TN E5333 “J LIBFEARY a % Pavia: E13593 Sm: - I V This is to certify that the dissertation entitled Ion Transport Processes presented by John H. Leckey has been accepted towards fulfillment of the requirements for Ph 0 D . degree in ChemiS try M Major professor Date 7/494, MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 MSU LlIlARllS “- RETURNING MATERIALS: Place in book drop to remove this checkout from your record. FINES will be charged if book is returned after the date stamped below. ION TRANSPORT PROCESSES By John H. Leckey A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1981 ABSTRACT ION TRANSPORT PROCESSES By John H. Leckey Ions flow due to electrochemical potential gradients so as to minimize the production of entropy. In a homo- geneous substance the electrochemical potential can be separated into a chemical potential gradient, the driving force for chemical diffusion, and an electrical potential gradient, the driving force for electrical migration. The development of a macrosc0pic set of partial differential equations from nonequilibrium thermodynamics, hydrodynamics, and electrostatics, which describes the isothermal trans— port of ions, is reported here. This system of equations contains a set of transport coefficients which result from the partial separation of the electrical and chemical ef- fects involved in ion transport. The characterization of these coefficients and the solution of this system of equa- tions for liquid—liquid Junctions and double layer relaxa- tions is also reported here. The solutions are in terms of numerical finite difference computer simulations and approximate analytical equations. All activity coefficients John H. Leckey have been retained in the numerical solutions and whenever possible in the analytical solutions. The time range of the dependence of the cell potential on single-ion activity coefficients has been delineated for several Ag/AgCl concentrationcells containing a single chloride electrolyte. A simple formula is obtained for the cell potential as a function of time, and in terms of Onsager coefficients, ionic valences, and the permittivity. The results have two major implications: (1) They can be used as guidelines for choosing appropriate electrolytes, time domains, and boundary conditions for future experi- ments to determine single-ion activity coefficients; (2) They suggest complicated interdependences between elec- trical transport processes, on the one hand, and single— ion activity coefficients, on the other. An analytical solution in terms of a perturbation series has been derived for the equilibrium double layer in the presence of a charge on the electrode. Activity coefficients appear explicitly in the solution. An ap- proximate analytical solution has been obtained, using Laplace transforms, for double layer relaxation following rapid electrode charging. The solution results in a simple formula for the potential difference across the double layer as a function of time. The numerical solu- tion of this problem has been used to delineate the range of validity of the time dependent analytical solution. Results are compared to previous studies. To Pam ii_ ACKNOWLEDGMENTS I wish to thank the Department of Chemistry for the financial support I have received from teaching assist- antships. I also wish to thank the General Electric Foundation and Amway Corporation for awarding me fellow- ships through the Department of Chemistry for the summers of 1979 and 1980, respectively. I am most grateful to Professor Frederick H. Horne whose teachings of nonequilibrium thermodynamics have been the cornerstone of this work. I appreciate his en- couragement and useful suggestions, which aided in the completion of this dissertation. I would like to thank Dr. Thomas H. Pierce for helpful discussions about computer programming, and especially for making available the matrix inversion package used in this work. I wish to thank Dr. Tom Atkinson for providing his curve plotting package MULPLT used to prepare the figures in Chapter 2, and Folim Halaka and Rob Thompson for help- ing me use this package on a PDP-ll computer. I am indebted to Bruce Borey for reading and comment- ing on several manuscripts, and to Dr. Stanley R. Crouch for critically reviewing this dissertation on very short notice. iii Most of all, I thank my wife, Pam, for her support, encouragement, and unbridled optimism. Her understanding throughout difficult circumstances has made the completion of this work possible. iv TABLE OF CONTENTS Chapter Page LIST OF TABLES. . . . . . . . . . . . . . . . . . . viii LIST OF FIGURES . . . . . . . . . . . . . . . . . . x I. INTRODUCTION. . . . . . . . . . . . . . . . . . l A. Phenomenology of Ion Transport. 1 B. Objectives of the Research. 3 C. Plan of the Dissertation. 5 2. ION TRANSPORT EQUATIONS . . . . . . . . . . 8 A. Introduction. . . . 8 B. Continuity Equations. 9 C. Nonequilibrium Thermodynamic Equations . . . . . . . . . . . . . . . . . 9 D. Electrochemical Potentials and Activity Coefficients . . . . . . . . . . . II E. Electrostatic Equations . . . . . . . . . . 1A F. Combined System of Equations. . . . . . . . 15 G. Transformation of Concentration Variables . . . . . . . . . . . . . . . . . 17 H. Discussion of Transport Coefficients for Transformed Equations . . . . . . . . . 20 3. ION TRANSPORT BOUNDARY CONDITIONS . . . . . . . Al A. Introduction. . . . . . . . . . . . . . . . Al B. Derivation of Kinetic Boundary Conditions. . . . . . . . . . . . . . . . A3 Chapter Page C. Applications of Kinetic Boundary Conditions. . . . . . . . . . . A8 A. NUMERICAL FORMULATION OF THE TRANSPORT EQUATIONS . . . . . . . . . . . . . . 50 A. General Finite Difference Approach for Parabolic Partial Differential Equations . . . . . . . . . . . . . . 50 B. Difference Equations for Ion Transport . . . . . . . . . . . . . . . . . SLl 1. Space and Time Grids. . . . . . . . . . 5A 2. Application to Ion Transport Derivatives . . . . . . . . . . . . . . 57 3. 'Boundary Conditions . . . . . . . . . . 58 A. System in Matrix Form . . . . . . . . . 62 C. Computer Simulation Discussion. . . . . . . 63 5. LIQUID-LIQUID JUNCTIONS AND SINGLE- ION EFFECTS . . . . . . . . . . . . . . 68 A. Introduction. . . . . . . . . . . . . . . . 68 B. Initial and Boundary Conditions . . . . . . 73 C. Simplifying Approximations. . . . . . . . . 7” D. Analytical Solution . . . . . . . . . . . . 75 E. Full Cell Equation Results. . . . . . . . . 79 F. Numerical Solution. . . . . . . . . . . . . 81 G. Conclusions . . . . . . . . .i. . . . . . . 95 6. DOUBLE LAYER RELAXATION AT BLOCKED ELECTRODES. . . . . . . . . . . . . . . 97 A. Introduction. . . . . . . . . . . . . . . . 97 B. Equilibrium Solution. . . . . . . . . . . . 100 C. Time Dependent Solution . . . . . . . . . . 120 vi Chapter D. Computer Simulation Results and Discussion. . . . . . . . E. Comparison with Previous Results. 7. SUGGESTIONS FOR FUTURE WORK A. Future Thermodynamic Studies. B. Program Improvement Suggestions APPENDIX A - TRANSPORT COEFFICIENTS APPENDIX B - DIMENSIONLESS VARIABLES USED BY PROGRAM "IONFLO" . . . . . . . . APPENDIX C - TRANSPORT EQUATIONS IN FINITE DIFFERENCE FORM. . . . . . . APPENDIX D - NEWTON'S METHOD AS USED BY PROGRAM "IONFLO" ' . APPENDIX E - LISTING OF PROGRAM IONFLO. APPENDIX F - SOLUTION OF SYSTEMS OF LINEAR PARTIAL DIFFERENTIAL EQUATIONS APPENDIX G - ANALYTICAL SOLUTION OF D FOR TIME DEPENDENT DOUBLE LAYER RELAXATION . . . . . . REFERENCES. vii Page 133 1A6 15A 15A 156 158 161 163 167 170 197 200 205 Table 6-2 6-3 LIST OF TABLES Comparison of Experimental Cell Potentials (EEXP) for Cell (A) with Cell Potentials (ECALC) From Equa- tion (5-20) Comparison of Equation (6-72) and Computer Simulation for 0.1 M KCl. A¢* Indicates Equation (6-72) Results. -1 a: a: 6 x 10 1 cm. 6 S of = 10‘7 C/cm2. Comparison of Equation (6-72) and Computer Simulation for 0.1 M NaCl. A¢* O. a Indicates Equation (6-72) Results. -1 6 x 10 1 cm. 6 S Q° = 10'7 C/cm2. Comparison of Equation (6-72) and Computer Simulation for 0.1 M HCl. A¢* a: a Indicates Equation (6-72) Results. -1 6x106 1 cm. S Q° = 10‘7 C/cm2. viii Page 82 139 . 140 lUl Table 6-A 6-6 Page Comparison of Equation (6-72) and Computer Simulation for 10"3 M KCl. A¢* Indicates Equation (6-72) Results. o = 6 x 106 s‘l. Q° = 10'7 C/cm2. a = 1 cm. . . . . ... . . . . . . . . . . . 143 Comparison of Equation (6-72) and Computer Simulation for 10'3 M NaCl. A¢* Indicates Equation (6-72) Results. 3 6 x lo6 s’l. Q° = 10‘7 C/cm2. a = 1 cm. . . . . . . . . . . . . . . . . . 1AA Comparison of Equation (6-72) and Computer Simulation for 10'3 M HCl. A¢* Indicates Equation (6-72) Results. a 6 x 106 8'1. Q° = 10"7 C/cm2. a 1 cm. . . . . . . . . . . . . . . . . . 1A5 ix LIST OF FIGURES Dependence of YSS at B equals zero. (0) NaCl, (d) KCl Dependence of YDD at B equals zero. (c) NaCl, (d) KCl Dependence of YSD at B equals zero. (0) NaCl, (d) KCl Dependence of YDS at B equals zero. (c) NaCl, (d) KCl Dependence of YSS for NaCl. (a) B o, (c) B = 0.5 M' Dependence of YDS on ionic strength (a) HCl, (b) LiCl, on ionic strength (a) HCl, (b) LiCl, on ionic strength (a) HCl, (b) LiCl, on ionic strength (a) HCl, (b) LiCl, on ionic strength -005 M-l, (b) B = on ionic strength 1 for NaCl. (a) B = -o.5 m‘ , (b) B = 0, (c) B = 0.5 M- Dependence of YSE/S on ionic strength. (a) HCl, (b) LiCl, (0) NaCl, (d) KCl. Page 25 27 29 31 3A 36 38 Figure 2.8 5.2 5.3 Page Dependence of YDE/S on ionic strength. (a) HCl, (b) L101, (0) NaCl, (d) KCl. . . . A0 Interfacial region at a phase boundary. . . A2 Uniform grid mesh for finite dif- ference equations . . . . . . . . . . . . . 52 Nonuniform grid mesh for improved ac- curacy near boundaries and for more efficient approach to steady-state equilibrium . . . . . . . . . . . . . . . . 56 Flow diagram for program IONFLO . . . . . . 65 Predicted charge density surface near the liquid-liquid interface at short time for a NaCl concentration cell operating between concentrations of 0.11 M and 0.09 M. B is zero for this simulation (y+ = y_). The time row at t = 0 has been omitted since it would obscure the lower portion of the surface. . . . . . . . . . . . . . . 86 Constant time planes from Figure 5.1. (a) 5 x 10.98, (b) 1 x 10-88 10-85 . . . . . . . . . . . . . . . . . . . 88 , (c) 2 x Charge density X§° cell position for different values of B at t = l x 10-85 for a NaCl concentration cell operating xi Figure Page between concentrations of 0.11 M and 1 0.09 M. (a) B = —l.OO M- , (b) B = O, (c) B = 1.00 M.1 . . . . . . . . . . . . . 90 5.A Cell potential vs time for cell. (A) with M = Na, SL 0.11 M, and 1.00 M'l, (b) l. S = 0.09 M. (a) B R B = o, (c) B = -1.00 M’ 92 5.5 Coalescence time Kg YDE for 0.1 M solutions, with 5% criterion. (a) HCI, (b) BaCl2, (C) KCI, (d) NaCl, (e) LiCl. . . . . . . . . . . . . . . . . . 9A 6.1 Contributions to s' from S2 and s3 for a 0.1 M NaCl solution with an 6 electrode charge of 2 x 10' C/cm2. b = -0.1; (A) s', (B) s (C) s . . . . . . 113 2’ 3 6.2 Contributions to d' from d1, d2, for a 0.1 M NaCl solution with 6 and (13 an electrode charge of 2 x 10- C/cm2. b = -0.1, (A) dl, (B) d1. (C) d2. d . . . . . . . . . . . . . . . . . . . (D) 3 115 6.3 Dependence on a of D at x = -a/2 as a function of time as predicted by Equation (6-65) for a particular NaCl 2 -6 solution. SB = 10- M, Q0 = 10 C/cm2, (A) a = 3 X 107 8-1, (B) a = l x 107 s_¥, (c) a = 6 x 106 3-1. . . . . . 126 xii Figure Page 6.A Concentration dependence of the relaxation of D at x = -a/2 for an electrode charging time constant of 107 s'1 as predicted by Equation 2 (6-65). (A) SB = A x lo' M, (B) SB a 2 x lo‘2 M, (c) SB = 1 x lo‘2 M. . . . 128 6.5 Effect of time on D versus z/a profiles as predicted by Equation (6-67). Q0 = -lo‘7 C/cm2, a = 6 x lo6 s‘l, = 5 x 1075, (A) lo'7s, (B) lo'8s, YDB (c) lo‘9s . . . . . . . . . . .-. . . . . . 131 6.6 Computer simulation of charge density difference versus distance from 6s for two 0.1 M NaCl solutions. a = 6 x 106 8-1, electrode at 10‘ Q0 = "10-7 C/cm2, 9 corresponds to B = -1 M‘1 B = 1 M'1 . . . . . . . . . . . . . . . . . 135 and p* corresponds to 6.7 Effect of B on Na+ and C1- activity coefficients. (1) yNa+ if B = 1 M“1 and yCl_ if B = -l M“1, (2) y: for NaCl as measured by Brown and MacInnes 1 [1936], (3) YCl- if B = 1 M‘ and yNa+ if B = -1 M‘l. . . . . . . . . . . . . 137 xiii Figure 6.8 6.9 6.10 Page Electrode charge versus time profile assumed by Newman [1973] for analytical solutions to the Nernst-Planck equa- tions . . . . . . . . . . . . . . . . . . . 1A8 Comparison of the experiments of Anson _£‘_l. [1969] (A), simulations using IONFLO (B), and Newman's [1973] ana- lytical predictions (C) for the relaxa- tion potential of HCl at a blocked electrode. a = 6 x 106 3-1, Q0 = 10'5 C/cm2. Time axis is logarithmic . . . 150 Comparison of the experiments of Anson EE.§l- [1969] (A), simulations using IONFLO (B), and Newman's [1973] analytical predictions (0) for the relaxation potential of KCl at a blocked electrode. a = 6 x lo6 s‘l, Q0 a lo‘5 C/cm2. Time axis is logarithmic . . . . . . . . . . . . . . . . 152 CHAPTER 1 INTRODUCTION A. Phenomenology of Ion Transport Application of a current impulse to an electrolyte solution causes the flow of ions. Positive ions move in the same direction as the current and negative ions move in the opposite direction. When the ions reach an elec- trode one of two possible events will occur. One possibility is the transfer of the ions' charge to and/or across the electrode-solution interface. This is done by either specific adsorption, an electron transfer reaction, or, as in the case of some intercalated electrodes, the simple migration of the ionsixux>the electrode (Armand 33 3;. [1978]). The other possibility is that the ions will re- tain charge and be somehow blocked from interacting with the electrode (Chang and Jaffé [1952]). In this case the current pulse is usually referred to as a charge impulse or coulostatic pulse (van Leeuwen [1978]). If the current is turned off and the interface remains blocked, the concen- Itration of the ions at the electrode will not increase indefinitely. This is so despite the fact that the charge on the electrode will be attracting all of the ions of the opposite charge to the electrode. The balancing effect, of course, is that the ions will tend to diffuse from the high concentration near the electrode to a lower concentra- tion in the bulk of the solution. Eventually these effects result in a fixed concentration profile of each ion and a time invariant electric field. Electrolyte solutions can also generate an electric field without any type of electric pulse. A diffusing electrolyte solution creates an electric field and,there- fore,a potential difference between any two points in the solution along the direction of the diffusion. This is because the positive and negative ions do not diffuse at exactly the same rate. Different masses, volumes, and interactions with the solvent cause different rates of diffusion. Of course, one type of ion does not get very far ahead of or behind the other type. The electrical force created by their separation keeps them from separating very far in the solution. The potential difference caused by this effect is nevertheless significant. This is the origin of the liquid Junction potential present in certain types of electrochemical cells. Although these two events seem only remotely related, they can be treated theoretically using the same non- equilibrium thermodynamic formalism. In the language of nonequilibrium thermodynamics, the ions will flow due to electrochemical potential gradients so as to minimize the production of entropy in the system (Onsager [1931]). Besides the above effects, the same set of nonequilibrium thermodynamic, hydrodynamic, and electrostatic equations describes many other important processes. The flow of ions through membranes, the movement of electrons and holes in semiconductors, and the performance of completely solid and polymer batteries can all be described macroscopically using thermodynamics, hydrodynamics, and electrostatics. The unifying concept is the competition between electrical migration and chemical diffusion. In a general way this dissertation deals with all of these effects, and in par- ticular with the detailed theoretical study of double layer relaxation at blocked interfaces, liquid Junction poten- tials, and how single-ion activity coefficients are related to both of these. B. ObJectives of the Research Although ion transport has been treated extensively in theoretical research (Turq and Chenla [1977], Vallet and Braunstein [1977], and Brady and Turner [1978] are some of the relatively recent studies.), it has not previously been approached using a full, rigorous nonequilibrium thermo- dynamic analysis. Rarely have activity coefficients or cross- diffusional effects been included in ion transport equa- tions. Their omission is valid for many types of non- electrolyte transport processes because in that case their influence is often small. In electrolyte solutions, however, they are significant for concentrations as low as 0.01M. In a 0.01M solution of BaCl2 in water, for instance, the cross-Onsager coefficient is 20% of the Ba Onsager coefficient. The mean molar activity coefficient of HCl at 0.01M is 0.906, £1 departure of nearly 10% from ideality. Both the cross-Onsager coefficient and the mean molar activity coefficient rise rapidly as the concentra- tion increases above 0.01M. In an effort to determine the effect of these terms on measurable quantities, such as potential differences, they have been retained throughout the numerical studies of the dissertation. They have also been retained whenever possible in analytical derivations. In a related obJective, the role of single-ion activity coefficients in electrolyte transport processes has been in- vestigated theoretically. Although they have not been determined experimentally yet, information about them is_ essential for further interpretation of many phenomena. As Conway [1978] has noted, detailed explanation of phase transfer reactions, electrodialysis, and ion-solvent inter- actions all require properties of individual ions. Hall [1971], [1973], [1977], and [1980] has extensively studied single-ion activities, and their relationship to the equilibrium thermodynamics of charged interfaces. The fact that single-ion activity coefficients deviate sig- nificantly from unity at concentrations larger than 0.01M, yet are undetermined, indicates that conventional steady- state experiments do not depend on them in an obvious way. With this in mind, most of the research has focused on time dependent phenomena. One of the obJectives has been to discover situations which might lead to experimental de- termination of single-ion activity coefficients. A fourth obJective has been to solve transport equa- tions analytically whenever possible. The analytical re- sults, although less accurate than the numerical ones, have the advantage that the final equation specifically reflects the effect that the input parameters (ionic valences, activity coefficients, Onsager coefficients, egg.) have on the final result. It is hoped that the results presented here will lead to a greater understanding of ion transport in general, in whatever type of system it occurs. More specifically, it-is hOped that these results will stimulate research in experimental and theoretical studies of time dependent ion transport phenomena. C. Plan of the Dissertation Chapter 2 begins with those fundamental equations of hydrodynamics, nonequilibrium thermodynamics and electro- statics necessary to formulate the ion transport equations. The rest of Chapter 2 deals with formulation of suitable ion transport equations and interpretation of the transport coefficients. Because of the tremendous importance of boundary conditions on ion transport phenomena, Chapter 3 is devoted entirely to their derivation and application. Chapter A describes the combination of the ion transport equations and boundary conditions into a general finite difference computer program. This program is the first of its type to retain all thermodynamic terms. Chapters 5 and 6 describe the application of the ion transport equations to two specific problems. Both chap- ters present analytical as well as numerical results. -Chapter 5 follows very closely Leckey and Horne [1981], and deals particularly with single-ion activity coefficients and their dependent liquid Junction potentials. The time range of the dependence of the cell potential on single- ion activity coefficients is delineated for several Ag/AgCl concentration cells containing a single chloride elec- trolyte. A simple formula is obtained for the Cell poten- tial as a function of time, and in terms of Onsager co- efficients, ionic valences, and the permittivity. Chapter 6 describes transient and equilibrium double layers created by a rapid current pulse at a blocked electrode. The equilibrium ion distribution equations are solved using a perturbation technique. At infinite dilution the solution approaches the Gouy-Chapman result. An approximate equa- tion for double layer relaxation is solved using Laplace transforms. From this a simple formula is obtained for the potential difference across the double layer as a function of time. This result and computer simulation results are compared to previous theories and the experiments of Anson 23 al. [1969]. The need for improved computer programs which account for concentration dependence of the dielectric coefficient, specific ion adsorption, and more sophisticated boundary conditions is discussed in Chapter 7. The possibility of studying liquid Junction potentials using thermal dif- fusion is also discussed there. CHAPTER 2 ION TRANSPORT EQUATIONS A. Introduction The macroscopic equations necessary for the description of ion transport are assembled from the fields of hydro- dynamics, electrostatics, and nonequilibrium thermodynamics. The systems dealt with in this dissertation are isotropic, nonreacting (except possibly at interfaces), isothermal, binary electrolyte-solvent mixtures not subJect to magnetic or gravitational fields. All concentration and electric field gradients are assumed to be in one dimension. Electro- neutrality is not assumed. The fundamental equations employed are not new. There- fore, more consideration will be given to simplifications, transformations, and in latter chapters solutions, rather than to derivations of these equations. Details of the derivations and assumptions involved in obtaining the start- ing equations are available from de Groot and Mazur [1962], Fitts [1972], Hasse [1969], Duckworth [1960], and Lewis and Randall [1961]. B. Continuitnyguations The concentrations of any two of the three components (solvent, cation, anion) are independent in a charged system. For the positive and negative ions as the two independent species, the relevant continuity of mass equations are (ac+/at) -(a/ax)(J+H + c+vo), (2-1) (ac_/at) -(3/3x)(J_H + c_vo) . (2-2) Here t is time, x is the space coordinate, c+ and c_ are the positive and negative ion concentrations, J+H and J_H are the Hittorf diffusion fluxes, and v0 is the velocity of the solvent. A Hittorf diffusion flux has as reference velocity the velocity of one of the components. In many cases it is useful to choose the velocity of the solvent for the reference velocity since it is often small enough to be neglected. C. Nonequilibrium Thermodynamic Equations The nonequilibrium thermodynamic equations which relate the fluxes in Equations (2.1) and (2.2) to electrochemical potential gradients are the Onsager flux equations for 10 coupled diffusion, l L, I + - 2++(an+/ax) + 2+_(an_/3x), (2-3) I c_. ll £_+(3u+/3x) + £__(3u_/3x) . (2-A) u+ and u_ are the electrochemical potentials of the cation and anion respectively, and the t are the Onsager co- 13 efficients.- These flux equations have been used in preference to the simpler Nernst-Planck flux equations for two main reasons. In a very dilute aqueous electrolyte solution, the ions are so far apart that long range Coulombic forces are the principal interactions, and specific effects are negligible. In this circumstance the flow of an ion is proportional to gradients only of its own properties and is not influenced by the presence of other distant ions. This reasoning leads to the development of the Nernst- Planck flux equations. At higher concentrations activity coefficients become specific, ionic conductances are no longer additive, and the Nernst-Hartley equation for the diffusion coefficient fails. Thus, intuitively, the flow of an ion must be influenced by the other ions present and the gradients of their properties as well. The macroscopic theory that includes these effects is nonequilibrium thermo- dynamics, which leads to the flux equations, (2-3) and (2-A). 11 A second reason for choosing these flux equations is the RiJ themselves. They are more fundamental than the common transport coefficients for special cases, such as the equivalent conductances or diffusion coefficient. It has been shown explicitly by Miller [1966] that any isothermal electrical transport process, no matter how complex, can be completely characterized once these A are known as 13 functions of concentration. D. Electrochemical Potentials and Activity Coefficients By strict thermodynamic definition u+ and u_ in Equa- tions (2-3) and (2-A) should be termed chemical potentials instead of electrochemical potentials. Since the only work terms being dealt with are electrical, however, Guggenheim's [1929] precedent will be followed and the term electrochemi- cal potential will be used for the "entire" potential. The electrochemical potentials for the ions are m + RT tn(c+y+) + z+F¢ , (2-5) 1: + II 11+ u = u_m + RT £n(c_y_) + Z_F¢ (2-6) where, ui0° denotes the infinite dilution standard state, R is the gas constant, T is the absolute temperature, Z1 is the charge number, F is Faraday's constant, and ¢ is the electric potential. The activity coefficient yi is on the molarity 12 scale, and c1 is the molarity of component i. The chemical potential derivatives of Equations (2-3) and (2-A) in the variables electric field, concentration, and ionic strength are (3u+/3x) = (RT/0+)(ac+/8x) + RT(v/2u+)(M+B)(BIc/3x) + z+F(3¢/3x) , (2-7) (3u_/8x) = (RT/c_)(30_/3x) + RT(v/ZV_)(M-B)(3Ic/3x) + z_F(3¢/3x) , (2-8) where IC is the ionic strength on the molarity concentra- tion scale, Ic = % (z+2c+ + z_2c_) , (2-9) and M and B are defined by M = (axing/316%,P , (2-10) B = (BAnGi/aIC)T’P . (2-11) yi has its usual definition v v+ v- Y+ = y+ 'y 3 (2‘12) 13 and 6+ is the mean molar activity coefficient deviation first introduced by Frank [1963]. /y (2-13) The exponents v and v_ are the positive and negative ion 4. stoichiometric coefficients, and v = v+ + v_. The motivation for this way of treating activity co- efficients is to take advantage of the fact that yi has been experimentally determined and tabulated (Lewis and Randall [1961]), whereas 6i has not. From Equations (2-12) and (2-13) it follows that (yioi)v/2V+, (2-1A) v/2v- (2_15) <<: ll (ye/5:) Thus knowledge of 5+ at a given ionic strength would yield knowledge of both y+ and y_. Following Goldberg and Frank [1972] it is assumed that 5+ depends on ionic strength according to in 6: = BIc . (2-16) For purposes of approximate analytical solutions B is assumed independent of ionic strength. For numerical 1A solutions B is given functional dependenciestx>match various activity coefficient theories. In a NaCl-H20 solution at an ionic strength of 1M, for example, a value of B of 0.5M-1 corresponds to y+ = 1.13A and y_ = 0.688; at B = 0, y+ = 0,833 and y_ = 0.833; at B = -0.5M’1, y = 0.688 and y_ = 1.13u. Thus, 4. at a given ionic strength, increasing B corresponds to increasing the "stability" of the positive ion relative to the negative ion. Clearly, knowledge of B would be very useful in determining the structure of an ionic solution in water on a molecular scale. E. Electrostatic Equations An additional relation between electric potential and concentration is provided by Poisson's equation (32¢/ax2> = -(Aw/e)D(X) , (2-17) where g is the permittivity of the medium and p(x) is the charge density, p(x) = F(Z+c + Z_c_) . (2-18) 4. Thus, (a2o/ax2) = -(AnF/e)(Z+C+ + Z_C_) . (2-19) 15 For computer simulation purposes (Chapter A), it is useful to replace the electric potential with the electric field E, E = -(ao/ax) , (2-20) and to replace Poisson's equation with the displacement current equation (Duckworth [1960]), H H) (aE/at) = (An/e)(l - z J+ — z J + (2-21) This conveniently introduces the current I as an inde- pendently adjustable variable. F. Combined System of Equations Combining Equations (2.1) - (2.13) yields Equations for the time derivatives of c+ and c_ in terms of more practical space derivatives than the continuity equations provide: (3/8x)[a++(3c+/3x) + a+_(3c_/3x) (3c+/3t) = + 8+E(3¢/3x) - c+vo] , (2-22) (ac_/3t) = (3/3x)[a_+(ac+/8x) + d__(ac_/3x) + B_E(3¢/3X) - C_vo] , (2-23) 16 where B+E = F(z+2++ + zl£+_) , (2-2u) B-E = F(z+i_+ + z_i__) , (2-25) and a++/RT = (2+,/c+) - %(z+/z_)(z+ - z_)[z+2+,(M+B) -z_2+_(M-B)]. a+_/RT = (£+_/c_) - %(zd/Z+Xz+-z_)[z+£++(M+B)-z_£+_(M-B)], a_+/RT = (z_+/c+) - %(z+-z_)[2+2_+(M+B)-z_2__(M-B)l, a__/RT = (£__/c_) - %(z_/z+)(z+-z_)[z+2_+(M+B)—z_2__(M—B)J. (2-26) The stoichiometric coefficients and ionic valenchxsare re- lated by v_ = z+[v/(z+ - z_)], v, = -z_[v/(z+ - z_)l (2-27) which are easily obtained from v+ + v_ = v, v+z+ + v_Z_ = 0 (2-28) 'The quantity [v/(z+ - z_)] is unity except for 2-2, 3—3, etc., electrolytes because in those cases v_ + 2+ and 17 v. + 12-1. Equations (2-19), (2-22), and (2-23) form a system of nonlinear partial differential equations. To the extent that vO can be neglected there are three dependent vari- ables, c+, c_, and o, and two independent variables, x and t. Of course, ¢ can only be determined to within an addi- tive constant. Because of the complexity of these equations, an exact solution for this system cannot be expected for any physically meaningful set of boundary conditions. Part two of each of Chapters 5 and 6 deals with simplifying ap- proximations that can be made for this system, after a suitable transformation of variables, in order to obtain analytical solutions. Chapter A contains a numerical tech- nique used in solving this system of equations. G. Transformation of Concentration Variables In the absence of chemical reactions, the isothermal transport of electrolytes can be thought of as the result of competition between an electric potential gradient and an activity gradient. These, of course, may oppose or reinforce each other. In order to separate the mass dif- fusion effects from the electrical migration effects, the following transformation is defined, 8 = %-[(c_/z+) - (c+/z_)] , (2-29) 18 % [(C_/z+) + (c+/z_)] . (2-30) U II S can thus be thought of as a sum or neutral electrolyte concentration, changing mainly due to diffusion. D can be thought of as a charge concentration, changing mainly due to electrical effects. In terms of these new variables, 0+ = z_(D-S), c_ = z+(D+S), (2-31) 0 = 2z+z_FD Equation (2-17) becomes, (32¢/3x2) = -KD , (2-32) where K = 8an+z_/e . (2-33) Besides the fact that Poisson's equation takes on a simpler form, this transformation has numerical advantages ‘when integrating the Poisson equation. Since the most important experimental measurement for this system is the puntential difference between two points, the calculation of 19 Ac from concentration profiles is very important. Because small changes in (c+ - c_) produce large changes in the po- tential difference, integrating Equation (2-19) can be risky. The use of Equation (2-32) eliminates this undesirable numerical effect. The differential equations for S and D are, from Equa- tions (2-22) and (2—23) (BS/8t) (ab/3t) where Yss (3/3x)[yss(aS/ax) + YSD(aD/3x) + YSEK-l(3¢/3x)-SVO], (2-3A) (a/afiErDS(BS/3x) + yDD(3D/3x) + YDEK'1(3¢/ax)-Dvol, (2-35) 1/2(a+++a__) - 1/2[(z+/z_)d+_ + [z_/z+)d_+], -1/2(a++-a__) - 1/2[(z+z_)a+_ - (z_/z+)a_+], -1/2(a++-a__) +1/2[(z+/z_)a+_ - (z_/z+)c_+], l/2(d+++a__) +l/2[(Z+/Z_)a+_ + (z_/z+)a_+], (2-36) (Ans/e)(z_e_E-z+e+E)=-(AnF2/e)(ziz++-zfz ), (“NF/8)(z_B_E+Z+B+E)=(unF2/€)(Zi£+++2z+z_£+_ + 2 Z_£__) = (An/e)x = ‘1 TD , 20 where A is the Specific conductance and TD is the dielectric relaxation time.' In Equations (2-36) and throughout this dissertation, 2+_ = 2_+. This is consistent with the use of the flux equations, (2-3) and (2-A). In terms of these same variables Equation (2-21), the displacement current equation, becomes (BE/3t) = (“w/E)I + KCYDS(BS/3X) + YDD(3D/3X) + -1 R YDEK (3¢/8X)]. (2_37) Thus the flux term in the displacement current equation is the same as the flux term in Equation (2-35). Equations (2-32), (2-3A), (2-35), and (2-37) form the set of general partial differential equations which are solved to various extents in Chapters 5 and 6. To complete the problem, a set of boundary conditions which can be used to model many types of processes is presented in Chapter 3. H. Discussion of Transport Coefficients for Transformed Equations The rather long expressions for YSS’ YSD’ YDS’ and YDD in terms of the 213, M, and B appear in Appendix A. For uni-univalent electrolytes in neutral solutions they, along with the expressions for YSE and YDE’reduce to: 21 RT YSS = §§E(£++ + 2£+_ + t__) + Ms (2++ + 22+_ + 2__) + BS (£++ —t__)] , (2-38) ' - BE[( ) + MS (2 A ) - BS (i - YDS - 2S 2-- - 2++ —- - ++ ++ 2£+_ + 2 )l, — BI YDD ‘ 2s (£++,+ 8--) . (2-39) _ BI YSD ' 28 (2-- 2++)’ =(uF2/m -£) YSE " 8 ++ —- ’ (2-A0) YDE = (“UF2/€)(2++ " 2£+_ + £__) The quantities YSS and YDD can be thought of as diffusion coefficients for "salt particles" and "charge particles" respectively. Thus, YSD and YDS are the cross diffusion co- efficients. The corresponding fluxes for these "particles" are S - YSS(3S/3x) + YSD(3D/3x) + YSEK_1(3¢/ax), (2-u1) I c_, I D ’ YDs(3S/3X) + YDD(3D/3x) + YDEK'l(a¢/ax) . (2—u2) I c_, I 22 These are not Onsager flux equations since YSD # YDS' They are quite satisfactory for describing experimental flows, however. The conventional Fick's law salt diffusion coefficient (Miller [1966]) is proportional to (788 - -1) YSBYDsYDE ' From the positive definite nature of the entropy pro- duction equation (Hasse [1969]) it can be shown that the quantity (£++ - 2% + 2__) is always positive, as are +- 1 and t__: From this it can be seen from Equations ++ (2-38) and (2-39) that 788 and YDD must then also be posi- tive. Thus salt and charge always tend to diffuse in a direction such as to eliminate the gradients of S and D, respectively. The signs of YSD and YDS depend on the rela- tive magnitudes of t++ and t__. Thus, for instance, a positive D gradient causes a flux of charge in the negative direction. This can be accomplished by negative ions dif- fusing in the negative direction and positive ions diffusing in the positive direction. If 2++ > t__, positive ions are diffusing faster than negative ions, YSD is negative, and there is a flux of salt in the direction of the D gradient. If 2_ < 2++, then YSD is positive, and salt and charge will diffuse in the same direction. Figures 2-1 through 2-A are plots of respectively, YSS’ YDD’ YSD’ and YDS versus the square root of ionic strength. The substances are HCl, LiCl, NaCl, and KCl. B has been set to zero for these comparisons. 23 Figure 2-1 shows that the highly mobile H+ ion increases the diffusion rate of HCl significantly and the highly solvated Li+ ion decreases the diffusion rate of LiCl. For KCl YSS is slightly larger than for NaCl, presumably because K+ is larger than Na+. The upturn in all of the YSS plots around 0.5M is due to the increase in their mean molar activity coefficients near this concentration. Figure 2-2 shows that with increasing ionic strength YDD becomes less than YSS' There are two reasons for this. For l-l electrolytes in neutral solutions there are no cross Onsager coefficients in the YDD expression as there are fer YSS' There are also no activity coefficient terms for these YDD expressions. Thus the upturn around li/z = 0.5Ml/2 is absent. These effects combine to produce the result that the ratio of charge to mass diffusion increases with increasing ionic strength. Figures 2-3 and 2-A show that YSD and YDS are generally smaller than YSS and YDD‘ This is because they contain a relatively large term which involves the cancellation of diagonal Onsager coefficients. The cross coefficients, YSD and YDS’ are negative only for HCl because it is the only electrolyte shown for which £++ > 2__. There is no upturn for YSD because it does not involve activity co- efficient terms. For KCl YSD and YDS fall significantly below the curves for the other electrolytes because of the almost complete cancellation of t++ and t__ 2A Figure 2.1. Dependence of YSS on ionic strength at B equals zero. (a) HCl, (b) LiCl, (c) NaCl, (d) KCl. 25 . H . N mazwfim Ectsmiyésm ad 06 3 9N SSA 17__()|, I‘SZwO / 26 Figure 2.2. Dependence of YDD on ionic strength at B equals zero. (a) HCl, (b) LiCl, (c) NaCl, (d) KCl. 27 m.m opswfim Azctemlyécm 9— 06 md 16 90 i i I _ n n l I! . n :13 // illlttlllllr, o Illllllltlllllllllltttr- o /h.. 3 ‘- db ‘- - q— d)— d- 00/. 9.0L |'_szulo / 28 Figure 2.3. Dependence of YSD on ionic strength at B equals zero. (a) HCl, (b) LiCl, (0) NaCl, (d) KCl. m.m onswfie Azctem2.o:tem 0.0 ad 3 29 c-(r- nu)— unh- d)- d)— du— 31 0.»: A a: nu . u / o N . 3 w 7v 8 _ o..- L o. n: 0.0 o.— 30 Figure 2.A. Dependence of YDS on ionic strength at B equals zero. (a) HCl, (b) L101, (0) NaCl, (d) KCl. :.m madman AzctemiyEcm Q6 06 *6 -r aL-g 1. fi— 31 db qr— ‘- .4- 0;: 90 Q- 80A. 9-0L l'_Szl.l..|0 / 32 Figures 2-5 and 2-6 show the effect of B on YSS and YDS respectively-for NaCl - H2O solutions. As B increases 788 decreases because 2__ > £++ for NaCl. When 2++ > t__ the Opposite is true. Qualitatively this means that if the more mobile ion experiences a significantly higher activity coefficient at lower concentrations, there will be less of a tendency for the electrolyte to diffuse in its own con- centration gradient. As B increases YDS decreases sharply with increasing ionic strength, not only for NaCl, but for all electrolytes. Because of the complex nature between the coupling of S and D there appears to be no simple ex- planation for this behavior of yDS. The coefficients, YSE and YDE’ as expressed in Equation (2-A0), are shown in Figures 2-7 and 2-8, respectively, for HCl, LiCl, NaCl, and KCl. The YDE/S curves are higher than the corresponding ySE/S curves because the YDE expression involves a sum of diagonal Onsager coefficients, whereas the YSE expression involves a difference. This reflects the fact that a charged particle will move faster in an electric field than a neutral particle. The complete, formal Onsager thermodynamic treatment of charged systems has not yet been accomplished. Presumably the Onsager flux equations corresponding to Equations (2-A1) and (2-A2) would contain coefficients similar to those discussed in this section. The phenomenological approach here is sufficient for the applications described in Chapters 5 and 6. 33 Figure 2.5. Dependence of YSS on ionic strength for NaCl. (a) B = -0.5 M‘l, (b) B = o, (c) B = 0.5 M'l. 3A m.m opswfim Azctcmiyéem ad Gd *6 od —— db u cub -- ‘- db N; H— w; 9— Q— fi— 88" 9-0L L‘SZLUO / 35 Figure 2.6. Dependence of YDS on ionic strength for NaCl. (a) B = -o.5 m‘l, (b) B = o, (c) B = 0.5 M‘l. 36 m.m osswfim Azctcmioztem 90 *6 «d 37 Figure 2.7. Dependence of YSE/S on ionic strength. (a) HCl, (b) LiCl, (C) NaCl, ((31) KC1. 38 e.m opswam Azctsmiyésm 0.0 *0 II 0.7- 0.— 38A) 0L0L L-'°wL—52w° / ‘5’ 39 Figure 2.8. Dependence of YDE/S on ionic strength. (a) HCl, (b) LiCl, (c) NaCl, ((1) KCl. A0 . w.m opswfim 133310.25 GO Q0 to _ . _ p n _ . _ . _ . _ I Q— ! 9N Q5 30/6) OLOL Lnow ,5ng / (S/ CHAPTER 3 ION TRANSPORT BOUNDARY CONDITIONS A. Introduction Any mathematical or physical description of a system is, at best, limited without a description of the events taking place at the system's boundaries. For the one- dimensional systems discussed in Chapters 5 and 6 there are two boundaries. At each boundary it is necessary to make a statement about the concentration and/or gradient of the concentration of each ion present in the system.. Since these boundaries correspond to the intersection of two phases, as shown in Figure 3-1, it is convenient to employ kinetic boundary conditions of the form, 1 = kifcio ‘ kibcié ’ (3'1) for the flux Ji of ion 1 through an interface of thickness 6 at each of the interfaces. The forward and reverse rate constants are kif and k1b (the directions are indicated in Figure 3-1), and Ci is the ion concentration. Boundary conditions like these have been used before by Chang and Jaffé [1952], Brumleve and Buck [1978], and have been A1 A2 P7 H H) W ‘F 27 I-‘° C‘ Phase L Phase R Interfacial Region Figure 3.1. Interfacial region at a phase boundary. A3 discussed by Perram [1980]. The purpose of this chapter is to determine the assumptions required for validity of boundary conditions of this type. This discussion fol- lows closely that of Horne, Leckey, and Perram [1981]. B. Derivation of Kinetic Boundary Conditions At equilibrium, the fluxes vanish and the Nernst Distribution Law (Prigogine and Defay [195A]) obtains for the ratio of activities ai in the absence of an electric field, _ _ _ e- aio/aio ' (Cicyio/cioyio) ‘ KiN ‘ eXp(‘AGi/RT) (3'2) where yi is the molarity based activity coefficient, K1N is the Nernst Distribution Constant, and AGG- the standard 1’ Gibbs free energy difference, is the difference in standard chemical potential across the interface, 8’ 11.3- 8 . (3-3) AGi g 6 - 1110 If an electrical potential difference, A¢ = ¢6 - c is O, maintained across the interface, then Equation (3-2) be- comes (aid/aio) = KiN exp(-ziF¢/RT) . (3-A) AA Equations (3-2) to (3-A) are readily derived from the equilibrium requirement that ”id = U10: with the electro- chemical potential given by .9. pi = “i + RT 2n a1 + ziFo . (3-5) If Ji = 0 in Equation (3-1), then (Cid/cio) = (kif/kib) ° (3'6) ThuS, by Equation (3-2), (kif/kib) = Kf for ideal solutions at equilibrium in the absence of an electric field and, by Equation (3-A), (kif/kib) = K exp(-ziFA¢/RT) for ideal iN solutions in the presence of an electric field. If cross terms are neglected, or if only one ion is moving through the interface, the flux of an ion in one dimension is given by -Ji = tii(3ui/3X) , (3-7) where iii is the Onsager coefficient and pi is the electro- chemical potential. The electrochemical potential gradient is, by Equation (3-5), ___= i i+ZFdl+——£ (3'8) A5 with d an yi Pi = 1 + ——————— . (3-9) d in C1 Although Equation (3-8) appears quite different from Equa- tions (2-7) and (2-8), the only real change is the inclu- sion of the general standard state term, (dug/dx) in Equa- tion (3-8). This is necessary because here the ions are passing from one phase to another, so the standard state must change . Substitution of Equation (3-8) into Equation (3-7) yields J = D dci + ziFDiCi Q9 + D ci 1 (3-10) 1 1 dx RT dx RT dx ’ where D = _ ’ (3‘11) and where activity coefficient terms have been neglected. Equation (3-10), rearranged, is a differential equa- tion for c1, ——+————=-——, (3-12) A6 with A1 = of + ztib . (3-13) If A1 is linear in x and if J1 and D1 are constants across the interface, then the solution of Equation (3-12) is xAA xAA _ i GRT 1 Ci - cioexp(- m) + J1 ———D1AAi[exp(- ——ORT) -1] , (3-1’4) with -e- = + .. - AAi Aci ziFdo . _ (3 15) The interfacial boundary condition, Equation (3-1), is obtained by putting x =(5in Equation (3-1A) and solving for J . This results in i D AA AA AA _ i i i i -l kif - _6RT [exp(- WJEl-exph RT” 3 D A (3-16) AA A _ i i i -l kin ‘ “(S-RT [1 ' exp(“ RT )1 The ratio of kif to kib is kif Au? + ziFAq) k1b - exp( RT ) , (3-17) “7 as required by Equations (3-A) and (3-6) for ideal solu- tions. If Z1 and Ac are positive and if ziFAo > Aug; then kif < kib' This is as expected since a positive ion would be moving into a region of higher electric potential as it moves in the positive x direction. Similarly, if Z1 is negative, Ac is positive, and ziFA¢ > Aug; then kif > kib' The other term in the exponent of Equation (3-17), Aug. can be approximated by the not very accurate Born equation, as Perram [1980] has done. z2F2 '0 i 1 1 An = (——---- , (3-18) i 231 SR eL where ai is the radius of the ion and ER and EL are the di- electric coefficients of the two phases. For Aug'> |ziF¢|, the Born equation predicts a larger rate constant corres- ponding to transfer to the higher dielectric phase. This is also as expected, since the higher dielectric will "stabilize" a charged particle more. Clearly the parameter 6, the thickness of the interface, is not readily accessible. It may be possible to estimate it from experiments which determine kif’ k and the distribution coefficient KiN ib for the ion between the two phases. Thus, Equation (3-1) is strictly valid if: (i) the neglect of cross terms in the flux equations is valid, A8 (ii) A? and ¢ are linear functions of x, and (iii) J1 and Di are constants across the interface. C. Applications of Kinetic Boundary Conditions One of the main features of Equation (3-1) is that it can be used for a wide variety of problems. It has been used in program IONFLO to increase the program's versatility. Without it, it would be necessary to rewrite large sections of code for each different set of boundary conditions. For liquid-liquid Junctions, discussed in Chapter 5 of this dissertation,the rate constants are set equal to zero because there is no flux of ions through the ends of the cell. For double layer relaxation studies, described in Chapter 6, the rate constants are set equal to zero at the blocking electrode since by definition no ions cross this boundary. At the reversible electrode the rate con- stants are set equal to infinity and the bathing concentra- tions are set equal to the equilibrium system concentrations. As mentioned in the introduction to this chapter, this type of boundary condition formulatiOn.has been used pre- viously by other workers. Chang and Jaffé [1952] used similar equations to study the effect of increasing rate constants at the electrode on the conductance. They did this for alternating current studies. Brumlene and Buck [1978] have used these boundary conditions to study num- erically a variety of effects. Among other systems, they “9 have studied charge steps at a blocked interface and mobile site ion exchange membranes. For the blocked interface simulation, they set the rate constants at the blocking electrode to zero, and observed the effect of current reversal on the ion distributions near the electrode. For ion exchange membranes they considered three ionic species, one trapped in the membrane by zero values for its transfer rate constants, the other species initially on either side of the membrane with finite rate constants. This simula- tion was used to determine selectivity properties of ion exchange membranes. Steele [1980] is using these boundary conditions to simulate the steady-state performance of various types of solid batteries. Perram [1980] has had considerable success in describing biological membrane transport with these boundary conditions. CHAPTER A NUMERICAL FORMULATION OF THE TRANSPORT EQUATIONS A. General Finite Difference Approach for Parabolic Partial Differential Equations The transport equations and boundary conditions des- cribed in Chapters 2 and 3, respectively, form a system of nonlinear parabolic partial differential equations. Such systems, in general, are not soluble by exact analytical methods. To get an accurate solution it is necessary to resort to numerical techniques, which, although approximate, can be used to get as accurate a solution as desired. The numerical technique of choice to solve the system of equa- tions of interest here is the finite difference method. The general presentation given in this section is based on Smith [1978] and Rosenberg [1969]. A review of numeri- cal simulations applied to ion transport is included in Section C of this chapter. To employ the finite difference method, continuous variables are replaced by discrete ones, and partial deriva- tives are replaced by finite differences. The resulting equations are algebraic rather than differential. To obtain discrete variables, the time-space solution domain 50 51 is replaced by a discrete grid as shown in Figure A-l. The solution is calculated only at the grid points. Any dependent variable U can be specified in time and space with its appropriate indices Ui,n A finite difference approximation to the first and second space derivatives of U at constant time is obtained and U in Taylor's series about the by expanding Ui+l,n i-l,n point Ui,n’ and taking the sum (for the second derivative) and difference (for the first derivative). If terms con- taining (Ax)3 and higher powers of Ax are omitted, the results are (3U/3x)i,n = (U - Ui_l,n)/Ax , (A-l) i+1,n (BU/8X)i,n = (Ui+l’n - 2Ui,n + Ui_1,n)/(AX)2 . (A-2) A finite difference approximation for the time deriva- tive of U can be made by combining Taylor's series expan- sions about any of the points +, o , or o as indicated in Figure A-l. The choice is mainly a matter of computa- tional efficiency. The use of point + (forward difference method) produces the simplest set of algebraic equations to solve, but suffers from a severe drawback. ’For this method to be stable the size of time steps, At, relative to the size of space steps, Ax, is very restricted. The use of point 0 (Crank-Nicholsen method) results in the most 52 1 IT+'I o o o o o -+ o o o .1 t f] O O o I (o C) O o o l1-I o o o o o o o o o o H l i+l )(-——-> Figure A.1. Uniform grid mesh for finite difference equations. 53 accurate time derivative approximation, but also produces the most complicated algebraic equations. A Taylor's series about point 0 (backward difference method) produces a time derivative analog that is unconditionally stable, but not as accurate or as cumbersome to use as the Crank- Nicholson method. After truncating terms containing (At)2 and higher powers of At (actually the (At)2 term vanishes for the Crank-Nicholson method), the time deriva- tive analogs become, Forward difference (aU/at)l,r1 = (Ul,n+l-Ui,n)/At ,' (4-3) Crank-Nicholson (BU/at)i,n+l/2=(Ui,n+l'Ui,n)/At’ (A-A) Backward difference (BU/at)i,n+l=(Ui,n+l-U1,n)/At' (14-5) Since these time and space derivative analogs can be applied at every interior grid point for each dependent variable, there will be a total of Jx(R-2) unknowns for each time row n, and Jx(R-2) angebraic equations. Here J is the number of dependent variables in the original partial differential equation system. Similar treatment of the boundary conditions produces a total of JxR equa- tions and JxR unknowns for each time row. If the original 5A partial differential equations are nonlinear, the system of algebraic equations will be so also. It then requires some sort of iteration procedure (Fletcher-Powell, Newton- Raphson, egg.) to solve them. The right hand sides of the equations in any case are formed from the solution to the system of equations at the previous time row. Starting from the initial conditions, it is then possible to step through time, calculating the solution to the partial dif- ferential equation system at each grid point, one row at a time. B. Difference Equations for Ion Transport The finite difference method described in Section A of this chapter can be applied to the system of electrolyte transport equations and boundary conditions detailed in Chapters 2 and 3 with several modifications. 1. Space and Time Grids It is well known that near charged interfaces quanti- ties such as concentration and electric field change rapidly over small space dimensions (<10"5 cm). To reduce trunca- tion error in the Taylor's series expansions, it is neces- sary to increase the number of space points in these regions. Since it is still necessary to transverse the entire solution domain With grid points, it is necessary 55 to use a space grid like that depicted in Figure A-2. This means that in regions where Axi and Axi-l are not equal, the finite difference space analogs are more com- plicated than previously stated. Equations (A-1) and (A-2) then'become, 2 _ 2 2 2 - (Axi) Ui=l,n}/[AxiAX1-1(AX1+AXl-l)J , (A-6) (82U/3X2) = [2Ax U -2(Ax +Ax )U i,n i-l i+l,n i 1-1 i,n + 2Aini_l,nJ/[AXiAxi_l(Axi+Axi_l)1 . (A-7) \ ‘o Although McAfee [1976] has theoretically develOped fairly general procedures for optimizing nonuniform space grids, a less cumbersome empirical approach is used here. Since the dependent variables change slower as steady- state is approached, it is useful to employ a nonuniform time grid to reduce the number of time steps necessary to reach a steady-state or equilibrium. A typical time grid of this sort is also shown in Figure A-2. Choosing the appropriate time derivative analog re- quires consideration of the relative computation times, complexity of the computations, and ease of finding and 56 n+loooooo o. o o 0000 t [1000... o o o 00.0 I n—[oooooo o o o 0000 I l-I l i+l Figure A.2. Nonuniform grid mesh for improved accuracy near boundaries and for more efficient ap- proach to steady-state equilibrium. 57 eliminating errors. It has been found empirically, that the backward difference analog performs best for this system of equations. 2. Application to Ion Transport Derivatives Equation (2-37) can be converted to finite difference form by a direct application of Equations (A-5) and (A-6). Equations (2-3A) and (2-35) contain derivatives of the form (a/ax)[P(U,V)W] and (a/ax)[P(U,V)aU/ax], where W is the electric field, U and V are the sum and difference concentrations, and P is a transport coefficient, 788’ YSD’ YDS’ YDD’ YSE’ or YDE‘ The discrete forms of these derivatives are best derived by repeated applications of Equation (A-6). They become, {(a/BX)[R(U,V)W]}i = P(6a2/YSS) (i.e., for t3105 s for a 1 cm cell), 78 to better than 1% accuracy, EJ(t) = -(u/w)K(vDS/YDE)(8§ - s§)exp<-R,lt> , (5-12) with R+1 E (w/a)2(YSS + You) = (n/a)2(a++ + c-_) . (5-13) (iii) For (5/YDE) (5/YDE), E is independent of B CELL 81 5 2 (iii) For (5/YDE) < t < (10 a /YSS), the cell po- tential can be expressed, after considerable algebra, in the particularly simple form ECELL=(-2RTt+/F)Itn(sg/s§) + tuna/mu . <5-20> where t+ is the transference number of the positive ion (see Appendix A). This is the classical electrochemical cell potential derived by Guggenheim [1929] assuming con- stant transference numbers across the cell. Table 5-1 shows a comparison between Equation (5-20) and some experimental results for concentration cells operating with liquid Junctions and moderate electrolyte concentration differences. The differences are quite small as long as the cell electrolyte concentration differences do not become too large. Although these cell potential comparisons are all in the constant cell potential time interval, they give some idea of the error induced by the linearization for longer and shorter times. The next section describes the use of program IONFLO to check the validity of the analytical solution for times less than (5/YDE). F. Numerical Solution In an effort to validate the analytical results at short times, program IONFLO was used to solve the system 82 Table 5-1. Comparison of Experimental Cell Potentials (EEXP) for Cell (A) with Cell Potentials (ECALC) From Equation (5-20). Cell sg/mol-dm‘3 sL/mol-dm'3 ECALC/10'3-v BEXP/lo'3-v KCl 0.10 0.040 21.16 21.17a KCl 0.10 3.00 -80.37 —80.A2a NaCl 0.10 0.080 _ A.13 - A.05b NaCl 0.10 0.040 -l7.09 —16.8ub HCl 0.10 0.078 9.9A 9.95a HCl 0.10 0.0u0 35.39 36.21? H01 0.10 0.005 103.u 118.8a aShedlovsky and MacInnes [1937]. bBrown and MacInnes [1936]. 83 of Equations (2.3A), (2.35), and (2.37) numerically. The V0 term was omitted from Equations (2.3A) and (2.35) for the reason discussed in Section C. For convenience the electric potential was replaced by the electric field in Equations (2.3A) and (2.35) according to Equation (2-17). The current I was set to zero since it vanishes for this experiment. Because the dependent variables S, D and E change rapidly near the liquid-liquid interface (X = 0) at short times, a non-uniform space mesh with a larger number of grid points near the interface was used. The 7 grid points were 10- a units apart near the interface and gradually expanded to lO'la units apart at the electrodes. Ninety grid points were used. Time steps of lO’lOs were used initially. The step size was increased as the cell approached equilibrium. Numerical stability was checked by comparison of results obtained using different step sizes. The numbers of spa- tial and temporal grid divisions were varied by factors of 3 and 10 respectively without change in result. For cell (A) operating at concentrations of SR = 0.11 mol-dm-3 and SL = 0.09 mol dm'3, non-linear computer simulation results and analytical solution results agreed to within 5% for all time ranges and all electrolytes. Since the only approximation in the computer simulation is neglect of the solvent velocity, 5% is a good estimate of the ac- curacy of the analytical solution for cell (A) operating 8A as indicated. Figure 5-1 shows the numerically predicted charge den- sity surface near the liquid-liquid interface for a NaCl concentration cell with an assumed value for B of zero (y+ = y_). Figure 5-2 shows the same cell at three dif- ferent times. As time increases the maximum value of the charge density decreases and moves away from the interface. Figure 5-3 is the same cell at 10-8s for three values of B. As can be seen, changing B changes the magnitude of the curve, but does not change its shape. This corresponds in the analytical solution to B appearing in the pre-ex- ponential factor but not in the time constants. 'Figure 5-A is a graph of cell potential versus time for cell (A) with M = Na, SR = 0.11 moi.dm‘3 and sL = 0.09 mol-dm-3. As a benchmark, the criterion that the cell no longer depends on B when the ECELL versus time 3 curves for B = 1 dm -mol-1 and B = -1 dm3-mol'l differ from the plateau region cell potential by less than 5% was chosen. With this somewhat arbitrary criterion, Figure 5-A indicates that from a time of 1.6 x 10793 on, the cell no longer depends on B. Figure 5-5 shows a graph of the coalescence time (tc) versus YDE for several electrolytes. Figure 5.1. 85 Predicted charge density surface near the liquid-liquid interface at short time for a NaCl concentration cell operating between concentrations of 0.11 M and 0.09 M. B is zero for this simulation (y+ = y_). The time row at t = 0 has been omitted since it would obscure the lower portion of the surface. 86 H.m ossmflm 87 Figure 5.2. Constant time planes from Figure 5.1. (a) 5 x 10'9s, (b) 1 x 10'8s, (c) 2 x 10‘8s. 88 m.m ensues . sob.) m o s N o m- T o- n- Figure 5.3. 89 Charge density YE cell position for different values of B at t = 1 x 10'8s for a NaCl con- centration cell operating between concentrations of 0.11 M and 0.09 M. (a) B = -1.00 M'1 (b) B = 0, (c) B = 1.00 m‘l. 3 90 m.m ossmfla sob.) o N- c. mwl mwt _ _ _ 91 Figure 5.A. Cell potential ys time for cell (A) with M = Na, SL = 0.11 M, and SR = 0.09 M. (a) B = 1.00 M'l, (b) B = 0, (c) B = -1.00 M‘l. 92 ON 0.. :.m oszmwm mmIO_\~ 0.. 0.0 _ n.mI m.m.. 93 Figure 5.5. Coalescence time ys YDE for 0.1 M solutions, with 5% criterion. (a) HCl, (b) BaCl2, (c) KCl, (d) NaCl, (e) LiCl. 9A l6- ' tc / I0"°s l l l I . 4 8 I 2 I 6 the / loss"| Figure 5.5 24 95 G. Conclusions The cell potential dependence on single-ion activity co- efficients vanishes very rapidly(t < 10785) for the systems investigated here. From Figure 5-5 it is clear that YDE is a good indicator of this effect for an arbitrary elec- trolyte. For the electrolytes studied, the relationship tC = (1'9/YDB) = (1.9 /AwA) = 1.9TD (5-21) approximates the coalescence time to within 5%. These results can be interpreted as follows. Initially in the interfacial region the electrical potential gradient is zero and the activity gradient is infinite. Consequently the positive and negative ions begin to diffuse from the region of high activity to the region of low activity. Since the aiJ are not equal, the diffusion begins at dif- ferent rates for the two types of ions. This soon creates a separation of charge and an electrical force that pulls the positive and negative ions back together. Thus a steady-state balance of diffusive and electrical forces is established. Before this balance is established the cell potential depends on the positive and negative ions separately. After this the forces and hence the ions them- selves are coupled in such a way as to remove their indi- vidual identities. Buck [1976] has similarly described the behavior of the Junction potential in terms of 96 differential ion motion. This model is consistent with the expression for yDE (the last of Equations (2_36)), 2 2 _ YDE = (HUF2/E)(Z+Z+++2Z+Z_£+_+z_2__) = (“fl/€)A = TD]. (5-22) Large values of t++ or £__ (HCl for instance) indicate rap— idly diffusing ions, large values of YDE and A, small TD’ and consequently a rapid vanishing of single-ion effects. The same is true for large values of 2+ and z_. For BaC12, the 2: factor offsets the relatively small values of 2++ and £__, which leads to relatively large values of YDE and A, small TD, and a small coalescence time. In terms of this model, the valence number of +2 for Ba++ creates a larger electrical force sooner. This apparently over- + to diffuse slowly, and thus whelms the tendency of Ba+ the coalescence time is relatively fast. For large values of e, YDE is small, TD is large, and the single-ion effect time is prolonged. This is consistent with the idea that a large value of 5 indicates that the positive and negative ions are relatively independent of one another. CHAPTER 6 DOUBLE LAYER RELAXATION AT BLOCKED ELECTRODES A. Introduction Double layer relaxation studies, both theoretical and experimental, have interested electrochemists to various degrees over the last 35 Years. Physicists and electrical engineers have also been interested in double layer relax- tion (also called space-charge decay), especially in semi- conductors and solid electrolytes. Laser and Bard [1976] note that for an intrinsic semiconductor the relaxation of free charge carriers resembles that of the cations and anions in the diffuse double layer at a metal/electrode interface following charge inJection to the metal. This is not surprising since the same thermodynamic and electro- static equations govern both processes. In an effort to study the kinetics of fast electrode processes, electrochemists have often been concerned with eliminating the double layer effects on electrode processes. Barker [1959] suggested the use of short duration current pulses as a means of eliminating the double layer charging current. Despite some experimental success by Reinmuth and Wilson [1962], Delahay and Aramata [1962], KooiJman 97 98 and Sluyters [1967], Weir and Enke [1967], Daum and Enke [1967], and Yarnitzky and Anson [1970], van Leeuwen [1978] has concluded that short duration pulse methods have not achieved the success expected of them because they in principle provide no more information than periodic tech- niques. Anson, Martin, and Yarnitzky [1969] have conducted one of the few experimental studies whose main goal was to study nonequilibrium double layers in the absence of electrode reactions. They studied the potential-time - behavior of several electrolyte solutions following coulosta- tic charge inJection into a hanging mercury drOp electrode. Their main conclusion was that the thickness of the double layer was not constant during rapid charge inJection, and that transient potentials calculated on the basis of a constant double layer thickness were in error. The research reported in this chapter is a theoretical thermodynamic study of systems similar to the ones investi- gated by Anson e£_al. Although the kinetics of fast elec- trode processes is not dealt with here, it is hOped that a better understanding of double layers in the absence of electrode reactions will lead to new ways to eliminate and/ or correct for them during electrode kinetic studies. More- over, the equations and their solutions are intrinsically of interest to both the macrOSCOpic and the molecular theorist. In an attempt to analyze the response to a coulombic 99 impulse applied to an electrode at equilibrium, Reinmuth [1962], starting from absolute rate theory and Fick's law, derived potential-time relaxation response expressions. These expressions were intended to cover the range from charge-transfer dominated processes to completely blocked electrodes. For Reinmuth's results to be valid, however, the solution must contain a large excess of supporting electrolyte. As Francescetti and MacDonald [1970] and Anson, Martin, and Yarnitzky [1969] have pointed out, there are many electrochemical and semiconductor systems where the use of supporting electrolytes is not feasible. In this chapter relaxation of the double layer at com- pletely blocked electrodes is considered in the context of the theory presented in Chapters 2 and 3. In an approach somewhat similar to that used in Chapter 5, both analytical and numerical results are presented. Section B deals with the equilibrium double layer in the presence of a charge on the electrode. The solution for the equilibrium case is in terms of a perturbation series with all activity coefficient effects included. Section C details an approxi- mate time dependent analytical solution using Laplace trans- forms. Section D deals with numerical simulation results using program IONFLO. 100 B. Equilibrium Solution The equilibrium double layer has been studied in great detail by chemists since the historic treatment of the subJect by Gouy [1910] and Chapman [1913]. Grahame [19A7], Delahay [1965], and Payne [1972] have presented extensive summaries of the topic. The only thermodynamic treatment which ineludes single-ion activity coefficients has been attempted by Hall [1977]. However, he presents no calculations. This section contains a systematic perturba- tion solution to the double layer problem. All single-ion effects are retained. At equilibrium the fluxes of all ions vanish, and therefore JS and JD, defined by Equations (2-A1) and (2-A2),- do also. This leads to the system of equations, 3E J = 0, J = 0, SE = KD , (6-1) for the ion and field distributions in the system in the presence of a charge on the electrode. Due to the presence of activity coefficients, Equations (6-1) cannot be solved exactly. Instead, a perturbation approach is used. In terms of S, D, and E, Equations (6-1) become YSS(dS/dx) + YSD(dD/dx) - YSEK‘lE = 0 -a/2:X:a/2a (6-2) 101 YDS(dS/dx) + yDD(dD/dx) - YDEK-lE = 0 -a/2ix5a/2, (5-3) dE/dx = KD -a/2:x:a/2. (6-A) Immediate simplification results from simultaneously eliminating the dD/dx term from Equation (6-2) and the dS/dx term from Equation (6-3). The results are dS/dx + (NS/H)E = 0 , (6—5) dD/dx + (ND/H)E = 0 , (6-6) where Ns = YSEYDS ' YDEYSS’ N0 = YDEYSD ‘ YSBYDD ’ (6-7) and H = K where the derivatives in the Taylor's series are AO = 1%5- 1% = l A: = -m(A:)2 A: = 0 A38 = 2m2(A:)3 128 = 0 _ (6-23) 82d = -2mb(>.:)3 Aid = -mb(A:)2 A: = b(l:)2 A3 = bl: Agd = 2b2(A:)3 )gd = [-2m(m+1)+2b2](A:)2 Substituting the expansions of s and d, Equations (6-20), into the coefficient expansions, Equations (6-21) and (6-22), and grouping according to powers of 6 yields 106 = S S A A0 + 6{Assl + A: d1} + 6 2{AS s32 + Add2 + 1 s 2 1A8 2‘sssl I AsdS 1dl I 2 dddl } I 5 ”{A sS3I Add3 I (6-2A) s s l s 3 + + Asssls2 I Isd(sld2 d2 5 l) I Idddld2 6Issssl I 1 s 3 l s 2 M 2 2 A 6Iddddl I 2Issdsldl I 2Isddsldl} I 0(5 )’ d _ d d d A - AO + 5(ASsl + Addl } + 6 ”(A Ss2 + Add2 + 1 d 2 d 1 d 2 3 d 2*sssl I Asdsldi I 2 dddl} I 5 {Ad s33 I Add3 I (6-25) d d d 1 d 3 ' Asssls2 + Asd(sld2 + d2sl) + Addd1d2 + EASSSSI + ExgdddI I 2Igsd81dl I Bxgddsl I u The perturbation approach outlined here is similar to that used by Horne and Anderson [1970] for thermal dif- fusion with one notable exception. The convergence of their solution relied partially on the smallness of higher order derivatives of their transport coefficients. Here no such smallness is expected. Instead the convergence properties of the solution rely on the magnitude of the charge on the electrode, q. If q becomes large enough it is probable that the solution, Equation (6-20),will diverge. 107 Application of Equations (6-20), (6-2“), and (6—25) to the system of Equations (6-14) - (6-18) and subsequent ordering of the terms according to the power of 5 yields the following system: dsn/dg + fn = o - 1/2 5 g g 1/2 , (6—26) d2dn/dE2-k2dn+dhn/d€ = 0 - 1/2 1 g g 1/2 , (6-27) den/d5 + k2dn = O - 1/2 3 g E 1/2 , (6-28) sn = 0, an = o g = 1/2 , . (6—29) en = -qn, ddn/dg = gn - hn g = -1/2 , (6-30) n: 1,2,3’°°' 9 where a second derivative of Equation (6-15) has been taken to permit substitution of Equation (6-16) into Equation (6-15). The fn, gn, and hn are defined by fl = 0 r2 = Agdlel (6-31) _ s s s 2 f3 ‘ Ao(d191 I d2el) I Assldlei I Addlel ’ 108 81 ' q gn = O n = 2,3,... , (6-32) h1 = O h2 = (A: + Ag)slel + Agdlel h3 - (Ag+kg)82el + (AS+A:)sle2+kg(dle2+d2el) + (6-33) s.d.s + sie1+%xfiddiel Higher orders of h and f are omitted for brevity. This system of equations can now be solved if equations of order n are solved before equations of order n+1, and if for a given n Equation (6-27) is solved before Equation (6-28). As in any perturbation method (Bellman [l96u]), there is no guarantee that Equations (6-20) will converge for 6 = 1. In this case convergence is not expected to be rapid for large charges on the electrode. Since this same system is solved numerically in the next section, the per- turbation solution is presented here only through third order. Through second order s, and first order d and e, the solutions are s 2 IO q 1 h2[k(€ 1/2)J (6 3”) S‘I'S= SI'l " a - l 2 2K2 cosh2(k) _ 2g d1 - k cosh(k) sinh [k(€-l/2)] , (6-35) 109 el = 33§%TE7 cosh[k(£-l/2)] . (6-36) At 298K for a 0.1 M bulk concentration and cell length of 1 cm, k is about 1x107. This means that, for instance, -exp(k) is an extremely good approximation to 2sinh (k). With this type of approximation for the hyperbolic func- tions, the results through third order are s' = l + (qu2/2k2)exp[-2k(£+l/2)J + . (6-37) (qu3/3k3){exp(-2k(€+l/2)]-2 expE-3k(€+l/2)]}, d' = -(q/k)exp[-k(g+1/2)] + (qu2/3k2) x {2exp[—2k(£+l/2)J-eXPE-k(€+l/2)J}- (q3/k3){(3/16)[A:+2(A§)2+Agd]exp[-3k(g+1/2)J (6-38) _ %(Ag)2exp[-2k(a+1/2)J-(1/16)EA§-%#{exp[-2k(a+1/2)J -exp[-k(€+l/2)]} - (q3/k2){<1/16)£A:+2(A§)2 + (6-39) Ad ]exp[-3k(g+l/2)] — 3(Ad)2ex [-2k(g+l/2)] - dd 9 d p (1/16)[A: - §§2 + AgdlexpE-k(a+1/2>J} , where the prime indicates the solutions are valid for k >> The identity A (6-40) Om CLO. has been used to simplify the 8' equation. The convergence properties of the s' and d' series are best evaluated at g = -l/2. At a = -l/2, the series for s' and d' become 3' 1 + (qu2/2k2) - (ASqu3/3k3) , (6-A1) d! -(q/k) + (qu2/3k2) - (q3/8k3)(l - f%Ag) , (6-u2) where the identity d 2 111 I has been used in the d' equation. Since an additional power of (q/k) appears in each progressive term of the series for s' and d', it is a good measure of the rate of convergence of the series. In terms of dimensioned variables (q/k) is (q/k) = (2'ITQ‘2/6RTSB)I/2 . (6-uu) Thus, small charges on the electrode and high concentra- tions favor rapid convergence. Small charges on the elec- trode indicate the system is not perturbed far from its original state, so the solution is well represented by the first two or three terms in the series. High concentra- tions indicate that for a given charge on the electrode the relative change in concentration at a given point in solu- tion is less than for low concentrations. If, for example, Q is 10-6 C/cm2 and SB is 0.1 M, then (q/k) is 0.29, and s and d are accurately represented by s' and d'. If, on the other hand, Q is 10"5 C/cm2 and S is 10-3 M, then B (q/k) is 291, and the series are apparently diverging. Figures (6-1) and (6-2) are plots of (s'-l) and d', as defined by Equations (6-37) and (6-38), versus distance from the electrode, showing the relative contributions of the terms in the series. The graphs are for a 0.1 M 6 NaCl solution with an electrode charge of 2x10- C/cm2, and b assumed to be —O.l. This corresponds to an activity , 112 Figure 6.1. Contributions to s' from 32 and s3 for a 0.1 M NaCl solution with an electrode charge of 2 x 10.6 C/cm2. b = -O.l; (A) s', (B) s2, (C) 53. 113 H.@ mpswfim ..o_>~\.-3 0N m. 0. n 0 d A d KWO.OI 00.0 000 9.0 nNd mmd 9.0 no.0 -mmd 13) ssamun/ (I 11A Figure 6.2. Contributions to d' from d1, d2, and d3 for a 0.1 M Nagl solution with an electrode charge of 2 x 10‘ C/cm2. b = -o.1; (A) d', (B) d1, (C) d2, (D) d3. m.w mpzuflh scion». ov mm 0» mm 8. m. o. n o. . a q a _ u _ . 115 ssamun/p 116 coefficient for Na+ of 0.71 and for C1' of 0.86. Two fea— tures of the graphs are readily apparent: (1) Even though (q/k) is greater than one (1.08) for these conditions, both series appear to be converging due to the numerical constants in the denominators of the perturbation terms. (2) The contributions from the terms containing Ag and Igd’ which depend on single-ion activity coefficients, are very small. Since the values of s' and d' decrease so rapidly as distance from the electrode increases, a more practical experimental variable is the potential difference. The potential difference in reduced form due to the equilibrium double layer is calculated from the electric field using Poissons equation in integral form, 1/ Aw = fe'dfi, (6415) -l/2 where Aw = (F/RT)A¢ , (6-46) and A¢ is the potential difference due to the double layer. The usual negative sign in Equation (6-A5) has been omitted to conform with standard electrochemical convention. Through third order the reduced potential difference is 117 Aw = awl + Aw2 + aw3 , (6-u7) where Awl = -q/k , (6-u8) Aw2 = -qu2/6k2 , (6-49) aw3 = q3EA: + Agd + 4(Ag)21/2Ak3 . (6-50) For the electrolyte solution whose d' profile is shown in Figure (6-2), the corresponding potential difference calculated using Equations (6-H5) - (6-A9) is -27.06 mV. I If b is assumed to be +0.1 (This corresponds to switching the activity coefficients from yNaI = .71 and yCl‘ = .86 sto yNaI a .86 and yCl’ = .71.) instead of -0.1, the cal- culated potential difference is -27.1u mV. Thus, accord- ing to these results the potential difference is the same to 3 significant figures for this solution. The potential difference seems insensitive to whether yNaI > yCl‘ or yCl‘ > yNaI' Although Hall [1978] has proposed that single-ion ac- tivity coefficients can be estimated from equilibrium measurements on electrolyte solutions at charged electrodes, when the electrode charge is small and specific adsorption is absent, it is difficult to confirm this from these 118 results. Exceedingly large or small single-ion activity coefficients must be postulated for the terms containing them to be significant. With Equation (6-A7) it is now possible to compare the results presented here with those of Gouy [1910] and Chap- man [1913]. They solved the Poisson-Boltzmann equation (Philip and Woodling [1970], Chen [1971], Olivares and McQuarrie [1975]) for the boundary conditions used in this chapter. Although their results only partially correspond with experrmental observation, they provide a logical starting point for comparison. In terms of the reduced variables defined by Equation (6-13), they predict the following charge-potential relationship q = —2 k sinh(A¢/2) . (6-51) Assuming for the present comparison that A: = l, Agd = 0, and A3 = 0 (i.e., m = b = 0, which is valid at infinite dilution), Equation (6—A7) becomes Aw = -(q/k) + (q3/2uk3) . (6—52) In order to compare this equation with the Gouy-Chap- man result, Equation (6-51) is expanded in a Taylor's series about Aw = 0. (This series always converges.) The result is 119 l 5 1920(AW ) + ...] 0 (6.53) q = -kEAw + §fi(aw)3 + Inversion yields Aw: -(q/k) +' (q3/2uk3) - (9q5/1920k5) + 0(q7) . (6-5A) Thus Equation (6-52) is identical to Equation (6-5H) through the second term in the series. There is little doubt that neglecting activity coefficients in the perturba- tion expansion produces a result completely equivalent to the Gouy-Chapman result. Another way of saying this is that the thermodynamic solution derived in this chapter is equivalent to the Gouy-Chapman result at infinite dilution. It is not surprising that the Gouy-Chapman result does not predict experimental results well at high concentrations and large electrode charges. At high concentrations A: begins to deviate significantly from one, and for large electrode charges higher order terms in the perturbation series, which almost certainly contain stronger activity coefficient dependence, become important. Although the failure of the Gouy-Chapman theory is generally attributed to its neglect of ions adsorbed on the electrode, it seems likely that part of its failure is due to the neglect of activity coefficients. 120 C. Time Dependent Solution In an effort to obtain the form of the analytical solu- tionfku‘double layer relaxation, the partial differential equation for D, Equation (2-35) has been Solved after assum- ing that: (l) v is very small; (2) YDS << YDD3 (3) YDD O and YDE are constant. Neglect of the v0 term is valid because there is no appreciable bulk flow during the relaxa- tion. The error introduced by neglecting the YDS term compared to the YDD term varies from system to system. As can be seen from Figures 2.2 and 2.A, YDS is about 1% of YDD for KCl and about 60% of YDD for HCl. Although for a given electrolyte YDD varies only about 10% up to 1 M, YDE is much more concentration dependent. Figure 2-6 shows it to be nearly linear in ionic strength up to l H. Al- though assuming YDE to be constant is a severe approximation, the results derived here will be compared to more exact computer simulation results in the next section. With these approximations, Equation (2-35) becomes (ED/3t) = YDD(82D/3x2) - YDED - a/2 : x g a/2 (6-55) Since there is no flux at the blocking electrode JD is zero there. Applying the above discussed approximations to Equation (2-42) at zero flux yields the boundary condi- tion at x = -a/2, 121 (ED/3X) = -(YDE/2FYDD)Q(’G) X = '3/2 , (6‘56) where Poisson's equation has been used to put the boundary condition in terms of the charge on the electrode, Q. At x = a/2, the reversible electrode location, the solution has its bulk prOperties, so the boundary condi- tion there is D = 0 x = a/2 . (6-57) Initially the solution is not charged, so the initial condition is D(x,t) = 0 -a/2 i x i a/2 . (6-58) Using a Laplace transform technique, as described in Ap- pendix G, the solution of this system of equations is D( t) (/ 1(2) f ”BRUNEI x, = k aFw n 2 2 n=0 <-1> dj‘ Q(t-U /YDE){exp[-U -1/2, <6-63) 123 and the complementary error function, erfc, is defined by 1/2 w -v2 erfC(Z) = (2/1r ) f e dv . (6-614) z For Equation (6-8) to be valid a must be less than YDE' In practice, this does not cause a problem since a is rarely larger than 107 s'1 and YDE is greater than 107 for most systems down to concentrations as low as 5 x 10'“ M. If a is greater than YDE the solution to Equation (6-59) takes on a different form, but nothing dramatic happens physically. It merely means that the charge has been applied to the electrode so rapidly that the ions have not had time to move far, until the electrode has been completely charged. In the other extreme, if the capacitor slowly charges the electrode, then a is small, w approaches one, and the solu- tion moves through a series of equilibrium double layers. Although this solution appears a bit cumbersome at first, it has several advantages over the Fourier series solutions derived in Chapter 5. Principally, it converges very rapidly for small times. It also takes on simpler forms for long times and for regions near the blocking electrode. As t goes to infinity, the exp(-at) portion of the equation goes to zero, and the remaining portion of the equation can be summed. The result is Equation (6-35), the first order solution for d. At x = -a/2 and all values of t, the solution for k >> 1 (k is much greater than one for all 12“ systems except possibly membranes.) reduces to D(-a/2,t) = (kQO/2aF)[erf(T) - w-lexp(-at)erf(wr)] , (6-65) where the error function, erf, is defined by z 2 erf = (2/nl/2) f e‘V dv . (6-66) 0 Figure 6-3 shows the dependence on a of D,at x = -a/2 as a function of time predicted by Equation (6-65) for a particular NaCl solution. As the electrode charging time constant increases, D at x = -a/2 reaches its equilibrium value much faster, as long as a is less than YDE' The difference in shape between the error function and expon- ential function is also evident from Figure 6-3. Figure 6-U shows the concentration dependence of the relaxation of D at -a/2 for an electrode charging time constant of 107 8‘1. Lower concentrations produce longer relaxation times and smaller charge densities at the elec- trode. Physically, the ions must travel further to reach the electrode for dilute solutions, so the relaxation time is longer. For regions near the electrode (x near -a/2) and k >> 1 it is possible to approximate Equation (6-62) accurately by the following expression Figure 6.3. 125 Dependence on a of D at x = -a/2 as a func- tion of time as predicted by Equation (6-65) 2 for a particular NaCl solution. SB = 10- M, Q0 = 10'6 C/cm2, (A) a = 3 x 107 s'I, (B) a = 1 x 107 s‘l, (C) d = 6 x 106 s'l. ‘1') m . m cLZHrTm s.0..0\« 0.0 0.? 0.? 00 .00 0M 0.N 0.. 0.. 0.0 0.0 0N 0 Figure 6.4. 127 Concentration dependence of the relaxation of D at x = -a/2 for an electrode charging time constant of 107 s"1 as predicted by Equation (6-65). (A) sB = u x 10‘2 M, (B) sB = 2 x 10‘2 M, (c) s = 1 x 10'2 M. B \ z.m mpzwfih sbro: 0.0 0...? 0...? 0..m 0.0 0..N 0..N 0“. 0... 0.20 000.0 10.0 .. 0.. mm W. a» 8 0 M m 4 ..ON WI u... .mN 0. g .00 L . :0 r0 129 D(z,t) = -(kQO/AaF){exp(kz/a)erfc[(kz/2ar) + TJ-exp(-kz/a)erfc[(kz/2aT)-T] - (6-67) w-Iexp(-at)[exp(sz/a)erfc[(kZ/2at) + wt]-exp(-kmz/a)erfc[(kz/2aT)-wr]]} , where z is the distance from the electrode and is related to x by z x + a/2 . (6-68) Figure (6-5) is a graph of D versus z/a as predicted by Equation (6-67) for three different times. For this com- parison the charge on the electrode at equilibrium is--10'7 C/cm2, a is 6 x 106 s-I, and YDE is 6 x 107 8. At 10.7 s the system is essentially at equilibrium. The potential difference as a function of time due to the double layer is calculated from Equation (6-67) by integrating Equation (2-32), Poisson's equation. The result is O A¢(t) = (aQ /ek){l-exp(-YDEt) + -;BE:;— (6-69) [eXp(-dt)-exp(-YDEt)]} , 130 Figure 6.5. Effect of time on D versus z/a profiles as predicted by Equation (6-67). QO = -10’7 C/cm2, a = 6 x 106 3'1, = 6 x 1078, (A) E 10-7s, (B) 10-85, (C) lO-Bs. 131 m.© opzwfim obSoE 0. 0. V m #0: 33.1113 - IOU-V0 0. I0 Q (0 132 where again a must be less than YDE' As t approaches infinity Equation (6-69) becomes A¢ = (aQO/kE) . (6-70) This is equivalent to Equation (6-48), the first order equilibrium potential difference. If a << YDE’ then Equation (6-69) becomes A¢(t) = (aQO/ek)£l-exp(-dt)] , (6—71) 1 E' ionic distribution. The change in potential difference with for t >> yB This situation indicates a virtual equilibrium time is due entirely to the change in the charge on the electrode. A more useful situation is when ak >> YDE’ This is true for most systems,except for membranes, extremely di- lute solutions, or slowly charged electrodes. For ak >> YDE’ Equation (6-69) becomes A¢(t) = (aQO/e){k-I[l-8Xp(-YDEt)1 + (6-72) a(YDE-a)-I[exp(-at)-eXp(-YDEt)1} In the next section this equation is compared to some computer and experimental results. 133 D. Computer Simulation Results and Discussion Program IONFLQ has been used to simulate the same type of systems as discussed in Sections A and B of this chapter. The goals in the simulations have been to: (1) delineate the range of applicability of the analytical results of this chapter, (2) determine the effects of single-ion activity coefficients on double layer relaxations, (3) compare the theory presented here with experimental results. Figure 6-6 is a charge density difference plot at 10"6 3 generated from two 0.1 M NaCl double layer relaxation simulations. The electrode charging time constant was 6 x lO6 3.1 and the charge on the electrode was -10'7 C/cm2. The only difference in the simulations was the 1 value of B. For B = -l M- (YCI‘ > YNaI) the charge density is slightly greater near the electrode than it is 1 simulation. Beyond about 5 x 10'8 1 for the B = 1 M- cm the charge density for the B = l M- simulation becomes larger. Physically this can be understood in reference to the activity coefficient curves for NaCl shown in Figure 6—7. Since the charge on the electrode is negative, the Na+ ions will tend to increase their concentration at the electrode relative to the bulk. If B is negative, then, from Figure 6-7, the relative change in the activity co- efficient for Na+ between the region at the electrode and the bulk is greater than for positive B. (i.e., The slope of curve 2 is steeper than curve 1 at 0.1 M.) Thus, there Figure 6.6. 13M Computer simulation of charge density dif- ference versus distance from electrode at 10‘65 for two 0.1 M NaCl solutions. a = 6 x 106 s‘1 1 B = -1 M- and p* corresponds to B = 1 M"1 , Q0 = -10-7 C/cm2 0 corresponds to 135 00 w.m ohswfim to. . Eo\x . o... 0% Qw. o“. 80... AS 0.0 - o._ . on - on 9. 136 Figure 6.7. Effect of B on Na+ and C1- activity coef- _ -1 ficients. (1) YNaI if B - l M and YCl’ 1 if B = -1 M- , (2) Y, for NaCl as measured by Brown and MacInnes [1936], (3) y01_ if 1 -1 B = 1 M’ and YNaI if B = -1 M 137 0N0 0&0 0...0 1.0 N..10 m.m omzmfim 2qu _.o 00.0 moo v0.0 «0.0 00.0 . l 00 0.0 CD 0 swmamaoo Kumov 0.. 138 is less of a tendency for Na+ to diffuse to the bulk, where it must increaSe its activity coefficient more than it would for negative B. In simpler terms the more positive B is, then the more Na+ enjoys the environment at the negative electrode. The opposite is true for Cl‘. Despite the fact that the charge density profiles are different for these two simulations, the potential difference across the double layer is the same for both of them to 5 significant figures. This result appears to be general for all electrolytes at all times during the relaxation process. Although different values of B produce different charge density distributions, the potential difference varies by ‘a virtually insignificant amount. Several simulations of relaxations due to extremely rapid electrode charging were made for 0.1 M solutions of NaCl, KCl, and HCl solutions. The time constant for the electrode charging process was set at 1011 5'1, 00 was 10"7 C/cm2, and B was varied from 1 M21 to -1 M-1. None of these showed any potential difference dependence on B, even for very short times (<10-9 s). This is quite different from the liquid junction potential behavior at short times dis- cussed in Chapter 5. The differential ion motion is ap- parently absent here. Several computer simulations have been made in an effort to assess the limitations of Equation (6-72). Tables 6.1—6.3 contain potential differences from selected 139 Table 6.1. Comparison of Equation (6—72) and Computer Simulation for 0.1 M KCl. A¢* Igdicates Equa- tion (6-72) Results. a = 6 x 10 s‘1 Q° = 10-7 C/cm2. a = 1 cm. t/S'10-7 A¢/V A¢*/V A¢/A¢* 0.125 ”3.9 “3.3 1.02 1.25 22.9 22.0 1.0“ 2.50 10.8 10.“ 1.0“ 5.00 2.42 2.32 1.0“ 7.50 0.54 0.52 1.0“ 10.0 0.122 0.117 1.0” -2I -2 + 1A.l 1.32x10 1.13x10 1.16 -3I -3 + 18.3 2.AAx10 2.18x10 1.12 25.1 1.u2x10’3 1.u0x1o'3 1.01 36.2 1.39x10"3 1.u0x10'3 0.99 +Time steps too large for the computer simulations at these time rows. 140 Table 6.2. Comparison of Equation (6-72) and Computer Simulation for 0.1 M NaCl. 00* Indicates Equation (6-72) Results. 0 = 6 x 1063-1. 0° = 10"7 C/cm2. a = 1 cm. t/s-10’ A0 /v A¢*/v A¢/A¢* 0.125 51.8 51.6 1.00 0.625 40.2 38.2 1.05 2.50 13.0 12.“ 1.05 5.00 2.91 2.76 1.05 7.50 0.647 0.617 1.05 10.0 0.146 .139 1.05 _2+ -2 + 12.7 3.2lx10 2.72X10 1.18 -3I -3 + 16.0 6.18x10 5.15x10 1.20 21.3 1.63::10‘3 1.54.:10’3 1.06 30.0 1.40;:10"3 1.391(10‘3 1.01 +Time steps too large for the computer simulations at these time rows. 1A1 Table 6.3. Comparison of Equation (6-72) and Computer Simulation for 0.1 M HCl. A¢* Indicates Equa- tion (6-72) results. a = 6x106 s-I. Q0 = 10'7 C/cm2. a = 1 cm. I t/s-10'7 A¢/v A¢*/v A¢/A¢* 0.1 ' 1A.5 16.A 0.88 0.5 11.6 11.u _ 1.02 1.0 8.60 8.97 1.02 2.0 u.72 4.65 1.02 u.0 1.u2 1.u0 1.01 7.5 0.176 0.172 1.02 10.0 4.021(10"2 3.83x10‘2 1.05 14.8 3.921(10‘3Jr 3.5ux10‘3 1.10I 19.0 1.61::10'3 1.56::10‘3 1.03 26.0 1.39x10”3 1.391(10"3 , 1.00 1. Time steps too large for the computer simulations at these time rows. 142 time_rows from these simulations for 0.1 M solutions of KCl, NaCl, and HCl, along with the corresponding results from Equation (6-72). There are relatively large deviations between the analytical and computer results for KCl and NaCl after 10-6 s and before equilibrium because the time step sizes were increased too rapidly during that part of the simulation. It is interesting to note that the devia- tions are about the same for all 3 electrolytes. This is somewhat surprising since one of the approximations which lead to Equation (6-72) was that YDS << YDD' At 0.1 M the ratios of YDS/YDD for HCl, NaCl, and KCl are -0.59, '0.21, and 0.018, respectively. Closer examination of the simulations, however, reveals that d2s/dx2, the term which multiplies YDS in Equation (2-35) is much smaller than d2D/dx2 for all three electrolytes. Presumably at higher electrode charges this would not be so, and the effect of omitting the st term would be more pronounced. These same simulations have been repeated at 10'3 M for the same three electrolytes. The results for selected time rows are listed in Tables 6.4 - 6.6, along with the results computed from Equation (6-72). Again the computer results are somewhat too high after 10-6 s and before equi- librium because the time steps have been set too large there. The relative deviations between the analytical and computer derived potential differences is slightly larger here than at 0.1 M. This is because the relative change 143 Table 6.4. Comparison of Equation (6-72) and Computer Simulation for 10"3 M KCl. A¢* Indicates Equa- tion (6.72) Results. 0 = 6x106 s'l. Q0 = 10'7 C/cm2. a = 1 cm. t/s-lO"7 A¢/v A¢*/v A¢/A¢* 0.250 1u19 1555 0.91 1.25 2268 2288 0.99 2.50 1293 1241 1.00 5.00 299 283 1.06 7.50 66.9 63.1 1.06 10.0 15.0 1u.1 1.06 1u.1 1.111+ 1.20 1.17+ 18.3 0.11Mr 0.107 1.35I 25.1 1.671110"2 1.63x10'2 1.02 36.2 . 1.1401(10'2 1.39'x10"2 1.01 fTime steps too large for the computer simulations for these time rows. 144 Table 6.5. Comparison of Equation (6-72) and Computer Simulation for 10"3 M NaCl. A¢* Indicates Equation (6.72) Results. 0 = 6x106 3‘1. 00 = 10"7 C/cm2. a = 1 cm. t/s-10“7 A¢/v A¢*/v Ad/A¢* 0.625 2502 2617 0.96 1.25 2628 2643 0.99 2.50 1594 1524 1.05 5.00 384 357 1.08 7.50 85.9 79.7 1.08 10.0 19.2 17.8 1.08 1u.1 1.82I 1.5a 1.18I 18.3 0.18144r 0.136 1.35+ 25.1 1.77x10‘2 1.59110‘2 1.10 3.62 1.40x10-2 1.39x10‘2 1.01 +Time steps too large for the‘computer simulations at these time rows. 145 Table 6.6. Comparison of Equation (6-72) and Computer Simulation for 10"3 M HCl. A¢* Indicates Equation (6-72) Results. 0 = 6x106 3'1. 00 = 10;7 C/cm2. a = 1 cm. t/s-10"7 Ac/v A¢*/v A0/A¢* 0.500 1052 1096' 0.96 1.00 875 864 1.01 1.50 657 6A3 1.02 3.00 268 261 1.03 5.00 80.7 78.7 1.03 7.50 18.0 17.6 1.03 10.0 4.03 3.93 1.03 1A.8 0.276I 0.23u 1.18I 19.0 3.70x10‘ 3.16x10’2 1.17I 30.9 1.39x10' 1.39x10'2 1.00 1ITime steps too large for the computer simulations at these time rows. 146 in YDE throughout the 10'3 M solutions is greater than it is for the 0.1 M solutions. (Assumption 3 of Section C of this Chapter is that YDE is constant throughout the solution.) The potential differences for the 10"3 M simulations are larger than those for the 10'1 M simulations because the relaxation is slower and because the equilibrium charge distribution is more diffuse. E. Comparison with Previous Results Anson 32 El: [1969] have experimentally measured poten- tial differences as a function of time for double layer re- laxations at dropping mercury electrodes. In a related study Newman [1973] has analytically solved the Nernst-Planck equa- tions for the boundary conditions discussed in this chapter. Besides the previously discussed thermodynamic assumptions inherent in the Nernst-Planck equations, Newman has made several other approximations in the process of solving them. He assumes that the charge on the electrode versus time is always of the form shown in Figure 6-8. Moreover, he assumes that while the electrode is being charged the ions move only due to migration in the electric field, and that after charging is complete, the ions move only due to diffusion. Figures 6-9 and 6-10 are comparisons among the ex- periment of Anson £3 a;., Newman's analytical predictions, 147 Figure 6.8. Electrode charge versus time profile assumed by Newman [1973] for analytical solutions to the Nernst-Planck equations. 148 w.w mpswflm . 0 “ Figure 6—9. 149 Comparison of the experiments of Anson e: §l° [1969] (A), simulations using IONFLO (B), and Newman's [1973] analytical predictions (C) for the relaxation potential of HCl at a blocked electrode. a = 6 x 106 8.1, 00 = 10-5 C/cm2. Time axis is logarithmic. 150 m.m opzwflm 3:. 0.0.. 0.0.. 0.0... 0..... ON... . _ 4 _ . 4 N0 0 I 10.0 m 1V0 lllllllmlllll/llll I 10.0 _ _ . A/(IOH) 0V Figure 6.10. 151 Comparison of the experiments of Anson 23 MM. [1969] (A), simulations using IONFLO (B), and Newman's [1973] analytical predictions (C) for the relaxation potential of KCl at a blocked electrode. a = 6 x 106 8.1, Q0 = 10'5 C/cm2. Time axis is logarithmic. 152 oa.c cpzwwm C. c. 00... OK... 0.0 .. 0.0: 0.0. .. . _ _ _ N0 I / .. 0.0 1. ..v.0 / 1 0 .0 L _ . . A/(IOXMV 153 and the simulations using IONFLO, for HCl and KCl relaxa- tions respectively. Time has been plotted logarithmically so that all results could be displayed on the same graph. Anson's transients are much longer in duration than either of the theoretical predictions. However, he reports varia- tions of a factor of 10 in the lifetime of the transients due to the shape of the capillary used in preparing the mercury drops. Because of the large charge on the elec- trode (10-5'C/cm2), the computer simulation results are subject to truncation error for the region near the elec- trode. The only definite agreement among the results is that the potential difference due to KCl decays slower than that due to HCl. CHAPTER 7 SUGGESTIONS FOR FUTURE WORK This chapter contains two sections: (A) suggestions for future thermodynamic studies of ion transport processes, and (B) suggestions for upgrading program IONFLO for future use and for validating it in its present form. A. Future Thermodynamic Studies (1) It does not seem possible at present to make ob- servations in the first 10.8 seconds of a liquid junction potential experiment. If the balancing of forces hypothesis is correct, it may be possible to slow the balancing time by initially imposing a temperature gradient on the system. This can be done by using constant temperature reservoirs at the two electrodes. A measurable potential difference will result (Agar and Breck [1956]). At steady state the concen- tration distributions will be approximately linear in space (Horne and Anderson [1970]). If the reservoirs are switched to a third, mean temperature reservoir, the con- centration gradients will begin to decay to zero, as will the junction potential and the entire cell potential. The decay time will be on the order of the relaxation time of the 154 155 thermal gradient in the cell (>1s). Besides the equations discussed in Chapter 2, the additional equation needed to analyze this experiment is the conservation of energy equa- tion (Hasse [1969]). This system of equations is quite com- plex, and a numerical solution may be required to analyze the data. (2) For many e1ectrode/electrolyte/electrode systems the double layer relaxation boundary conditions are more complicated than those used in Chapter 6. For these systems it is necessary to account for specific adsorption of ions at the electrode. To analyze these systems thermodynamically requires adding another region to the solution domain with its corresponding boundary conditions. This new region would be too narrow to allow for conventional ion transport within it, so the main equation to solve there would be Poisson's equation. The solution to the equations in this region and in the previously discussed region would be joined by kinetic boundary conditions as described in Chapter 3. This is also a complex problem and would probably re- quire a numerical solution. (3) Turq 22.21: [1979], using radioactive tracers, have reported the appearance of transient concentration gradients from initially homogeneous species in multi- component electrolyte solutions. This experiment could be analyzed analytically by a direct application of the matrix method described in Appendix F. Other more complicated 156 electrochemical and biological multicomponent systems could be analyzed numerically. Miller's [1967] ternary system data and n-component projections would be useful in these extensions. (4) By necessity, Onsager coefficients have been cal- culated from measurements on essentially electrically neutral solutions. Thus, it is not known how they depend on 0+ and C_ separately. Pikal [1971] and others have postulated various models from molecular considerations. It should be possible to test these models for macroscopic manifesta- tions using program IONFLO. (5) It is generally accepted that the dielectric co- efficient is not a constant, particularly near charged inter— faces where the water dipoles become oriented in preferential directions. Although the electrostatic implications of a non-constant dielectric coefficient are more fundamental (Perram and Barber [1974]) than the subject of this dis- sertation, the equations necessary to test various dielectric models are contained in this dissertation. Again, a numeri- cal approach may be necessary. B. Program Improvement Suggestions (1) IONFLO has only been used for the boundary condi- tions discussed in Chapters 5 and 6, and for a narrow range of electrode charges and concentrations. It should be validated for other applications such as membrane transport 157 and solid electrolyte transport. (2) At present IONFLO reads in expansion coefficients of Onsager coefficients and activity coefficient deriva- tives. These are used in function subroutines to form the_ Y transport coefficients. These functions are called col- lectively about 5000 times per iteration per time row by subroutine COEMAT and RHSVN. It would be more efficient to form expansions of the Y transport coefficients as functions of S and D directly from the input data at the beginning of the simulation. These direct expansions of the transport coefficients could then be transferred to streamlined func- tion subroutines for use by COEMAT and RHSVN. (3) IONFLO should be redimensioned to permit simula- tions of multicomponent systems. Temperature could be treated as a component with some minor adjustments for en- thalpies of mixing. (4) The dielectric coefficient should be programmed to have the capability to depend on the dependent variables. (5) The space grid and boundary conditions should be expanded to permit use of multi-region domains. APPENDIX A TRANSPORT COEFFICIENTS The y in Equations (2-3A) and (2-35) are 13 788 (RT/2z+z_)(ae - bg) YSD = (RT/22+Z_)(-af - bh) (A-l) YDS = (RT/2z+z_)(-ce + dg) YDD = (RT/2z+z_)(cf + dh) , with m ll 0‘ u (Z_/c+)(z+£++-Z_£_+), (z+/c_)(z+2+_-Z_£__). O I Q. I - (z_/c+)(z+2+++z_£_+), - (z+/C_)(z+2+_+z_£__), e=1-%(z+/z_Xz+-z_)2c+(M+B), f=1-%(z+/z_)(zi-zf)c+(M+B), g=1-%(Z_/z+)(z+-z_)2c_ X = a can be represented by the equations LJ(X) = BJ j = 1,2,...RR , (D-2) where each value of J represents a row of the matrix. Expansion of L (2) in a Taylor's series about the RR J variables yields, 1 RR RR 32L - ( ) A AY AY + (D-3) 2 RR 3 L 1 J 2 - ————) (AY ) + 2 i=1 3Y12 1 ’ 167 168 where L = LJ(?O) and AYi = Y O (D-A) o J 1'Y1 Truncation of the Taylor's series after the AY1 term and substitution of Equation (D-3) into Equation (D-2) produces RR 31; 1-1 (§§i)§_§OAYi = R3 - L0 J = 1,2,...RR . (9—5) This can be written as the matrix equation, £0 , (D-6) IN) I QLY= where the matrix elements of g are defined by, L}? ) . (D-7) G = ( A A 13 1 X‘XO m K: This suggests the following iteration procedure: 1. Choose 20. 2. Calculate g and 20 using Equations (D-4) and (D-7). 3. Solve Equation (D-6) for 4:, A. Use the last of Equations (D-A) to calculate 2. 5. Compare £2 and 2. 6. If IAYi/YII i = 1,2,...RR is less than some specified tolerance, the iteration is terminated and the program proceeds to the next time row. If not 169 + 20 and the procedure is repeated starting with Step Ir<> Of course, there is no guarantee that this procedure will lead to convergence for all initial choices of 2°. Program IONFLO has two options for choosing go. Option A chooses 2° to be the solution to the previous time row. This is simple to use and the time step can always be decreased until the procedure converges. Option B projects the solution at time row n to the solution at time row n+1 using space derivatives and transport coefficients evaluated completely at time row n. This projected solution is then used as go. These values of 20 are almost always more ac- curate than the ones from option A, so larger time steps may be used. Option B is usually more efficient for rapid approach to the steady-state of the system. APPENDIX E LISTING OF PROGRAM IONFLO The following input deck executes program IONFLO: ATTACH,MAIN,IONFLOBINARY. ATTACH,X,JHLLINPACKBINARY. ATTACH,Y,JHLBLASBINARY. ATTACH,T,AZCOlBINARY. ATTACH,U,INTEGRATIONS. ATTACH,W,EQNlBINARY. ATTACH,V,GRIDPRBINARY. ATTACH,R,JHLPLOTBINARY. LOAD(MAIN) LOAD(X) LOAD(Y) LOAD(T) LOAD(U) LOAD(W) LOAD(V) LOAD(R) EXECUTE. The following is a listing of program IONFLO: 170 OOCOOOOOOOOCOOOOOOOO 000000000000OOOOOOOOOOOOOOOOOOOOOOOO000000000 171 PRCXSRAM IONFID (INPUT,GJT‘PUT,TAPE GIIOIJ'I'PUT‘HIAPE 50=INPUT,TAPE 7) PRERAM IONFLO SOLVES THE SYSTEM OF EQJATIOI‘B GENERALIZED NERNST- PLAMIK, POISSCN DISPLACDIENT‘ CURRENT FOR THE TRANSPORT OF A BINARY EIECTROLTT'E IN A FINITE SPACE IXNAIN. DERIVATIVES ARE APPROXIMATED USING FINITE DIFFEREMZE. 2T'YPESOFBOUNDARYVALUE PRCBLEMSAREPESIBLE. KEY=1 ISE‘CRT'HE DIFFUSION POTENTIAL KEY=2 IS FOR KINETIC ELECTRODE MODELIM; BOI‘H EMPLOY A NON-UNIFCIM SPACE MEH WHICH DEPENE ON THE MEAN CQCENIRAT‘ION OF THE EIECT'ROLYTE AND THE INPUT PARAMETER INPUT ALL INPUT IS READ THRGBH SUBROUT'INE AIO. THIS SIBROUT'INE SHOULD 3E INSPECTEDT‘ODET‘ERMINE PROPERGRDERAI‘DFGRMATOF INPUT‘ THEFOLLONING ISALISTOF meVARIABLESAmmEIRMEANm. TITLE DESCRIPTION OF SIMULATION KEY 1 FOR DIFFIBIQI POTENTIAL SIMULATION 2 5m KINETIC ELECTRODE MODELING R NLNBEROFSPACE POINT‘SUPTOIEB Q NLHBEOFTIMERGIS ISEI‘ NIMBER 05‘ W SPACE POINTS AT EACH ELECTRODE on SIDE OF ACE IJK FRACT‘ICN OF TIME ROVS TO BE WRITTEN IST'OP NINBER OF TIME ROVS T‘IME STEP SIZE ARE‘CCIBT'ANI‘ M14 SPECIFIES A TIME ROI HR WHICH MANY INTERNAL PARAMETERS BE WRITTEN IN UNFORMAT'I‘ED OUT‘HJT. THIS GENERATES A EDT WPPUT‘ AND SHOULD ONLY BE USED FOR DELGGM. WHEN NOT USE SEI' W GREATER THAN Q. NPLT‘ FRACTIQIOFTIMERONSTTRT‘AREWRIT'IENT'HATARET‘OBE PLOT'I'ED WILL OF IN IOFF 1 Fa? C(MPLETE ITERAT‘ION mm a PG? no ITERAT‘ICN IW GWRITESOUPPUTCNLYFCRIASTITERATIWOFEPCHTIMEROI SPECIFIED BY IJK lWRITESCXJ'I‘PUT‘FCREVERYITERATIwOFEACHT‘IMERm SPECIFIED BY IJK ISTUPIAST‘TIMEROIITT‘RT'CURRENT‘ISTURNEDINFE RAEE DETERTINES RATE AT WHICH GRID SPACING EXPANIB FROd SMALL STEP SIZE TO LAEE (NE. 171‘ INITIAL TIME STEP SIZE IN SECQ‘JE F DETERMINE RATE OF TIME STEP SIZE INCREASE. TOL MAXIMIMAILOIEBLE FRACT‘IQIAL CHAEEFROJICNEITERAT‘IW T‘O T'HENEDCI‘FORPRERAMTOPREEDETOT‘HEWTDIERGV ZP CHAEE NIHBER Em POSITIVE ION ZM CHAEE NIMBER Fm NEGATIVE ION 0 0000 0000 0000 0000000000000000 00000 000 0000 00000 000 0000 0 00000000 000 000 172 20 FINAL m (N ELECTRCIE IN C/OI"*2 ALF TIME Cm Fm CHAEE INJECTIW A CHARACTERISTIC LEMSTH IN (N GDE MEAN 2m VALUE IN MOL*S*C/G*CM'*3 SIZE LEICT'H w SYSTEM IN 04 EPM MEAN DIELECTRIC CIEFFICIENT IN C*C*S*S/G*QT**3 84 MEAN ELECTROLYTE CGCENT‘RATION IN MOLE/04”“ [030,169 KSLNDA,KSAI~ooo Q05 “8 “SE :9me 0301 EQEQQ‘OQQO‘U‘ID aawromméw SQ ’KDCDQO‘U‘thN EGOQQO‘U‘MNH "DMO‘UIIBW SQ \OQQGUDUNQQS mummwat-‘aas \lmthUQQGQQG O‘UIMNQQGGQG thwNO-‘GEGGQQ waQQQQQGQQ UNGQQQQQGGS NHSGQGGQEQG ass SQQBSGGQG OMQSQQQGGGQ mm bGGQQGQGQ awkwaaaasaa mthNEQSQQS GthUNl-‘QQQSG O‘U‘bNQQQQSQG mmbwmaaaaaa OSU'IDWNI-‘QGGQQ O‘U‘lwaQQQGGQ GUIWNQQQGGG O’SUlchUNI-‘QSQQG GMMQQQSQQQ GMhUNQQQQQG O‘thWNt-‘Q&QQQ 000000 000 0000 MM 180 secs aaaa as aaaa scam see» $330 013% omqm mqmm 5 9 GUI-DU DION 0 90 WDKDM REAL KDO,KSO,KDL,KSL,KDR,KSR Q>DW seem WFWQQ “>0“ QQ‘OM wymmq a>mmq EEOC)“ mywcoq $3,me QQ‘OGQ QEQWQ QEEQQ INTEGER R RR RSTOP 80M wl%%}a%éy BSGAEIlaRSSGRSAgASA KDR KSR SR SL m DL C /ELEC/ EPS'Z£ I I I I I I I I I I $333“! IBflP=IMI-5 DENIUZDBTHEMmflmmHXTOZfim DO 130 91:1m .16 A3%I?J 865.5 MEDBUNSflfllflflflXEUflflflSAflEDflImfllfimimmmMN ammmmusxrww 1352'=D1§S£11{*DEL(1)/2. 0 2%} F1 = 1.91/01 =KSH/DzD F3 = zSS(s D)/D22 T1 = F1+ 5‘2 + F3 T2 = zSSP S,D)*Y(1)/D22 %; 3 T2. /2 2mm) T5 = i/Aégg ° T6= ABD(11 1)= F1 : 2DS(S 2D)/D22 F1+ 52 zmpés, Wig/022 -ZDEPPSé:DLY -HEP *D -5354; :g; M” 2. -2DDPS 5“€ KISS/D SDQSD Z 22 -233; S 2,} 5:3} 655352333333 I MIIINIIIIIIII HH D afidfifig S§:$ .‘DZ 10,2/D g/(z 0 ) 4ZDDS:I%¢R2 IIIIIIIAIIIIIII'IIIIIIIIIII :SSEEHESR 2+T3+T4+T5+T6 2+T3+T4+T5+T6 T3+T4+T5+T6+T7+T8+T9 00000 181 T =ZDSDS,ZD*Yl/D2 T = musi' 133* vf23/D2 22 T7 = -2DSD S,D D*Y 4 /D22 T8 = -ZDDD S,D *y 5 /D22 T9=-ZDEDSD*Y2/EPS Tm -ZDE (S, )* ( )/(2.a* D2) ABDéll, =T1+T2+ T3+T4+T5+T6+T7+T8+T9+T1z 2 2 = KsyEPS 1 = -is S D?) (2.9MD2?D T2 = -ZSE SP P)/§2. 2* 2) ,3?) = Ti + T T1 = —z E S D))/(2.D*D2)J T2 = -2DE SP P)/+2. 3* 2) D 91,3) = T1 + 2 ABD 11,3) = ABD8, 4 = —zSSS,D/D22- a. 5*ZSEPSP,DP/D2 *Y3 $8 9644):” -2DS S, D /D22 - a.5*2DEP SP, DP /D2 *2 3 ABD 7,5) = -zSD S ,D)/D22 —. 0.5*ZSED SP,DP /D2 *2 3) ABD 8,5 = -ZDD S ,D)/D22 — 0.5*ZDED SP,DP /D2 *Y 3 ABD 9,5 = ABD6 6 = 0.9) ABD (3,6) = 3.9 ABD(8,6) = 2.0 LING THE INTER ASSEMB IOR MATRIX Em IT IS CQNENIENT T0 LOOP IN MULTIPLES 0F 3 BEAUSE THERE ARE 3 EQJATIOI‘S AT EVERY GRID POINT m 1Q1K=4 ,RSTOP,3 1013“; 1)/3 fK-33 K-3 K+3 3:13: 2&3 +1{)/2.a K DP= = WEY yKfi + Y(K+l))/2.l =DEL KM + DELéKK) + Y(K))/2.a SPP:Y T2 = -ZSSSM*:1§K-3/DEM T3 = *YK-Z T4 = ZSSP ”A“? éfim a ZSEPSMQ RA {Dézmm/AD T4 + T5 + T6 :3" {53“‘3mm ZDDP 'm *YiKMD 2138 9451.113“) RAMK+ 81313913 =I=T1+ +334. +)44+T5+T6 ABD 15,K- g? 16,;333 = zm(S, TD)*RAT/(EPS* AD) T2 = -ZSS 134W K-3 /DEN\1 T3 = -2 T4 = ZSSD T5 = 25145ng K+l)/DEM 182 a 2S? Sm mo*RAT*Y K+2 T1D(13£11§D -§4='T1/&§42 4344 + T5 4- T6 T2=- ZDSDSM,DM*YK-3/Dem >0 9‘6?“ ”Maw DDD 94:34 *Y K+1)/DEm a T1 + T 4 + T5 + T6 51135, D) *RAT/(EPS AD) 919 0.0 T2 = zss SP,DP /DENK T3 = 255 sum *le T4 = -zs S,D nap“ T5 =( -zsap D*S K4m +2)/mm ABD11K)=- ST1+T2+T3+T4+Ts zfiSSM m/DENK *YK/+l /1-:ps T4 = -ZDEP=§ §*Y K+2 *SUB/MUL ABD(12 K T4 T1 a 266 ES,D 1*Y K-Bi *RAT EPS*AD T2 = zDDpi S'D *Y K-2i BPS*AD T3 = -zDSpé /PS*AM T4 = -2DDPS*YK+1S*S/(RPS* ) T5 = -ZDEP D*Y K+2 /EPS T6 = -ZDSP S' D *Y K+3 / RA'IV‘EPS*AD T7 = -ZDDP S' D *Y K+4 / RA'I'*EPS*AD T8 = -ZIB(= .151 SUB/(E AMUL ABD(13K)= +T2+T3+T4+TS+T6+T7+T8 T1 = zSD SP, E)/D15:KK T2 = 281835;? T5 = l-ZSED?S?DE*SLSB*Y(K+25/1§MUL M(g+ ='=1+T2+T +T4+T5 T2 = ZDD SP, DP m T ..T T $6, TT) T4 = -z Sé 'AD 395m: T5 = - T6 = -2DED{S' D2.*S+KT *Y(K+ 2W A13D(1%ls‘5 +T +T4+T5+T6 T1 2 SM K-3 i*RAT/ EPS*AD T2 = D*Y K-Zi *RAT EPS*AD T2 = -ZDSS 1“P1101803 PS*AM m)“. T = (EPS*AM ) T5 = -ZDED 5' D3 2/EPS T6 = -2DSD S, D *Y K+3 / RAT*EPS*AD T7 = -2DDDS D*Y 54(4/ /RAT*EPS*AD T8 = -zw( ,b) AMUL) ABD(12K+1'=T1+T2+T3+T4+T5+T6+T7+T8 $2 a 2% $95” DPP ”5ARAQBAD) «r3 = -zsaism sub Timur. ABD(9 K+2 2D 1 + T2 + T3 = ins: Dm *RAT/AD T2 : :Tégafgseéagéwm 36 K - T1 + T2 4- T3 533,5 2’4"” ' §DZ112§§P 1 + T2 T2 =- zssp§SR,Dp;‘*Y(K)/DENK O 000 101 183 Tit-5335:; 592?" %§& mflR‘E‘RT T6 = -ZSEPS 113/DEM“ Y(K+2)/(AD*RAT ABD(8,K +3)=PT% +T4+T +T6 T1 = -2Ds SR, T2 = 205? SRIDR *Y K) IEmix/Diana?” T3=ZwPSPDP*YK+1yD T4 = -2DS R mm T5 . -ZDDP SR, DR *Y K+4 /DENK T6 = -ZDBP SRR DRR Y(K+2)/(AD*RA ABD9K+3T1+2+T3+T4+T+T6 ABD1RéK+ £21): -ZDSMD)/(E:*RAT*AD) T 1 T2 = ZSSD SR, DR *YK $31 a ZSDD SR' {:DP *Y ”135% T5 = -ZSDD SR, DR *YK T6 = -ZSED SRR DRR/)1.2 (K+2)/(AD*RA ARD(D7,K+4)=T1+ +T3+T4+T +T6 T1 = -ZDD SR, DR /D T2 =- ZDSDS RIDR *YK )K/DE NK T3 = ZDDD SR, DR *Y K+1)/DENK T4 -2DSD SR, DR *YK K+33/DENK g =- -Z-I uSR'D a -2DBRS SRRD Rzz (K+2)/(AD* RATE ABD8,K+4 TiD+ 2+T3 3T6+T4+T+ ABD 9,K+4 521.30 (S, D) / ( EPS*RAT*AD 5 3‘ 2‘. 5 a; IIIIIIIO) 0.. ASSEMBLIMSTHEMATRIXEWARISINGFRCMIHEW CCNDITIOINB AT X=A B§2==D§EL RRR11{§RBL(R-1)/2. a S a Y 3(RR- D . Y -1 ‘ -zss S, D/D 2+ *ZSEP $14,114 /D2 RR -Zm28: DI/flRz 2+ 5 ZDEP SM ,mI/ngwi RRI -23DS,/DD 2+ *zsman,m/D2*YRR is: 62% 22+ 2 i an? .5 .*5 .5 .*ZIED5 Slhm/DZWRR a U R E lllllllllllllllll F éSE:(S§:fi D D:*Y1RR)Z£2. 6*D2) §D(11,RR-2$ + T3 + T4 + T5 + T6 52 : 21% ”(ESéD) /D22 T2=zDRRSD*YRR/ 2.a*D2 T3 - -2w (1%, R) 4 (RR-iszz I 1 102 000000000000 0000 1'3 1814 _—_- * _ $4 = g8§PES,D ,D YéRR 2)/D22 $3 = IRl PP(R2* Y = + T6D = -2DSR S ,R) ;R(RR;1R)/D22 WI g= +B+M+fl+fi D13I RR-z = T = 2 $3 3 2:3 SéD/ 22 T4 = -ZSSR R, /*Y RR-Sg/DZZ T5 = SDDS *Y RR-4 $9 3 ZSSERS,6E*YRRR- RR 22 T8 = -zs (é, )* (RR- )/ERS T9 = ZSED(S D)*YiRR)/ 2. 2 *D2) fiW-Qf 15 +n+w+m+m+m+m+m T2 = $2 3 29D SzD/ /D22 T5 = -2DSR(S,/EPS )*Y(RR-5)/D22 $9 3 EESDDDD, S bDlyYéRRR RR-2 ‘4’ 232 T8 = 22 MR *Y RR-l /D22 T9 = (R, ;* ERR- )/ERs T10= 2DED s, D *Y lRR)r/(2.E; D2) gmz’m ‘3: 2+T +M+fl+m+w+m+m+na ABD 9 éRR = 25 S E/ 2. E*D2) + 23 SM, DM / 2. 0*D2 1,RR) RDR(R,R) }(2. 2*D2) + ZDR(SM, DR)}(2.E*D ) IF M .NE. MMM [GO To 123 D01 WRITE2 (61,1 iRBD(IM,JM) %Tbé2x, CONTINUE RETURN END SUBROUTINE RIBVN (R,IEL,U1‘, Y ,B) WIEN IFLAG=0 TIE FIRST TIME RIBV IS CALLED AT A GIVEN TIMER SUBmmE CQIPUTES 'IIERHS VECTOR FORTHE H%TION) AT 'I'HEN-I-lTHTmERONFROI'I'I-IESOLUTIONTOTHEMA'IRIX TIE N TH TIME RON. WHENIFIAG=1, RHSV COIPU'I'ES'IIE 'IIECURRENI‘ VALUEOF'ITEAUXIIARYVEC'IOR. INTIEMAINPRCXSRAM'I‘I‘IISVEC‘IOR WILLBEAIJDEDTOTIERHSVECTORTOFQQM'HECURRENTRHSVEXHOR FOR A GIVEN ITERATION. R IS THE NLMBER 0F SPACE PTS. TH IME RON IF IFLAG=G AND PREVIOUS ITERATION STEP IF IF LAG=1. B RR) IS THE RIGHT HAND SIDE VECTOR IF IAG=0AND ANDTIE AUXILARY VEC IFIF LAG=1. THE (NLY DIFFEREI‘CmE‘S’o BE'IWEEN THE RIB VECTORAND THE TEAUX VECTOR AREmTI-IE'I'EMSDNO HEVINGDI‘ANDINTIEBGJNDARY MS. REALIGDOKSEJQDLIISSLWKSAKDAKDRKSR INTEGER DTMENSIDN' DE 32 Y 62 B 322 /Rc::TR/ KRéiRSé,I<15L(I(SL),KDA,KSA, KDR,,KSR,SR Siam, DL CDMMDN/ /ELEC/ EPS,ZI 000 00000 185 RR=3*R RS'IOP3RR-5 vm ELEMENTS ARISIMS FRCM THE Bum CWDITIOBB AT X30 D2 =- DEL(1 /2.0 D228 ay 811??“ )*DEL(1)/2. 0 DaYé 353—{5/02 ms-Kéa 302 P3a-z 1‘1: 21322442323? $3 45 41363.8: ASZ %:NZ%SE(§§%§5 912859432 2)*) (31m) T5 = ZSD S,D /D22 *Y 54 B 1) 8 T1 + T ‘1' T5 ‘1' =- (-KD02 - +zm(s, D)4/D22)*Y(1) 2‘2 Ksa/Dz S2 " 2%.”?232965’ 22 '12: 214/22 3+F4)*Y2 $2 = guns /2. 6*02) + 2&(39. DP)/(2. amznwm . . ... s 4.22;. . $32; =1032/ saw? 5 T4 T5 T . -§:ng/ 5M5 ) 3 3): 21 + +1- VEC'KR Em FRO! THE INTERIOR SPACE POINTS THE m-LmP IS IN MULTIPLES OF 3 SINCE ‘IHERE ARE 3 EGJATIOBB AT EACH SPACE POINT 91 K=4 ,RSTOP,3 no 1 22 " 2:2 53 34 = Y K-3 + Yam/2.2 em = Y K-3 5 =- Y n+3 39': 5:) x+3 YK) /2.a w= YK—Z YK+))/20 B": 2 2.2? ngan 1;: + Y(K+1))/2.a Ian-02:01:24 +03 ADéKK) 222 =22 $22212 3‘ mu. 2 DE *DEL an a DEL 81:“) - 0:1}.(131) F1 1. 0 F2 3 -2 SP, DP /DENK m-4$m T4 - (n + Film + 3)*Y(K) 000 000 000 ' 000 186 F122: Iéé'sig'fiwfifi 2F131>+S'Fé/EPS+ F3 2(*Y( +1 = ZSD sp Pinpgwi Rig/Dam VECTOR ELEMENTS mm s summon BiK)=T1+T2+T3+T4+T5+T6+T7 T =- we SM,m)*Y(K-3 BN4 ) $2. : 22. 22.22222 3.“... T4 T5; Exzaoép, W13? +4i/6Dm T63-ZDDSP,DP*YK+ ZDE(S D5 (F1 + ’F ”Yaw-1% zusuésp pr *Y [(+33 ) -z 13.335022. Ma 21238 pp ,Dpp)/(RAT*AD) 208 s, ,Défs UBAMULé T10 = (F 2 + 3)*Y(K+2) m swarms mm D EQUATION 8991) =T1+T2+T3*+ mgv+TS+ +T6+T7+T8+T9+T10 a -2Ds s, D: K-3 EPS*AD; = -zmé§lg Diwlgx-zz/ 3332;422:929} T3 = zns *MULLL =21!) ,D* *S 3%; -13' 126312 m/éps)m .1) T68 m S, D *YéIzG-Bg/ S‘EPS PS*RA'NAD Wflwa,D*YKM/EPS*RAT*AD WWMBEGIATIW BK+2 =T1+T2+T3+T4+TS+T6+T7 351339 “I {38333333 I II II II II II II 101 C vacrm Em ARISIMS E'Rm 'IHE BGJNDARY CQIDITIOIS AT XSA D 8 =DEL éR-l 0 éR-1{£DEL(R-1)/2. 0 S 3 RR. §L YEAR 1 ’ 3132235131ng (RR -5;/022 T2= 28D 3933' *Y -4/D22 F1 = -1. F2 = - F3 ... -z 88(2DS T3 ... (F i’éz + m:3)*Y(RR—2) F1 = - 2 F2 . -ZSD s D) F3 = 233( .15 $15: a (1.25st ,D) 2. 3)2'132)“ + 4zsr: SM 130/(2. 0*D2))*Y(RR) 322-... 2, .352 2.43" ..2 T a 2m ”(ERR-5 T/D22T4 T2 = no s'D *Yf ran-4 /D22 5» ‘ ’2 1?ng F1: 15+ 8&2 *Y(RR-2) JUN 0000 100 00000 187 -Zw 82D)2 3222138 (F1(+ D5fi3+g4w §3£WI(ZDB3S,D)432.B*D2;+izw3é‘13M/(2oWDZHWmR) gmgrfRR-l) 13% ‘22 ' " SIBRUJTINE RIBVO (R,IEL,UI‘,Y,B,M,M4) SUBROUI‘INEREBVOGENERA’IES'IHEPORTION OFTHEBVEII'IURTEETIDES MGMWITHEACHITERATION. REAL KDflfiKSO, Lm'II'DKSL' ,KDA ,KSA,KDR,IGR mmm 81113125 ION EgméapsgfiéixgggkL ,9ch ,1+ zm-ZPM)/(S+D+1.0) T4=U4*V4 ZIEPSTl-i-TZ-T3-T4 END 9; **3 6 **3 2 G \ + 2PM 0/ S - D + 1.0)**2 8.53% NO END §%§S+D+1.m+m-AB NCl-‘NCD-‘NC HNOO II II SEESSSSSSSSSEZ “llllllllllllllllllllllllll fl??? D + 1 awrz u4*v" 4 ' ZSSD=T1+T2+T3+T4 RETURN 33'"? :r'6 .1. C NN “55 II ll O War-(.3 1' l- t N V I HH 0 o b-b-s GU“ A V mm .1328 M 555355 195 $303 9%1'45’35/"2 RENRN COMQJ/ IF (S .13. -1.a) S a -1.0 + 1.08-la ZPP=C1*S-D+1.0g+C$2} s-D+1.a;**3 zm=c3*s+o+1.a +C4* S+D+1.0**3 ZPM=C5* ((S+1.0;**3)+C6) s+1.a)**2 ZPPD=-C1)-1.5*C§2;* Tés' + .0) zmn=-C()+1.5*C$ (+D+1.a) Vls-PP-ZPM 'r = U1+v1) %- S+D;m.a)*m w2= S+D+ .m**2 T2: UZ+V2 zsw-T1+T RETURN END WIONZIEDSD coum/ 6(4) ZPP=C1*S-D+1.0 +c * S-D+1.0**g zm=c3*s+o+1.I +C *S ‘r s+n+1.a** ZPM=CS* ((S+1.¢£**3)+C6) S+1.0)**2 ZPms—C§1)-1.S*C523‘* és- + .0) m=C()+1.§*C) (+D+1.I) ~ AM=C7 -C(8) (3+1.a) fiac9 name“ 3%”?ngng lfl)**2 = o - + 0 33:0?“ ¥2=$.%SS+D+1.0)**2 znsna 1+T2+T3+T4 END mm 2339 s D) cm Nun's éq GDE,EPM,A,ZP,ZM,CCN ccmou {G (3&4) IF (S c To -10”) S a -100 + 1.03-10 zppp . cu) + 1.s*C(2)*SQRT(s - n + 1.91) 196 E'é‘z‘fi : §$3§4E$é§§$filifl§$§ 1’ ‘z’pfap 178211 END mum 21189 S D) x comm mum's én GDE ,EpM,A,zp,zn,cm ccmw c; C(é4 IF [gs .mcs = -1.a+1.aE-1a 299 ='cTi +1. S*C2*SQRTS-D+ map: + 1. 5*c 4 *SQRTS szp 1. 5 21189 : 3. 894é—éGSgfi*A*A*(ZPP + 21149 pitgzfitfiflm END SED S D Cwmm% ITS}C S36”: ,EPM,A,ZP,m,C(N COMO! /GREEK/ ZPPD= :ngZéJ.)+-11(:(S *Cé £21§$LEDS£ D+1+1.0) mamas-03914111111 lZDPPD RETURN END nuc'rxou 2). -s D) comm Nun's: éa éGm, EPM ,A,zp,zm,cm comm /GREEK/ ZPPD=-C1:S§*&i£2*ST +115“) man-+1.5C DD++) ZDEDS 3x8 4E-06*A1A* ZPPD RETURN END APPENDIX F SOLUTION OF SYSTEMS OF LINEAR PARTIAL DIFFERENTIAL EQUATIONS In order to solve Equations 5-6 for S and D, they are expressed in matrix form as pg = o , (F-l) where E is the linear operator matrix and Q is the solu- 0 tion vector. If w is defined such that DET(g)w = o (F-2) then it can be shown that, Ui = COFJi(P)w , (F-3) where J can be any i (the cofactor can be expanded along any row). For our system, 197 198 (3/3t) - YSS (a2/ax2) 'YSD(32/3x2) + YSE (32/3x2) (a/at)-YDD(aZ/ax2)+yDE , (F-u) ‘YDs C} II The equation for w is [A(3u/8xu)+B(33/3t X2)+C(32/3x2)K32/3t2)+G(3/3t)1W = 0, (F-S) ~where A = YssYDD ‘ YSDYDS’ B = “(Yss + YDD) ’ (F-6) C G: = YSEYDS ' YDEYss ’ YDE ' Since W, in this case, must be an odd function about x = O, the solution is expanded in a Fourier sine series, W(x,t) = f, bk(t) sin (kflx/a) . (F-7) k=1 Substitution of Equation (F-7) into Equation (F-S) produces the ordinary differential equation 199 bg + [G - ”A - - (kn/a)203bk = o (F-8) The solutions of Equation (F-S) are bk(t) = L+k exp(-R+kt) + L_k exp(-R_kt) , (F-9) where R+k and R_k are given by Equation 5—8, and L+k and L-k are determined by using Equation (F-3) and the initial conditions. The result is Equation (5-7). APPENDIX G ANALYTICAL SOLUTION OF D FOR TIME DEPENDENT DOUBLE LAYER RELAXATION The time dependent system of equations in Section C of Chapter 6 is solved using Laplace transforms. Because this is a time dependent problem the following set of reduced variables is used in this appendix for simplicity. d = D/SB E = x/a ” = YDDt/32 k = a2YDE/YDD (G‘l) a