evidence which substantiates the model that in binary mixtures, ether and chloroform cross associate to form bimolecular clusters. Tracer diffusion data indicate that the same cross association of ether and chloroform occurs when nonassociating carbon tetrachloride is added to binary ether-chloroform mixtures. 114 115 The friction factors for ether and chloroform are concluded to be the same in the binary ether—chloroform system and the ternary ether-chloroform—carbon tetra— chloride system. It is also shown that the friction factor of chloroform is concentration dependent for both the binary and ternary systems. The Hartley-Crank equation (equation (71)) and the Darken equation (equation (90)) are shown to be valid for nonassociated binary systems. It is the opinion of this author that the intrinsic model of diffusion is very likely the true mechanism by which diffusion occurs. This opinion is based on the fact that the intrinsic model leads to equations from which mutual diffusivities can be predicted from measurements of tracer diffusivities and viscosity and, in the case of associated systems, from an equilibrium constant which can be determined from thermodynamic data. It would seem more than fortuitous that this would be the case if the intrin- sic model were not correct. FUTURE WORK Theory More work should be done in the deveIOpment of the chemical model to remove one of its primary faults; the assumption that the species form an ideal solution. The assumption was made to derive an expression for the compo— nent activities in terms of measurable quantities. One method of attack can be seen by recalling that the activity of a component is equal to the activity of its monomeric species. Since the component activities can be calculated from vapor pressure data, for example, and since the mole fractions of the species can be deter- mined from spectrosc0pic techniques, the activity coeffi- cients of the monomeric species can be determined as a function of composition. Most likely, the species form a nearly ideal solution, so any simple, thermodynamically consistent set of equations (multi-component van Laar, Porter, etc.) can be fit to the activity coefficient data. The resulting expressions could then be used to modify equation (120) to give a more accurate thermodynamic description of the system of interest. 116 117 The most logical extention of the intrinsic model of diffusion is to multicomponent systems. Kett and Anderson (76) have made a study of the intrinsic model as applied to ordinary diffusion in ternary systems. It would also be of interest to investigate relationships between tracer and ordinary diffusion for these systems. Experimental The accuracy of the capillary technique could be increased if the average concentration (in equation (227)) could be measured several times during the run. A tech— nique of this type was used by Mills (51) for materials labelled with an isotope whose nuclear emission is capable of penetrating the walls of the capillary. Counters were placed outside the capillary. When carbon - 14 or tritium labelled materials are used, absorption of the relatively soft beta particles in even the thinnest capillary walls precludes the use of Mills' method. Instead, the capillary itself could be constructed of a scintillation crystal that is inert to the solvent action of the system of interest. Calcium floride is one material that is resistant to many mixtures of this type studied here. The crystal could be viewed from the side by a photomultiplier detection system and the count rate determined at prescribed time intervals. 118 Anticipated problems are in the areas of contamination of the crystal with radioactive materials and poor conversion of radioactive disintegrations to detectable light pulses. Another area for future experiment work is in the area of activity determination. To check the validit of BlnAA ___ I 81nXA must be known quite accurately. Much of the activity data the diffusion equations, the thermodynamic term, in the literature is very bad. Therefore, it would be desirable to have good activity data on carefully selected systems. APPENDIXES APPENDIX I Fortran Computer Program for the Determination of Equilibrium Constants C THIS PROGRAM FITS TOTAL PRESSURE DATA TO THE TYPE (B) C ASSOCIATION MODEL. NOMENCLATURE: W = WEIGHT FRACTION, C X=MOLE FRACTION, A=VAP. PRES. OF A, B=VAP. PRES. OF B, C PI=TOTAL PRES., AK=EQUIL. CONSTANT SZZXX = 0. SXX = 0. 8Z2 = 0. SXXZ = 0. READ 100, T READ 102, A, B DO 1 J=1, 9 READ 102, WA, PI WB = 1.-WA XA = WA SB - 1.-XA XX=XA*XB A = ((XA-XB)*(A-B)-PI)/(PI-A-B) 22 = z*z szzxx = szzxx +z2*xx sxx = sxx +xx 522 = 822 + 22 sxx2 = sxx2 + xx*xx 1 PRINT 103, XA,XX,Z2 SLOPE = (9.*SZZXX-SXX*SZZ)/(9.*SXX2-SXX*SXX) AK = -SLOPE/(SLOPE+4) PRINT 100, T 2 PRINT 102, SLOPE, AK 100 FORMAT (F7.4) 102 FORMAT (2F7.4) 103 FORMAT (3F7.4) END 120 (l) (2) (3) (4) (5) (6) APPENDI X I I Experimental Procedure Bulk solution of the desired concentration was prepared and placed in a bulk solution container. The container was in turn placed in a constant temperature bath for at least one hour before the run was started. Capillary solution of the same chemical concentration as the bulk solution was weighed out. A small portion of one component consisted of labelled material, added with the component. The capillary solution was degassed by allowing the solution to equilibrate at approximately 30°C. The capillaries and frits were washed three times with acetone and dried. Three cells were assembled. The bulk solution was removed from the water bath. Each cell was filled in turn and placed into the bulk solution. The container was then returned to the water bath and diffusion allowed to proceed for three to six days. After the run, 5 cc. of scintillation liquid were measured into each of three counting vials. 121 (7) (8) (9) 122 Each capillary was then emptied by the following procedure. (a) (b) (C) (d) (e) The frit holder was removed. The cell was inverted into the counting vial. The funnel-shaped portion of the screw cap was filled with scintillation liquid. The metal foil disk was punctured and the capillary flushed with 10 cc. of scintillation liquid. The porous frit was removed from the counting vial. The cells were cleaned, reassembled, filled with capillary solution and emptied as per the above procedure. The vials were placed in the scintillation counter and the activities were determined. If the count rate of any one of the three samples deviated more than 5% from the mean activity of the three, the run was discarded. APPENDIX III Solution of Equation (226) 2 ac; 3 CE 39 = Di 2 (a) 82 ac; 82 = O at z = O, 6>0 (b) (226) ac; 1 -D3 82 = R C: at z = L, 6>O (c) C1 = Cio at 6 = 0, L>z>0 (d) Following the general procedure of separation of variables, a product solution is assumed of the form CI = F(z)- W(6). Then, the differential equation becomes: 1 d2F 1 aw 2 F ___—2 = W d—e— = - X , (III-l) dz 1 a constant. Or, d F + AZF - o (III-2) 123 124 and d? 2 IE 4' D:A q} = 0. (III-3) The product of the solutions of these two equations is: C3 = (klcos(Az) + k2 sin (Az)eXp(-A2D: ). (III-4) Also, so: a: = (kzxcos(Az) - k, Asin (Az)eXp(—A2D; ). (III-5) To satisfy boundary condition (226) (b), k2 from equation (III—5) must be zero. Then with k2 = 0, combine equation (III—4) and (III-5) with boundary condition (226) (c) with the following result: 2 . _ _ l _ 2 * _ -eXp(-A D:6)k,ASIn(AL) — fifif eXp( A Die)klcos(AL) (III 6) Or, after simplifying, = * - COT(AnL) RDiAn (III 7) where the subscript n on A indicates the infinite number of roots. Since the differential equation is linear, this infinite number of solutions may be added. Then, from 125 equations (III-4), * = _ _ Ci 2 kneXp( AnDie)cos(Anz) (III 8) n=1 Now, substitute this eXpression into boundary condition (226) (b), multiply both sides of the resulting equation by cosAmzdz and integrate between 2 = O and L. Then, L L CEOXCOS(Amz)dz =Jr2: knCOS(Anz)COS(Amz)dz (III-9) n: o 0 When m ¢ n, the integral on the right side is zero and when m = n, it is: E + —l— SIN(AnL)COS(AnL)) (III-10) k ( n 2 2An The integral on the left is: —lSIN(AnL). (III—ll) A n So, * L 2Ci0 SIN(An ) k = . o (III-12) n AnL + SIN(AnL)COS(AnL) 126 The final solution is then: l “ SINA L * eXp 0 n=1 AnL + SIN(AnL)COS(AnL) (III-13) _2* ( AnDie)COS(Anz). The solution in this form is not useful, for the concentration ratio is not conveniently measured as a function of position and time. It is more convenient to measure the ratio of average concentration to initial concentration as a function of time. An expression for the average concentration can be obtained by combining (III-l3) and (III-l4). *dz * = 1 (111-14) j? o Ciavg j? dz 0 C The result is: Ciav w SINZBn ——E?§ = 2 1:1 &n(8n + siancoan) exP (227)(a) D?6 2 1 -8n 2 127 where n n (III-l6) Equation (III-7) can now be rewritten as: RD; COTBn = Sn L (227)(b) By defining the new variable, Bn’ the concentration ratio can be written as a convenient function of two dimension— RD? D?8 1 1 less groups, __f and ——§ . APPENDIX IV Fortran Computer Program for Diffusion Coefficient Calculations c R=NUMBER OF TERMS IN THE SUM c w = RD*/L c AN=BETA N PRINT 100 D0 8 J=1, 100 R = J R = R/SO DT = R w = 50.0 R = 1.0 CAVG = 0.0 DO 11 = 1,10 7 AN = (2. *R-l.)*3.l4l6/2. 3 G=AN-(AN/W-COTF (AN))/(l./W+1./SINF(AN))**2) IF (ABSF(AN/W-COTF(AN))-.00001)2,2,4 4 AN = G GO TO 3 2 AN = G SUM = (SINF(AN))**2/(AN*(AN+(SINF(AN))*(COSF(AN))))* lEXPF(-AN*AN*DT) CAVG=CAVG+SUM IF(SUM/CAVG-.OOOOl)5,5,6 6 R=R+l.0 GO TO 7 5 CAVG = 2.*CAVG PRINT lOl,CAVG,W,DT,R R=1. CAVG=0.0 1 w=w+5.0 8 CONTINUE 100 FORMAT(*0*10X*CAVG*1OX*W*12X*DTL2*lOX*R*) 101 FORMAT (*0*,4(7x,F7,4)) END 128 APPENDIX V Determination of Friction Coefficients From Tracer Diffusivities If the diffusion relations derived in this work are to be used for quantitative predictions, the friction coefficients, fi, must be specified. For components of a nonassociated system or for a nonassociated component of an otherwise associated system, the tracer diffusiv— ities are given by equation (79) kT I 79 D1 PIT ( ) 1 Given viscosity and tracer diffusion data, one can calcu- late fi at any concentration. The Situation is more complex for an associated component of a mixture. When component A associates it— self to form dimer clusters, 0 -kT —l-—l-——-l—+—l——. (164) Here it is not possible to calculate a value for both fl and E11 from viscosity and tracer diffusion data at any one concentration. However, the frictional coeff1c1ents 129 130 can be calculated directly from the eXperimental tracer diffusion intercepts. Using equation (135) and l' Hospital's rule, it can be shown that O I . X limit 1 _ XA+O x ‘ 1' (V'l) P It follows directly from equations (V—l) and (164) that limit D* = -——. (V—2) XA+O A flr‘ Therefore, fl can easily be calculated from the intercept of DA at XA = 0. Also from equation (135), 0 limit f1 _ 1 + (1+4K)1/2 (V_3) XA+1 xA 1 + (l+4K)l/2 + 4K and hence, 1/2 I a. = —- 751-1 12‘1302 + 4K + ——: )1 Xz+l 1 fl]. 1 + ( + 11 ~ The friction coefficient, fll' can now be calculated from this equation after fl has been determined from equation (v-2). 131 The following limiting eXpressions are applicable when the two components associate with each other to form a dimer cluster: limit D* = ——— (V-S) xA+1 A nf1 ' . . kT llITllt D5 = FF- , (V-6) XB+1 ' 2 limit D1: = 1%: K i 1 f—l— + 715— (v-7) X +0 1 f A 12 and . . kT 1 1 K 11m1t D* = — + :— (V-8) XB+O B r] K + I I; £12 After fl and f2 have been calculated from equations (V-5) and (V-6), f can be calculated from either equation 12 (V-7) or equation (V-8). If different values are obtained, an average E12 can be used. 0.0 0.216 0.26 0.34 0.523 0.677 0.82 1.00 0.0 0.08 0.18 0.29 0.42 0.50 0.75 0.76 0.98 1.00 APPENDIX VI EXperimental Tracer Diffusion Data at 25°C. Table III. Tracer Diffusion Data for the Benzene (A) - Carbon Tetrachloride (B) System D3 x 105 (cmZ/Sec) DE x 105 (cm2/sec) -- 1.32 1.57 _- 1.58 -_ -- 1.48 1.76 __ -- 1.71 2.00 _- 2.14 __ Table IV. Tracer Diffusion Data for the Chloroform (A) — Carbon Tetrachloride (B) System 2 DA x 105 (dmz/sec) DE x 105 (cm /sec) -_ 1.32 1.57 -- _- 1.49 1.82 -- __ 1.73 2.02 -- 2.23 -- -_ 2.04 __ 2.16 2.44 -- 132 133 0.0 0.027 0.027 0.101 0.201 0.642 0.642 0.0 0.04 0.42 0.44 0.65 0.82 0.85 0.95 1.00 Table V. Tracer Diffusion Data for the Methyl Ethyl Ketone (A) - Carbon Tetrachloride (B) System X D* x 105 (cmz/sec) A B 0.0 1.32 0.25 1.55 0.40 1.88 0.50 2.00 0.73 2.35 Table VI. Tracer Diffusion Data for the Acetic Acid (A) - Carbon Tetrachloride (B) System DA x 105 (cmz/Sec) DE x 105 (cmZ/sec) __ 1.32 2.31 2.33 1.80 1.65 1.27 1.28 Table VII. Tracer Diffusion Data for the. Ether (A) - Chloroform (B) System D; x 105 (cmZ/sec) DE x 105 (cm2/sec) -- 2.44 2.15 2.44 3.05 -- -- 2.65 -- 2.80 6.23 -- -- 3.55 —- 4.13 8.75 -- 134 Table VIII. Tracer Diffusion Data for the Ether (A) - Chloroform (B) - Carbon Tetrachloride (C) System [xC — 0.5] XA Dg x 105 (cmZ/sec) DE x 105 (cmZ/sec) 0.0 -- 2.03 0.1 2.32 -- 0.1253 -- 1.99 0.2505 2.81 -- 0.2510 -- 2.11 0.3411 -- 2.13 0.3751 -- 2.17 0.3742 3.38 -- 0.4511 - 2-30 0.5014 4.62 -- APPENDI X VI I EXperimental Mutual Diffusion Data at 25°C. (X is the average concentration of the solutions above and below the boundary in the diffusion cell.) Table IX. Benzene (A) Mutual Diffusion Data for the - Carbon Tetrachloride (B) System Chloroform (A) 5 2 XA DAB x 10 (cm /sec) 0.0150 1.450 0.2548 1.439 0.4936 1.451 0.7656 1.755 0.9964 1.945 Table X. Mutual Diffusion Data for the — Carbon Tetrachloride 5 2 XA DAB x 10 (cm /Sec) 0.0265 1-505 0.4311 1.644 0.7389 1.752 0.8255 1.796 0.0966 1.881 0.9418 1.972 135 (B) System APPENDIX VIII Experimental Density and Viscosity Data at 25°C. XA 0.0 0.0301 0.1711 0.2953 0.5606 0.7864 1.0 C xA 0.0 0.1118 0.2318 0.4915 0.8450 1.0 Table XI. Density and Viscosity Data for the Benzene (A) - Carbon Tetrachloride (B) System n (Centipoise) p (g/cc) 0.8888 1.5945 0.8794 1.5639 0.8393 1.4745 0.7992 1.3902 0.7205 1.2036 0.6563 1.0386 0.5970 0.8736 Table XII. hloroform Density and Viscosity Data for the (A) - Carbon Tetrachloride (B) T1(Centipoise) System 0 (g/cc) 0.8888 0.8263 0.7611 0.6608 0.5680 0.5401 1.5945 1.5744 1.5600 1.5322 1.4955 1.4798 Table XIII. Density and Viscosity Data for the Ether (A) - Chloroform (B) - Carbon Tetrachloride (B) System XA r1(Centipoise) O (g/cc) 0.0 0.6475 1.5307 0.1031 0.6298 1.4428 0.2501 0.5702 1.3288 0.3495 0.5189 1.2509 0.5010 0.4343 1.1370 136 AH m AH im iM iN io NOMENCLATURE algebraic complement of the element, 0 ., in the G determinant k3 activity constant in equations (100) and (101) concentration of a Species, (moles/cc) number of components concentration of a component (moles/cc) mutual diffusion coefficient (cmz/sec) tracer diffusion coefficient (cmZ/sec) intrinsic diffusion coefficient (cm2/sec) a diffusion coefficient de ined relative to the volume fixed plant (cm /Sec) friction coefficient (cm) determinant of stoichiometric coefficients heat of mixing (calories/mole) heat of reaction (calories/mole) molar flux relative to the medium fixed reference plant molar flux relative to the mass fixed reference plane molar flux relative to the number fixed reference plane molar flux relative to the component fixed reference plane 137 ’3 W W 1"!