NONEQUILIBRIUM THERMODYNAMIC THEORY OF TRANSPORT IN MEMBRANES Thesis for the Degree of Ph. D. MICHIGAN STATE UNIVERSITY ‘ JlNG-SHYONG CHEN. 1971 LIBRARY Michigan Stave University '"fl ." . o. 1" This is to certify that the thesis entitled Nonequilibrium Thermodynamic Theory of Transoorf In Membranes presented by JIng-Shyong Chen has been accepted towards fulfillment of the requirements for Ph.D Jame in Chemistry Majorprofessor Date Dec. 9, I97I 0-7639 . -' * 9V\’\ “ ‘1: " BIKBING * HUAE 8 3%? 800K BINDERY INC. " LIBRARY BINDERS dsrnuman. memes] ABSTRACT NONEQUILIBRIUM THERMODYNAMIC THEORY OF TRANSPORT IN MEMBRANES BY Jing-Shyong Chen The steady-state transport theory that accounts for the osmotic flow of fluids through biological mem- branes is obtained, subject to certain simplifying assumptions which may be removed in subsequent work. The principles of hydrodynamics and nonequilibrium thermo- dynamics are used to describe the simultaneous transport of mass and electric current and the osmotic flow rate in a fluid system undergoing osmosis. In order to under- stand better the osmotic flow phenomena in biological systems, the following particular subjects are thoroughly investigated: (1) the distribution of ions in charged membranes, including the effect of molar volumes of ions, (2) the solution of Poisson's equation modified by in- clusion of the dependence of the dielectric constant of the salt solution on the salt concentration and the electric field, and (3) the distribution of pressure Jing-Shyong Chen including the effects of ionic concentrations and electric field. Their contributions to biological systems are dis- cussed. Moreover, the range of applicability of such classical monuments as the Boltzmann equation and the linearized Debye-Hfickel theory is examined. After these fundamental problems are dealt with, the transport theory is obtained subject to membrane -3 C/mz surface charge density of o z 10 (C = Coulomb; m = meter), which is apparently reasonable biologically. A capillary model is used for the membrane separating two aqueous salt solutions of different concentration at the same temperature. The theory is valid for total ionic concentrations between about 0.004 molar and 2 molar, a range which includes systems of biological in- terest and more concentrated systems. We assume that there is no hydrostatic pressure difference across the membrane. The steady-state solutions of the differential equations derived are obtained for the case of a symmet— rical, binary electrolyte, yielding the average osmotic flow rate as a function of initial concentrations. The result of the theoretical section is then used to explain, at least qualitatively, the osmotic flow phenomena ob- served experimentally. In particular, the theoretical equations predict both the shape of the experimental curves of flow rate versus concentration and the direction of flow through charged membranes. NONEQUILIBRIUM THERMODYNAMIC THEORY OF TRANSPORT IN MEMBRANES BY Jing-Shyong Chen A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1971 To My Parents ii ACKN OWLE DGMEN TS The author wishes to express his appreciation to Dr. Frederick H. Horne for his guidance throughout the course of this study. Appreciation is extended to Mr. James D. Olson for his computer assistance. Appreciation is also extended to the Department of Chemistry, Michigan State University, and the National Science Foundation for providing financial support. Thanks are particularly due to his wife, Mei—horng (Betty) for her assistance and encouragement during his difficult time. iii DEDICATION TABLE OF CONTENTS ACKNOWLEDGWNTS O O O O O I O O O O O I O O O O O O 0 LIST OF TABLES O O O O O O O O O I O O O O O O O O 0 LIST OF FI LIST OF AP Chapter GURES O C O O O O O O O O O O O O O C O PENDICES O O O O O O O O O O O O O O O O I I 0 INTRODUCTION 0 I O O O O O I O O O O O O O O A. B. C. II. THE A. B. C. D. E. Membranes and Membrane Phenomena . . . . . Motivation . . . . . . . . . . . . . . . . The TheSiS I O O O O O O O O I O O O O O 0 TRANSPORT EQUATIONS . . . . . . . . . . . Introduction . . . . . . . . . . . . . . . Chemical Potential . . . . . . . . . . . . Hydrodynamic Equations . . . . . . . . . . Principles of Nonequilibrium Thermodynamics . . . . . . . . . . . . . The Differential Equations Describing the Osmotic Flow of Fluids Through a Charged Membrane . . . . . . . . . . . III. EQUILIBRIUM DISTRIBUTION OF IONS IN AN EI‘ECTRIC FIELD I O O O C O O C O O O O O A. [310061 Introduction . . . . . . . . . . . . . . . Derivation of Modified Boltzmann Equation. Molar Volume and Charge. . . . . . . . . . Uni-univalent Case . . . . . . . . . . . . Discussion . . . . . . . . . . . . . . . . IV. RADIAL DISTRIBUTION OF ELECTROSTATIC P Am B. C. OTENTIAL IN A MODERATELY CHARGED CYLINDER. IntrOduCtj-On O O C O O O O O O O O O O O O Poisson's Equation . . . . . . . . . . . . Simplified Approximations. . . . . . . . . iv Page ii iii vi vii viii UTUJH l—’ 10 13 18 24 33 33 44 45 47 49 49 51 53 Chapter 6)"!!th Perturbation Scheme. . . . . . . . . . . Solutions of the Differential Equations. Discussion . . . . . . . . . . . . . . . Conclusion . . . . . . . . . . . . . . . V. THE STEADY-STATE DISTRIBUTION OF PRESSURE IN MODERATELY CHARGED SYSTEMS . . . . . . A. B. Introduction . . . . . . . . . . . . . . The Macroscopic Osmotic Pressure Between the Aqueous Phase and the Membrane Phase in a Moderately Charged System . The Macroscopic Osmotic Pressure Across a Moderately Charged Membrane. . . . . Osmotic Flow of Water in a Charged Membrane . . . . . . . . . . . . . . . Conclusion and Discussion. . . . . . . . VI. THE STEADY-STATE PHENOMENOLOGICAL THEORY OF OSMOSIS IN CHARGED SYSTEMS. . . A. BIBLIOGRAP APPENDICES Introduction . . . . . . . . . . . . . . Transport Equations. . . . . . . . . . . General Formula for uO . . . . . . . . . Special Formula for no and its Consequences . . . . . . . . . . . . . The Working Equation for Anomalous Osmosis. . . . . . . . . . . The Onsager Transport Coefficients . . . Results and Discussion . . . . . . . . . HY O O C C O O O O O O O C O O O O O O O Page 62 65 74 79 81 81 86 91 92 96 98 98 100 104 107 115 120 122 131 135 LIST OF TABLES Table Page 4.1 Value of 6a and V: of some uni-univalent electrolytes at 25°C (Sanfeld, 1968) . . . . . 53 4.2 Some calculated values of IO/Il and l/Il . . . 69 4.3 Values of a and b (k = 0,1,...,5) . . . . . 76 k k 4.4 Some calculated values of W, W , and W - W . -3 -2 O O O T = 298°K, [cl = 10 c m , a = 4.2 A, and Ka = 1.692 . . . . . . . . . . . . . . . . 77 5.1 Some calculated values of w and N' at 20 20 various salt concentrations. c (i) = 3 3 0 2.0 x 10 mole/m . a = 17.6 A. o C/mz. t = 25°C. . . . . . . . . . . . . . . . 94 II |—' O 6.1a The Onsager transport coefficients for Hzo-NaCl at 25°C 0 O I O O O O O O 0 O O O 123 6.1b The Onsager transport coefficients for H20-KC1 at 25°C. . . . . . . . . . . . . . 124 6.1c The Onsager transport coefficients for HZO-LiCl at 25°C . . . . . . . . . . . . . 125 vi LIST OF FIGURES Figure Page 4.1 The Dielectric constant D0 of water as a function of field strength [E] (in volts per cm). . . . . . . . . . . . . . . . . 54 4.2 Plot of the dimensionless parameter w as a function of Kr. . . . . . . . . . . . . 78 5.1 Plot of NRC (in joules per m3) as a function of the logarithm of the salt concentration . . 95 6.1 Plot of U0 1 (in cm2 per sec) as a function of the logarithm of the salt concentration for negative charge. . . . . . . 127 6.2 Plot of UO 2 (in cm2 per sec) as a function of the logarithm of the salt concentration for positive charge. . . . . . . 128 vii LIST OF APPENDICES Appendix Page A A Derivation of the Chapman-Gouy Equation For a Charged Cylinder. . . . . . 135 B Derivation of Equations (4.87) and (4.88). 139 viii CHAPTER I INTRODUCTION A. Membranes and Membrane Phenomena Transport phenomena across membranes, charged or uncharged, are encountered in many areas of life and physical sciences. For instance, chemists and chemical engineers would like to understand the mechanisms of membrane transport so that they would be able to fabricate membranes of any desired property or properties. Biolo- gists, on the other hand, are interested in understanding the behavior of complex cell membranes in terms of estab- lished physicochemical principles. A membrane, in simple terms, is a phase, inter- nally heterogeneous or homogeneous, which acts as a bar- rier to flow of some of the molecules and ions present in the liquid and/or gases in contact with it. Most membranes, except the obvious ones such as oil membranes, are to be considered heterogeneous. Membranes may also be classified natural or artificial. All biological mem- branes are natural membranes. Functionally, a membrane must be more or less active in the selective transport of some ions when used as a barrier to separate two solutions or phases, unless it is too fragile or too porous. Most artificial and natural membranes have been found to carry ionogenic groups either fixed to the three dimensional membrane matrix, as seen in a well-characterized ion exchange membrane, or adsorbed, as found in some colloidal systems (Kobatake, 1959). Ionogenic groups and pores (space occupied by water) in a membrane attribute certain func- tionality to the membrane. Thus a membrane may be con— sidered permselective and/or semipermeable depending on its functionality. The phenomenological transport prop- erty that controls the former is the transport number or transference number, whereas the latter is determined by the so—called reflection coefficient (ratio of the actual hydrostatic pressure required to give zero net volume flow to that which is required if the membrane were truly semi— permeable) introduced by Staverman (1951; 1952). In the absence of external strains and external nmgnetic and gravitational forces, the driving forces that may produce a flow or flux of molecular or ionic species through a membrane separating two solutions are (l) gra— dient of concentration, (2) gradient of pressure, (3) gradient of temperature, (4) gradient of electric potential, and (5) chemical reactions. These forces may Operate in various combinations and may generate a number of transport phenomena. To understand membrane phenomena, it is necessary to prOpose or design various transport systems or model systems consisting of membranes with par— ticular and predetermined specific properties. B. Motivation The transport of solutes and water through living systems has been the subject of rather extensive study since the 19th century. In spite of many experiments, the fundamental mechanisms involved in the transport phenomena exhibited by various living systems have re- mained relatively obscure. There are many reasons for our failure to advance more rapidly, but a number of the problems encountered in the study of living systems arise from the complex nature of living cells when considered from the point of View of a membrane system. Thus, trans- port phenomena in various cell membranes are vital to nmny biological systems. Among many theoretically-oriented investigators, Kobatake et a1. (1964) were the first to initiate a study based on the principles of nonequilibrium, continuum thermodynamics. They were quite successful in attempting to explain the experimental findings (Grim, 1957) with regard to the osmotic flow of fluids through biological membranes. Since then, various other continuum theories (sf. Toyoshima, 1967; Fujita and Kobatake, 1968; Gross and Osterle, 1968) have been develOped. Moreover, nonequilib- rium thermodynamics is often used to study the discontinu— ous fluid-membrane systems. The difference in the ap- proaches between the continuous and the discontinuous fluid-membrane systems will be discussed in more detail in the next chapter. Most of the previous theories, although adequate in many cases, are, in general, limited in the following respects: 1. Use of transport equations that are restricted only to dilute aqueous salt solutions (for detailed analy- sis, see, for example, Chapman, 1967). 2. Certain transport parameters are treated as constants. 3. The effect of pressure-induced current on the overall transport phenomena is neglected, implying that the pressure is considered independent of salt concentra- tions and electric field. 4. The dielectric constant of pure water is used in Poisson's equation, implying that the salt solution is very dilute and the electric field at the membrane sur- face is negligibly small. S. The Debye-Hackel theory is used arbitrarily and inconsistently; in particular, it is usually assumed (Kobatake and Fujita, 1964) simultaneously that the Debye length K-1 is very small and that the ionic strength is very small. 6. The classical Boltzmann distribution of ions is used with no attempt at justification. The main purpose of this work is to obtain a phenomenological transport theory for osmotic flow in biological systems that is not subject to the above re— strictions and that, hopefully, will resolve some of the discrepancies appearing in the literature (cf. Kobatake and Fujita, 1964). In addition, we hope to use the theory to explain experimental observations of osmotic flow phenomena. In general, it is hOped that our theory can be used to describe better and to understand further the membrane systems and the mechanisms of transport processes encountered in living cells. C. The Thesis In the following chapters, the steady-state trans- port theory that accounts for the osmotic flow of fluids through biological membranes is obtained. We make use of the principles of hydrodynamics and nonequilibrium thermo- dynamics in describing the simultaneous transport of mass and electric current and the osmotic flow rate in a fluid system undergoing osmosis. In order to understand better the osmotic flow phenomena in biological systems, we in— vestigate (1) the distribution of ions in charged mem- branes including the effect of molar volumes of ions, (2) the solution of Poisson's equation modified by inclusion of the dependence of the dielectric constant of the salt solution on the salt concentration and the electric field, and (3) the distribution of pressure including the effects of ionic concentrations and electric field. Their con— tributions to biological systems are discussed. Moreover, the range of applicability of such classical monuments as the Boltzmann equation and the linearized Debye-Hfickel theory is examined. The theory is obtained subject to membrane surface charge density of o z 10-3C/m2 (C = Coulomb; m = meter) which has been reported (Fair and Osterle, 1971) to be sensible biologically. A capillary model is used for the nembrane separating two aqueous salt solutions of differ- ent concentration at the same temperature. The theory is valid for total ionic concentrations between about 0.004 molar and 2 molar, a range which includes systems of biological interest (Woodbury gt_al., 1970) and more con- centrated systems. We assume that there is no hydro— static pressure difference across the membrane. The steady-state solutions of the differential equations derived are obtained for the case of a symmetrical electro- lyte, yielding the average osmotic flow rate as a function of initial concentrations. The result of the theoretical section is then used to explain the osmotic flow phenomena observed experimentally. CHAPTER II THE TRANSPORT EQUATIONS A. Introduction In this chapter the differential equations which describe macroscopic transport phenomena are presented. Specialized equations used in the study of osmotic flow of fluids through charged membranes are deduced, together with appropriate boundary conditions. We consider only continuous, isotropic fluids in which no chemical re- actions occur and which are subject to certain driving forces but not to a magnetic field. For more detailed analy- sis of the equations used in the study of various transport phenomena, refer, for example, to works by Horne (1966), Kirkwood and Crawford (1952), de Groot and Mazur (1962), Fitts (1962), Katchalsky and Curran (1967), and Haase (1969). The more fundamental, rational-mechanical ap- proach of Truesdell (1969), Bartelt and Horne (1970), and others is not required here. Katchalsky and Curran, and Haase, have also con— sidered membranes. A general description of transport phenomena in membranes and an extensive list of references may be found in the recent book by Lakshminar— ayanaiah (1969). However, nonequilibrium thermodynamics is frequently applied to the discontinuous fluid-membrane system which is particularly convenient for experimental analysis but is not readily subjected to a theoretical treatment. The system is not regarded as a continuum, and the state variables are not continuous functions of space and time. Here the phenomenological or transport coefficients are functions of the fluid and of the mem- brane. This approach is a "black box" approach in that no questions are asked about the interior of the membrane. Alternatively, if we are able to describe the de- tails of the inside of the membrane, the fluid-membrane system could be regarded as a continuum and the state variables as continuous functions of space and time. In this approach events in a small volume element are examined and the macroscopic properties derived by averaging over the entire volume of the system being studied. Conse- quently, it will be necessary to construct a dynamic model and then correlate the model with the macroscopic phenome- nological description obtained experimentally. The aver- aging process will require the choice of a proper refer- ence frame for the flows. Kobatake and coworkers (1964) and Osterle and coworkers (1968,1970) and Manning (1968) have also used the continuous approach. 10 B. Chemical Potential The driving forces needed to produce a flux or flow of molecular and/or ionic species are the gradients of the chemical potential of all species in the system (e.g., Haase, 1969). In the absence of a magnetic field, the total chemical potential “a of Component a may be written (Horne, 1966) _ O “a — “a (T,p) + RT Quad + Ma? + zaFw (2.1) where am is the activity of Component d, a = X f p (2.2) T is absolute temperature, p is pressure, R is the gas constant, Ma is molecular weight of d, F is the gravita- tional potential, za is the ionic charge per mole of a, F is Faraday's constant, and w is the electrostatic poten- tial, and where u: and the activity coefficient fa are defined relative to an appropriate standard state in which there are no effective external fields. Note that fa is a function of temperature, pressure, and composition. Since we treat only aqueous solutions here, the appro- priate standard state is the pure solvent (denoted by the running index 0), o _ . _ _ _ “a - 11m [pa RT in Xa Mar zaFw] (2.3) x *1 o ll whence lim f = 1. (2.4) a We are primarily concerned with the gradient of the chemical potential. Using the chain rule for dif- ferentiation of “a (T,p,xB,F,¢) we have, in general nota- tion, v-2 Vua = uaTVT + papr + 8:0 uaBVxB + uaFVF + pawvw (2.5) where = = _’ ' t uaT (Bud/8T)p,xa,F,w sa (partial molar en ropy) “up = (Bud/BP)T,XG,F,w = Va (partial molar volume) “GB = (Bud/3X8)T,p,P,w = RT (Bin xa fa/BXB)T,p.F,w “a? = (Bud/3F)T,p,xa,w = Md “aw = (and/aw)T,p,xa,P = zaF (2.6) Thus, v-2 = -_ _ + + . 2.7 vua sa VT + van + 830 “a8 VxB MaVF zaFvw ( ) 12 The Gibbs-Duhem equation including the external forces is (Horne, 1966) v-1 2 xa vua = -E VT + V vp + Mvr + EFvw , (2.8) d=0 where _ v-l _ v-l _ s = Z x s , v = Z x v , a=0 d d d=0 d d _ v-l _ v-l v-l M = Z x M , z = X x z ,0 = Z x u . (2.9) d=0 a a d=0 d d d=0 a QB For the case we shall treat, wherein the gravitational potential is negligible and the temperature is uniform, (2.8) becomes, upon division by V, v-1 2 c Vu = Vp + EFVW (2.10) d=0 a a where the molarity is defined by ca = na/V = xa/v (2.11) and where 2 = z c z = (p/fi)? . (2.12) Note that E and E vanish for electroneutrality. 13 For most of our purposes, (2.7) is not the most useful form for the chemical potential gradient. Instead, taking the gradient of (2.1) directly, we have _ o Vua — Vua (T,p) + RV T in aa + MaVF + zaFvw , (2.13) which becomes, for the case of negligible gravitational potential and uniform temperature, _ —o Vua - Va Vp + RTVQn ad + zaFvw , where o _ _ . _ (Bud/8p)T’P’w — v — 11m v , and where V x RTVRn a = Z d = 8 Note that ---0 Va + RT (Sin fa/Bp) V d T,p,xa C. Hydrodynamic Equations Equation of Continuity of Mass + RT (sen fd/ap)T,p,x (2.14) (2.15) Vp (2.16) a (2.17) In the absence of chemical reactions, for a fluid containing v components the v independent equations of continuity of mass are 14 (dp/dt) + pv-E = o , (2.18) p (dwa/dt) + V-ja = 0 , (2.19) where p is density, t is time, 6 is the center of mass, or barycentric velocity, and Wu and 3a are mass fraction and diffusion flux of Component a, respectively. The barycen— tric velocity 3 is defined by v 1 + _ + u = Z w u , (2.20) d d d=0 where Ed is the velocity of Component a relative to a laboratory frame of reference. The diffusion flux 3a is defined by Era = pa (Ea—E),a= 0,....,\)"l o (2021) Note that pa = wap. The diffusion fluxes, however, are not all independent, v-1 2 a=0 3a = o . (2.22) Substantial time derivatives d/dt are relative to local time derivative B/at by (d/dt) = (a/et) + E-v . (2.23) The operator "V" is defined by 15 v = I (e/ax) + j (a/ay) + E (e/az) , (2.24) + where i, j, and k are the unit vectors of a three dimen- sional Cartesian coordinate system. For more detailed analysis in the use of various coordinate systems, see, for example, Irving and Mullineux (1959). Navier-Stokes Equation The Navier-Stokes equation which relates the velocity of barycentric frame of reference of fluids to the external forces is (Horne, 1966) p (dE/dt) + V[(% n-n')(v-E)] 2v-nsymvfi + OX " VP I (2.25) where symVE is the symmetrical part of the tensor V5, and where n is the coefficient of shear viscosity and n' is the coefficient of bulk viscosity, both taken as noncon- stant. In writing (2.25) we have used the equation of motion, 7* -> p (dE/dt) — v-E = pX (2.26) + + I where G is the stress tensor, given approxrmately by the Newtonian linear phenomenological relation Q++ 2 + 3 + = - [p + (-3- n-n') (V°u)] I + ZHSYmVu . (2-27) 16 + and where pX, the net external force, is pi = - pVF - zpvw (2.28) Introducing (2.23) into (2.25) and rearranging, we obtain 0 (BE/at) = pi - vp + nvzfi + (% n+n')(7(v.fi) —> —> - puV.u , (2.29) where we take n and n' to be constants. If the system considered is at steady state, (2.29) becomes, by setting B/Bt = 0 and rearranging, + 2+ 1 + + + Vp - ox = nV u + (3 n+n') V (V.u) - puV.u . (2.30) However, most fluids except very dense ones are essentially incompressible. For an incompressible fluid, the density p is constant in time and position. Thus according to (2.18). v-E = o . (2.31) The Navier-Stokes equation for an incompressible fluid in a system at steady state is, then, for constant n, vp - pi = nvzfi . (2.32) 17 Energy Transport Equation The general equation of continuity of total energy is (apEi/at) + v-EE = o , (2.33) T where 3E is the total energy and where ET is the total T specific energy, E = E + % u . (2.34) where E is the specific internal energy and u2/2 is the local kinetic energy of the center of mass. Note also that ++ u = u°u (2.35) Although it may be preferable (see Bartelt and Horne, 1970; Gyarmati, 1970; and Ingle, 1971) to obtain the kinetic energy by summing over the kinetic energies of the components, the difference between the two definitions is negligible for the present purpose (see Horne, 1966). The energy transport equation can be expressed as .+ p (dE/dt) + v-jE = E:VE - pE-§ (2.36) where 3E is the internal energy flux not due to bulk flow 18 ¢ +_ + I + + 3E - ouE + u'0 - (I + puzF) w (2.37) T where the electric current I is the sum of the partial + currents 1 a v-l v-1 1 = 2 Id = z pafia (za/MG)F , (2.38) d=0 d=0 and where v—l _ pz==paio (Wu/Ma)zd = Z (2.39) D. Principles of Nonequilibrium Thermodynamics The above treatment of nonequilibrium system is as far as we can proceed from hydrodynamics alone. In order to proceed in the discussion of the general theory of irreversible processes, we could describe the elegant, rational, fundamental approach of Truesdell (1969), Mfiller (1968), Bartelt (1968), Bartelt and Horne (1970), Gyarmati (1970), and Ingle (1971). Alternatively, we could present the conventional, heuristic approach as exemplified by de Groot and Mazur (1962), Fitts (1962), and Haase (1969). For simplicity, we choose the latter course, and we emphasize that the more fundamental ap- proach apparently gives the same results for the simple systems investigated here (Bartelt and Horne, 1970). 19 We need to introduce two fundamental assumptions regarding the system under consideration. The first assumption is: Postulate I: The principle of local state For a system in which irreversible processes are taking place, all thermodynamic functions of state exist for each element of the system. These thermodynamic quantities for the nonequilibrium system are the same functions of the local state variables as the correspon- ding equilibrium thermodynamic quantities. The second assumption is: Postulate II: The assumption of locally linear fluxes + . I The fluxes 3a are linear, homogeneous functions of the forces Ya. That is, = 2 L V. (2.40) The forces are "driving forces" for the fluxes. The phenomenological coefficients de' are independent of the forces. The diagonal coefficients Lad relate conjugate fluxes and forces, while the off-diagonal elements La a! (aia') give rise to cross phenomena which are produced due to interference when twotransport processes take place simultaneously. As in the case of postulate I, postulate II is apparently valid when the system is close to equilibrium. Thus, both postulates apply to systems 20 with small space and time gradients of the local thermo- dynamic variables. Based on both postulates, we now treat the linear phenomenological theory for nonequilibrium systems. The Gibbsian equation for dE is dE = Td§ - pdV + z “a dwa+-dF + dew , (2.41) Applying the chain rule for differentiation of E (T,p,wa,F,w), we obtain by the principle of local state and various thermo- dynamic formulas, dE/dt = ('C'p-pVB) (dT/dt) - (TVB-pVB') (dp/dt) v-2 — — dr d4 4 ago (Ea-Ev_l)(dwa/dt) + E? + ZF 33 (2.42) where C? is specific heat capacity at constant pressure (and external fields),B is thermal expansivity, _-—-1 - = V 3V 3T 2.43 B ( / )p.wa,F.w ( ) 8' is isothermal compressibility, B' s - V—l'(8V/3p) (2.44) T,wa.P.w and Ea is partial specific internal energy, Ea==(3E/3ma) (2.45) T.p.mB,F,w ‘ 21 Note also that WI II M 3 t13| a d (2.46) Application of the chain rule for differentiation to the equation of state p = p (T,p,wa) gives a similar relation, do/dt = - pB(dT/dt) + 08' (dp/dt) 2 v-2 _ _ - 0 ago (Va—Vv_l)(dwa/dt) (2.47) Substitution of (2.18), (2.19), (2.23), (2.42), and (2.47) into (2.36) yields, after rearrangement, + v-2 _ _, p6 (dT/dt) - T8(dp/dt) = ¢ -V°q - z j . (H'-H p l a=0 d where o1 is the entropy source term for bulk flow, I Z + ¢l - (o + pI):Vu , (2.49) 3 is the heat flux -)- f V_l _. q = 3E - 20 1a Ha . (2.50) a: Ha is partial specific enthalpy, and _' _ _ Ha — Ha-+I‘+-M Fm , ——1—)F4 . (2.51) 22 It has been shown (Fitts, 1962; Kirkwood and Craw- ford, 1952; de Groot and Mazur, 1962; Haase, 1969) that the driving forces conjugate to q and jg are Vin T and By postulate II, the linear phenomenological fluxes are v-l -> .— -q = QTT Vin T + QEOQGTVTUG + V-l _ -ja = QGT Vin T + Z 9&8 VTUB’ d=0,...,v-l (2.53) B=0 where Q's are the phenomenological, or Onsager coefficients. These coefficients are not all independent since, by (2.22), v—1 2 Q = 0, B = 0,...,v-l (2.54) d=0 dB Further, due to the requirement of positive definite entropy production, it has been shown (Bartelt and Horne, 1969) that 2 Q = 0, a = 0,...,v-l (2.55) 23 Thus, for the v independent fluxes 3T = a, 30' 31"°"3v-2’ the linear phenomenological equations are ‘36 = QdT Vin T + 820 ”as VT(“8‘“v-1) ' a = T, o, l,...,v-2 . (2.56) For most purposes, it is more convenient to use (2.53), which is permissible because of (2.55). If the fluxes and forces of (2.56) satisfy Onsager's (1931) condition, which seems very likely but cannot be shown theoretically (Cole— man and Truesdell, 1960; Andrews, 1967), then the matrix of Onsager coefficients is symmetric; i.e., QdB = QBG for 8,8 = T, 0, l,...,v-2. However, since one of the goals of experimental studies of transport phenomena is the verification of Onsager's Reciprocal Relations (Miller, 1960), we do not use them. Substitution of (2.7), and (2.52) into (2.53) yields + v—l _ -Ja = QdT VinT + 320 908 (VB/MB) Vp v-l v-2 + Z 0 2 (u /M ) VX 8:0 as y=0 BY 8 Y v-l + Z (ZS/MB) FVw, d=T, 0, l,...,v-l (2.57) B=0 “as 24 General discussion about the physical implication of various phenomenological coefficients can be found elsewhere (Fitts, 1962; de Groot and Mazur, 1962; Haase, 1969; Horne, 1966). E. The Differential Equations Describing the Osmotic Flow 9f Fluids Through a Charged Membrane The System and the Simplifying Assumptions We are mainly concerned with the derivation of the steady-state phenomenological theory for the trans- port of fluids undergoing osmosis through biological mem- branes. As has been mentioned in Chapter I, the trans- port system for the study is composed of a charged, continuous membrane separating two electrolyte solutions of different concentration at the same temperature. Before we can start to write the equations for our transport system, we must take full account of the fundamental problems which are pertinent to our present investigation. As has been recognized for years, a normally grown living thing is, from thermodynamic point of view, an Open system in a stationary state (the flows of all the species are constant in time), inside of which irreversible processes occur continuously and slowly (Haase, 1969). If, moreover, the fluxes and the gradients are small, we may use the principles of nonequilibrium 25 thermodynamics such as those which enable us to use equilib- rium properties (postulate I) and linear phenomenological relations for the fluxes (postulate II). Moreover, for the system in question we consider (a) an isothermal, incompressible fluid; (b) a capillary model for the membrane in which uniform cylindrical capillaries penetrate across the membrane with flows of all components in the direction of the capillary or axial axis; (c) absence of chemical reactions; (d) absence of external magnetic, gravitational, and cen- trifugal fields; (e) constant viscosities; and (f) slow motion, i.e., a quasi-steady—state in which local time derivatives are zero. These requirements can be realized experimentally, and do not, in principle, introduce error. It is advantageous to introduce them into the phenome- nological theory, the net effect being a simplification of the differential equations. We do not, however, make use of the previous as- sumptions such as (i) constant diffusion coefficients and ionic mobility in the flow equations; (ii) constant dielectric constant of water in Poisson's equation; (iii) radial-independent pressure; (iv) the classical Boltzmann distribution of ions; and (v) the linearized Debye-Hfickel theory for "all" cases. Moreover, (vi), the assumption that Ka>>l, where K is the reciprocal of the effective 26 thickness of electric double layer (Debye length) and a is the radius of the capillary, is found here to be not always necessary. In the next several chapters we shall discuss these assumptions in more detail. Our main pur- pose is to derive, subject only to a minimum number of assumptions, the steady-state phenomenological theory that describes for the transport of fluids across bio- logical membranes. The Differential Equations and the Appropriate Boundary Conditions We now can set up the differential equations necessary for describing the transport system in question. Making use of the principles of hydrodynamics and non- equilibrium thermodynamics and subject to the conditions considered above, the relevant transport equations, rela- tive to the local center of mass are + v-l _ v-l v-2 -3 — 2 a (v /M ) vp + x Q 2 (u /M ) VX d B=0 B B B B=O d8 Y=0 BY 8 Y v-l + = — 8:0 0&8 (zB/MB)FVw, a T, 0, l,...,v l (2.58) + + _ pu Vwa + V 3a —0 (2.59) v-2 + _ _ .+ _ 1 . _'_—l -TBu-Vp — ¢l V q ago 3a V (Ho Hv-l) . (2.60) Vp + erw = nvzfi . (2.61) 27 Only (2.58) and (2.61) are considered in what follows; the other equations deal with effects too small to be observed in membrane transport in ordinary electrolyte solutions. The partial electric current relative to the local center of mass is e _ e la — (Zn/Md) ja , (2.62) and the total current is given by (2.38). Poisson's equation, which relates the electro- static potential w to the excess charge sz is V°DVw = -4nsz = ~4WEF , (2.63) where D is the dielectric tensor of the electrolyte solu- tion, taken as nonconstant, and z and 7 are given by (2.41) and (2.12). The units of w are volts throughout this work. At this point our description of the system under considera— tion is, in principle, complete. In problems of electric conduction, however, it is useful to relate the diffusion fluxes and electric currents of all the species to the velocity of the solvent and thus to adopt.the Hittorf reference system (Haase, 1969). The solvent-fixed (SF) frame (i.e., (jo)o E 0) is extremely convenient and transference numbers are usually referred to it (Miller, 1966). The reference velocity is now + denoted by the velocity of solvent molecules, uo. 28 Moreover, since the solutions are relatively dilute, it is useful to use molarity ca instead of mole fraction xa to describe composition. Before we can write explicitly the transport equa- tions for the SF reference frame, it is essential to re- quire considerable knowledge about conversion from a reference velocity Ea to a second velocity 5b. For simplicity, we follow the treatment of Haase (1969). We consider, in the general case, the transition from one diffusion current density + —> ana = ca (um-ma) . (2.64) to a second diffusion current density -> + bja = ca (ua-wb) . (2.65) ZFor fluid systems, the reference velocity Ea and 3b are so chosen that the relations + V—l —> ma = :o(wa)aua (2.66) v-1 2 (w ) = l (2.67) d a a: v-l + + wb = Z (wa)bua (2.68) 29 Z (w )b = l (2.69) d=0 hold. The (wa)a and (wa)b are the normalized weights, i.e., "weight factors" for averaging the velocities sub- ject to the normalization (2.67) and (2.69). For the barycentric velocity, defined in (2.20), the "weight factors" are the mass fractions. It can easily be shown from (2.66), (2.67), and (2.64) that v-1 2 [(wa)a/Cal d=0 T J = 0 . (2.70) 30. and from (2.68), (2.69), and (2.65) that v-l —? — Z [(wa)b/ca] bja — 0 . (2.71) d=0 Combination of (2.64) - (2.71) yields the conversion re- lation between two frames of reference: v 1 -c 2 a = B [(wB)b/cB] aje . (2.72) 0 We are, however, primarily concerned with the con- VeIflSion from the barycentric system to Hittorf's reference System. Let = 3 /Ma. baa = (3') (2.73) 30 Note that in Hittorf's frame of reference, (we)o = l, (Wm)O = 0, or = 1,2,...,v-l , (2.74) v-l so that (2.69) is immediately satisfied, i.e., Z (wa)O = 1, d=0 and + _ (30)O — 0 (2.75) Hence we obtain from (2.74) and (2.72) (3 ) = 3 /M - c I /c M , (2.76) yielding the conversion from the barycentric system to Hittorf's reference system. By definition of jo, the two reference velocities are related by +_—>- 7 u — uo (l/Oo)jo . (2.77) Consequently, if SF reference frame is chosen for the system, the diffusion flux and electric current of species a can now be eXpressed, relative to SF reference frame, as -r V_l o —o \)-1 o -(3 ) = 2 9 v Vp + Z 9 RTVin a d o B=l d8 8 B=l d8 8 v-1 0 + E 9 z Fvw,cx= 1,2,...,v-l (2.78) B=l d8 8 31 ) (2.79) where 038 are the phenomenological coefficients in SF reference frame. The Navier-Stokes equation in SF refer- ence frame then becomes, according to (2.77), Vp + EFVw = nVZ E - nV2 ELI; o o (2.80) 0 We assume, for the Navier-Stokes equation only, that the difference between E and Go is negligible and, therefore, that _ 2+ Vp + ZFVw = nV uO (2.81) The appropriate boundary conditions for the problem are c := c (0) at x = 0, and c = c (i) at x = i do do do do PO(X=i)-po(x=0)=0 0 = 0 (0) at x = 0, and 0 = 0 (i) at x = i , (2.82) where Coo (0) and Cdo (i) are initial molar concentrations of Component d at the boundaries, pO (x = i) and p0 (x = 0) are initial hydrostatic pressures at the boundaries, 0 is partial electric potential along the axial axis, and 0 (0) and 0 (i) are initial values of 4 at the boundaires. The appropriate boundary conditions to w and 110 are dw/dr Sue/8r = 0 dw/dr 4no/Da, and u = 0 at r = a , (2.83) O 32 where c is the surface charge density fixed on the mem- brane, Da is the dielectric constant of the solution at the wall, and uo is the velocity of solvent molecules or the SF reference velocity along the axial axis. The experimental transport parameters, for some binary electrolyte solutions, which are necessary for the study have been obtained based on the SF frame of refer- ence and can be found elsewhere (Miller, 1966). CHAPTER III EQUILIBRIUM DISTRIBUTION OF IONS IN AN ELECTRIC FIELD A. Introduction Before we can obtain the solution of the steady- state phenomenological theory which describes osmotic flow phenomena in biological systems, it is necessary to acquire considerably more knowledge of the local distribution of ions, charge density, electrostatic potential, and pres- sure in nonequilibrium system. This information is essen— tial in describing transport phenomena in membranes and thus must be resolved and discussed beforehand. In the present chapter we derive an equation for the local dis- tribution of ions in an electric field and then obtain the excess charge density in the system considered. The principal enabling step is the assumption that the system is in chemical equilibrium radially, i.e., that the chemical potential of each species is uniform radially. The fundamental problem in a study of a membrane system or any electrochemical system in which the charges are unequally distributed (i.e., 2 ¢ 0) in the volume or 33 34 on the surface is the knowledge of the macroscopic dis- tribution of ions at each point of the system since the presence of a net charge produces a macroscopic electric field which tends to displace charged species. From knowledge of this distribution it will be possible to calculate several electrochemical properties and thus to describe accurately the behavior of electrolyte solutions in electric fields. However, as has been known for years, there are great difficulties in calculating the distribution of ions in the field, due partially to the lack of a coherent theoretical formulation, and partially to the mathematical difficulty in solving the problem. For these reasons one often uses the classical Boltzmann distribution equation without adequate justification (Sf. Kobatake and Fujita, 1964; Fujita and Kobatake, 1968; Gross and Osterle, 1968). Thus, it is assumed for each species a that _— __ _ B xa — Xa eXp ( zaFw/RT) — xa , (3.1) where Ea is the mole fraction of d in the absence of an electric field and where, for subsequent use, we denote the Boltzmann equation mole fraction by x2. Many assumptions have been made in conjunction with (3.1); the usual ones are: (a) existence of point charges, which implies that molar volumes of ions are 35 zero; (b) absence of polarizable components; (c) absence of activity coefficients; (d) uniformity of pressure, which implies among other things that the pressure- induced gradient of the electrochemical potential of a charged species is negligibly small. Despite the fact that various authors have attempted to correct the Boltz- mann equation, often empirically, some by introducing a finite volume (Sparnaay, 1958), some by introducing polarization, and others by introducing pressure, none of the theories is complete. Moreover, many of these authors treat the parts of the double layer very close to the interface as special media, very different from the parts remote from the interface. Further, the micro- sc0pic nature of these parts of the layer give rise to difficulties in their description in terms of macroscopic quantities such as polarization or pressure. The most complete theory for the distribution of ions in an electric field so far has been that of Sanfeld (1968) derived by means of the local thermodynamic method (Prigogine, 1953; Mazur, 1953; Defay, 1954). Unfortunately, the theory is limited only to equilibrium systems and is valid only for very dilute solutions and gas mixtures. Our purpose now is to calculate this distribution in the general case and to apply our result for calcula— tion of the excess charge density in the system. 36 We consider continuous, anisotrOpic fluids in which no chemical reactions occur and which are subject only to the external electric field. Unlike many other investiga— tors, we make full use of the equations of chemical thermo- dynamics, including particularly all pressure terms. At present, we only treat systems (particularly, biological systems) in which the surface charge density fixed on the surface or on the membrane phase is such that the effects of polarization are small enough to be neglected. This requirement can always be realized experimentally without introducing error, but it could be modified in subsequent extensions of our theory. B. Derivation of Modified Boltzmann Equation Consider a charged, continuous phase (i.e., a membrane, a barrier, or a cylinder) which is static or through which there are steady-state flows of matter. We assume an isothermal electrolyte solution containing solutes which are dissociated in a neutral solvent, absence of chemical reactions and external fields except electric field, and a capillary model for the phase con- sidered in which uniform cylindrical capillaries penetrate across the phase. Our results also apply to any single, non-capillary cylinder. 37 When the only external force is electrical, the gradient of the molar electrochemical potential “a of Component a in the absence of polarization may be described, according to (2.14) by —O Vua = Va Vp + RTVin ded + zaFVw, d = 0,1,...,v-l (3.2) where V: is defined by (2.15). The gradient of “a in the radial direction is, by (3.2), _ —o Spa/8r — Va (Sp/3r) + RT (Bin xafa/ar) + zaF (aw/3r) . (3.3) ‘We assume that the chemical potential of each species is uniform radially, Qua/3r = 0, d = 0,1,...,v-l . (3.4) 'That is to say, all the radial fluxes are zero everywhere in the system; chemical equilibrium obtains radially. lqote, however, that the gradient of “a along the capillary > 3, w + 3% V + ... (4.9) 56 For better than 1% accuracy, we therefore require 0.01 >> 3% 92 + §%-w4 + ..., (4.10) which yields IVI < 0.245 (4.11) So, for Ital < 0.245 (in fact, for [Val < 0.49), the Chapman Gouy equation becomes 1/2 Iol == (2CERT) IT | , (4.12) a and (4.11) yields the following restriction on lol//E , 2 (|o|//E) < 0.245 (28RT)1/ (4.13) For T = 298°K, R = 8.314 J mole-l K'l, and e = 7 x 10‘10 C V-1 m-l, (28RT)l/2 = 1.86 x 10"3 c mole‘l/2 m'l/2 (4.14) Therefore, for sinh [Val to equal IVaI to better than 1% accuracy, it is necessary for 3 1/2 m‘l/2 (4.15) (|o|//E) < 0.45 x 10‘ c mole- Concentrations of important metal ions in biological systems (Woodbury et al.. 1970) lie in the range of 0.004 < c < 0.4, in units of moles per liter, while ion concentra- tions are often outside this range in other systems of 57 interest. We take c i 2 mole liter"l = 2 X 10-3 mole cm-3 = 2 x 103 mole m-3 since the validity of the simple linear correction to the dielectric constant, (4.4), would be questionable for higher concentrations. Then in order for both (4.15) and (4.16) to hold, Iol may be as large as 3 C m-Z. On the other hand, for c = 0.004 mole 6 mole cm-3 = 4 mole m-3 be larger than 0.9 x 10"3 C m_2. Typically, charge den- 3 20 x 10- 1 liter- = 4 X 10- , Iol must not sities are of the order of 10— C m_2 (Fair and Osterle, 1971). However, for I0] 2 10-3 C m-Z, we require c 2 4 x 10.3 M. Otherwise, for fixed lo] 2 10.3 C m-Z, smaller concentrations lead to values of IV] too large for (4.7) to hold. For such small concentrations (which do not appear to be of biological interest in any case), there exists no analytical solution of Poisson's equation. Fair and Osterle (1971) have obtained numerical solutions 5 -3 2 for c z 10- M and Icl : 10 C m- . We hereafter restrict consideration to 0.004 M 5 c 5 2 M (4.16) 3 2 and Iol z 10- C m- , the actual upper bound on Io] being that given by (4.15) for the actual composition in question. Working backwards, we now accomplish item (iii) by using (4.3), IEI z(dV/dr)a = IoI/ea (4.17) 58 For lo] 5 20 x 10'3 c m“2 and e z 7 x 10"10 a 7 c v"1 m‘l, v m‘l, so that IhlEl2| 5 3 x 10‘16 x 9 x 1014 [El 5 3 x 10 = 0.27, which is less than 1% of the dielectric constant D0 of water. Since this is the maximum correction to D, we confidently neglect the contribution of IE]2 to the dielectric constant and write D=Do+0c+6c (4.18) By Table 4.1, the maximum contribution of the concentration terms to the dielectric constant would be for c+ = c_ = 2 M = 2 X 10-3 mole cm-3; for, say, 2 M LiI, D = 78.3 - 22 - 14 = 42.3, a 46% reduction. The concentration correc- tion term contributes less than 1% for total salt concen- trations less than about 0.05 M, depending on the particu- lar ions involved. The next task is estimation of xa and ca. For total salt molarity 5 2 M, Ea is at most 2 M and — — —0 ~ - —o 3 -l — x = c v ~ c v = 18 cm mole c , O. 0!. 0. C1 0 O. E 5 0.002 mole cm‘3, Pa 5 0.036, i0 3 0.928 (4.19) d For uni-univalent salts, by (3.35), B _ x0 = l - (1-xo) coshV , (4.20) and for (1—ib) 3 0.072, [9| 5 0.245, 59 x: 1 1 - 0.072 x 1.03 = 0.926, F — x3 = (1-I ) (coshW-l) 5 0.00216 0 O O (4.21) Thus, correction terms proportional to (xO-xg) in (3.22) - (3.25) are entirely negligible compared to unity. Correction terms due to volume terms, however, are not always negligible. For, say, 2 M KCl, by Table 4.1, W (0.036/0.928) e" 0.063 . R Thus,for better than 1% accuracy, our equations from Chapter III become, with v—l __ — — _ -o —o W's—8:1 (xB/xo) [eXp ( zB‘FHWB/Vo) . the following: xa = xa exp (-zaV) (1/6) = (1—V*)/§5 VS _ B — -o _-* = — _ ca — (Xa/XOVO) (1 v ) ca exp ( zaV) (8.8/l8) + (0.036/0.928)ew(18/18) (4.22) (4.23) (4.24) (4.25) (4.26) 60 0—1 —_ ——O _ - -— Z — (l/xovo) ( 2 xa za exp ( zaV)) (l v*) d=l v-l _ = ail ca za exp (-zaV) (4.27) In obtaining (4.25) — (4.27), we neglect the difference between v* and V* compared with unity since v-l _ - — —o —o v-l _ _ —o —o = v* + 8:1 (xB/xo) (l-exp (-zBV))(vB/VO) (4.28) and, for small IV] and a uni-univalent mixture, —* _ * ~ ‘ — “0 ’0 v v ~ ( 8 (XS/X0) (VB/V0) cations - z (E /i) (V°/V°)) (4.29) anions B O B O For 2 M KC1, e.g., with IT] = 0.245, 7* - v* 2 0.02 X 0.245 = 0.005 (4.30) Thus, for the prescribed conditions the equations of Chapter III reduce to the simple Boltzmann equations. If we further specialize to the uni-univalent case, we have 61 7 = 2 E e—w - Z '5 ew cations a anions = - 2c sinh? = - ZCV , (4.31) where we have used (4.7) and where c = z E = 2: E (4.32) C a O cations anions Moreover, the equation for the dielectric constant is D + 2 6 E e + z 6 E e (4.33) 0 a a! I cations anions D and Poisson's equation is l. d D dV _ 2 _ 2 EEE r (5;) a? - (8TTCF /DORT) W — K W , (4.34) where K = F (8nc/DORT)l/2 (4.35) is usually called the reciprocal of the Debye length. The boundary conditions on (4.34) are (d‘Y/dr)r=0 = 0, (dV/dr)r=a = (4noF/RTDa) , (4.36) where —V V — a — a Da = D + 2 6a ca e + Z 6 ca e (4.37) cations anions 62 D. Perturbation Scheme The concentration dependence of the dielectric constant must be taken into account since it strongly affects the dielectric constant. However, it complicates the Poisson equation, (4.34), so much that it still cannot be solved directly. In order to take account of this last complication, we solve (4.34) by a perturbation scheme defined by 6 = DO + a (5+ c+ + 5_ c_) (4.38) and @ = W + E? + 62? + O (63) , (4.39) o l 2 When a = l, D = D and. Q = W. The parameter a is simply a bookkeeping device which allows us to keep track of the inclusion of the concentration dependence of D. In order to obtain an explicit perturbation expression for c+ and c_, we note that 2 e — l i W0 i 6W1 + O (a ) l 2 + 7 W0 + ewowl + ... (4.40) and (4.33), (4.38), (4.40) yield A _ l 2 D = D + e{ Z 6 c [l—W + — W —'€W (l-W ) + ...] O cations a O 2 O l O + Z 65[1+wo+%wg+ewl(l+4’)+ ]} anions a a O = D0 + e{ ( z a E + 2: 5 Ea) (l + % 42) cations a anions a O — < 2 5a Ea - 2: 5 E ) we} cations anions a + o (52) = D + eD + O (82) (4 41) o 1 ’ ° where D1 = ( 2 6 E + X 6 E ) cosh? . a a . d a o cations anions - ( Z 6 E' - Z 6 E ) sinh? . a . a o cations anions + - . = D cosh? - D SinhW (4.42) o o Substitution of (4.41) and (4.39) into (4.34) yields (l/r) d/dr {r (1 + eDl/DO + o (52)) (dwO/dr + edwl/dr + O (52))} = K2 (40 + 54 + o (32)) , (4.43) l 64 with boundary conditions, from (4.36), (4.39), and (4.41), 2 (d‘Po/dr)r=0 + (d‘l’l/dr)r=0 + O (8 ) — O , (4.44) 2 (dwO/dr)r=a + e (dwl/dr)r=a + o (e ) l = (4noF/RT) (00 + eDl (a) + o (52))" (4.45) The zeroth order differential equation is, from (4.43), (l/r) (dr (dWO/dr)/dr) = K2 we (4.46) with boundary conditions (d‘PO/dr)r=0 = 0, (d‘PO/dr)r=a = 4woF/DORT (4.47) The first order differential equation is, from (4.43), (l/r) (dr (dWl/dr)/dr) = - (l/r)(dr(D1/DO)(dwo/dr)/dr) + K2 41 , (4.48) and the boundary conditions are (d‘l’l/dr)r___0 = O, (d‘Pl/dr)r=a = - 4noFDl(a)/D: RT (4.49) Clearly, higher order equations could be obtained if de- sired. 65 E. Solutions of the Differential Equations Then and Substitution of Variables First, make the substitution -1 g = Kr r = EK ga = Ka , d/dr = K (d/dg) r The zeroth order equations are then or (1/5) (d5 (dwO/da)/dg) = w O dsz/de2 + (1/5) (dwO/dg) - w = o , O (Mo/dag:O = 0, (dWO/d§)€= = 4woF/KDORT E a The first order equations are 2 d Wl/dg q1 (E) 2 — (1/5) (da
> 1. Recalling that the motivation for the requirement (4.15) was that we (Ka) , (4.64) 68 wished to use (4.7), we may easily modify (4.15) in light of our results here. Formally, all we need is (|o|//E i 0.45><1o‘3/[10(Ka)/Il(na)] c m"1/2 mole-l/Z (4.68) where the denominator is tabulated in Table 4.2. For example, for Ka = 0.5, it is necessary that (Iol//E) : 0.11 X 10"3 C ml/2 mole-l/z; for c = 2 M, the requirement becomes Iol < 5 C m ; and so forth. ~ First Order Solution By (4.42) and (4.7), as supported by (4.68), D1 = D - D WC (4.69) Then by (4.57), <31 (6) = — D31 {[D+ we - D- W2] - 0' W52} + = - (AD DO) IO (g) 2 "’ 2 2 4 + (A D /Do){[IO(€)] + [11(6)] } ( .70) Now the left-hand-side of (4.56) is the same as the left-hand—side of (4.54), and we therefore know that usoo . n 5 .H m>uso .ny mo cowuocsm 0 mm 5 “mumamumm mmmacowmcmfiflp map mo poamllm.¢ .mflm +.Hv_ om H mvm.o o 1 1 I l. 1 1“ ‘ i d 4 1 1 q u q ‘ I 55. 5 p _ _ mmm.o m mam.H mmma.o mvmo.o II ll (0 .Q 79 According to our analysis, we notice that W - To is nonpositive. This indicates that the inclusion of the concentration dependence of the dielectric constant in Poisson's equation actually results in the decrease of the overall potential function at any point in the system. However, the decrement W — WC is very small. As a result, we can anticipate that the solution represented by 90 + 91, where W is the first-order solution, should give an 1 excellent approximation to V. For larger values of Ka, more complicated numerical analysis is required, but there is no conceptual difficulty. For smaller values of c, the contribution of the concentration dependence of D will of course be less, and 41 will be less important. G. Conclusion Including the radial dependence of the dielectric constant in Poisson's equation, we have obtained analytically, in the general case, an improved potential distribution equa— tion for an uni-univalent electrolyte system, subject to 3 to 2 mole/2. (4.68) and for salt concentrations of ~ 10- The equation can be used to study various biological sys- tems. A practical potential distribution equation is also derived for the case of Ka < 2 where a = 4.2 g. Conse- quently, the physico-chemical properties of aqueous uni- univalent electrolyte solutions in biological systems and other charged systems can now be interpreted more accurately. 80 In the derivation of the equation, we have required sinh W = W. Recently, MacGillivray gt_§l. (1966) have demonstrated that the solution of the Poisson-Boltzmann equation for a charged cylinder can be approximated by the solution of the Debye—Hfickel equation even when the molarity of the electrolyte is low, i.e., when T >> 1, provided the cylinder is "moderately charged," that is, when ZnIOIF a/DORT < 1. For the example mentioned in the preceding section, it can be shown that 2NIO|F a/DORT = 0.0118 at 25°C. We have here obtained explicit conditions (4.68) on the ratio of IOI to /E and have thereby eliminated the need for restriction on K itself. In fact, we have not used Debye-Hfickel theory as such at all. CHAPTER V THE STEADY-STATE DISTRIBUTION OF PRESSURE IN MODERATELY CHARGED SYSTEMS A. Introduction In the study of electrochemical systems, knowledge of the pressure distribution has been realized to be essen- tial in the understanding of interfacial diffuse double layers and thin liquid films (Verwey, 1948; Derjaguin, 1966; Herwitz, 1964; Defay, 1963; Sanfeld, 1966). The difficulty in obtaining a simple equation for the pressure distribution is primarily due to the lack of a theoretical formulation for the problem. As a result, many authors have either neglected its existence or considered the pressure to be independent of electric field (9:. Kobatake and Fujita, 1964) without justification. As we shall see in the next chapter, the effect of the pressure is impor- tant and vital to the understanding of osmotic flow phenomena of fluids across a charged, continuous phase; i.e., a membrane. Previously, no explicit, practical pressure distribution equation has been obtained, and its various important applications have not been dis- cussed. 81 82 In this chapter, our primary purpose is to make use of the potential distribution equation derived in Chapter IV and to derive, in the general case, an explicit formula for the pressure distribution at any point in a moderately charged system where the conditions of Chapter IV are satisfied. We consider a nonequilibrium system at. steady state, an isotropic fluid, and absence of electric polarization. A capillary model is adOpted for the mem— brane. We are primarily concerned with the uni-univalent case. One of the fundamental approaches in the study of flow phenomena in membranes, charged or uncharged, is the knowledge of macroscopic osmotic pressure in the system. The differential equation that governs the pressure of fluid at any point in the system is, in the general case, by (3.3) r dp = - (RT/6:) d in X0 , (5.1) where p is the pressure at any point in the system, x0 is the mole fraction of water at any point in the system, w . I V0 is the molar volume of pure water, R is gas constant, T is the uniform, absolute temperature, and 9n fO is assumed constant. From (5.1) we immediately obtain p = - (RT/52) in X0 + B , (5.2) 83 where B is the constant of integration. This equation can be used in a variety of ways. For example, the difference in pressure between any two points a and b in a system is the osmotic pressure "ha: “ha = p (b) - p (a) = - (RT/7:) 2n [xo(b)/xo(a)1 (5.3) By using this equation and the results of Chapters III and IV, we could calculate n a between any two points in the b membrane. We are, however, primarily concerned with osmotic flow through the charged membrane. We hereby define “mo as the steady-state value of the macroscopic osmotic pressure between the membrane phase m and the aqueous phase in compartment 0 (the other compartment will be called compartment L), “mo = p (m) - pO (0) . (5.4) where p (m) is the average pressure in the capillary and po (0) is the average pressure in compartment 0. More- over, define 020 as the steady-state value of the osmotic pressure across the charged membrane. 0&0 = p (1) - p (O) , (5.5) where p (2) and p (0) are the average pressures over the cross section of the capillary at, respectively, 2 and 0, 84 the positions of the interfaces of the membrane with, respectively, compartments L and 0. From (5.2), (5.4), and (5.5), there follows (rmo = - (RT/R73) 2n [Seem/Xe (0)] (5.6) 020 = — (RT/(7'2) SLn (502/3300) , (5.7) where xom is the steady-state value of the average water concentration in the capillary, x0 (0) is the average value of water concentration in compartment 0, and £02 and i are the steady—state values of the average water con- 00 centrations over the cross section of the capillary at, respectively, 2 and O. In order to obtain the relative distribution of water, i.e., [Eom/Xo(0)] or (Egg/£00) at steady state, we could use directly the results of Chapters III and IV to obtain the desired averages. It is more illuminating, however, and somewhat simpler to focus attention on the pressure itself. Hence, our second goal in this chapter is to make use of the pressure distribution equation and to derive the practical equations for "mo and n as lo functions of the salt concentrations and the electric field for use in the study of transport phenomena of water in charged systems. Consider a charged, continuous phase separating two aqueous electrolyte solutions of different concentration 85 at the same temperature. We wish to derive, in the general case, the pressure distribution equation for the system. By (3.7) I 8p/8r = - EF (aw/er) . (5.8) By (4.27) (the conditions which make it valid are there— fore required here), v-l _ (Bp/ar) = - (RT) 2 c 2 [exp (—z 9)](89/8r) (5.9) a a a a=1 This is easily integrated to yield v-l _ P = P + RT 2 C [eXp (-z W) - l] , (5.10) a a 0:1 where po is the value of p for zero electric field. For a uni-univalent system, (5.10) becomes p = pO + 2cRT (cosh? - l) , (5.11) where we have used (4.32). We have now obtained explicitly the pressure distribution equations which take account of the concentration dependence of the pressure. Since osmotic flow is our concern here, we are interested in the longitudinal, or axial, dependence of the pressure. Differentiation of (5.11) yields 86 (Sp/8x) = (ape/3x) + 2RT (cosh? - l) (ac/8x) - RTE (aw/8x) , (5.12) where we have also used (4.31). If we now assume that the last term on the right-hand-side is negligible, we have (Sp/3x) = (apo/ax) + 2RT (coshW - l) (ac/3x) (5.13) This assumption is reasonable on two counts: (i) we do not expect the potential difference across the membrane to be significant, and (ii) the average net charge in the solution inside the membrane must be small. In any case, we are interested here in illuminating the contribution of the concentration gradient to the pressure gradient, and we therefore use (5.13), wherein W is taken to be the value of Fw/RT for E, the average salt concentration in the membrane. B. The Macroscopic Osmotic Pressure Between the Mem- brane Phase and the Aqueous Phase in a Moderately Charged System Making use of the pressure distribution equation derived above, in the following treatment we derive an explicit formula to represent the macroscopic osmotic pressure "mo between the membrane phase m and the aqueous phase 0 for a moderately charged membrane, separating two uni-univalent electrolyte solutions. 87 We introduce (5.11) into the following equation 5 (r) = g pdx/(z-O) , (5.14) I 0 where E (r) is the average axial pressure, and where 2 is the length of the capillary or the thickness of the mem— brane. Then, there results 2 '5 (r) = p0 dx/(l—O) + 2RT {cosh‘l’ - l} (:cdx/(i—O) (5.15) I 0 where we again, and for the same reasons, take W to be its average in the membrane. Clearly, the evaluation of integrals in (5.15) re- quires knowledge of the explicit expressions of p0 and c as functions of x. We assume, as a first approximation, that po (x) and c (x) are linear functions of x. That is, we may write p0 = (APO/2) X + pO (0) (5.16) c = (Ac/2) X + c (0) , (5.17) where ApO and Ac are defined as Apo = po (2) - po (0) (5.18) Ac = c (12,) - c (0) (5.19) Combination of (5.16), (5.17) and (5.15) yields the ex- plicit expression for the radial pressure in the system, 88 E (r) = {pO (2) + pO (0)}/2 + RT (coshw-1}{c(2)+c(o)} (5.20) We rewrite (5.26) as O '5 (r) - p (0) = Ape/2 + RT {cosh‘P-l}{c()2.)+c(0)} (5.21) The macroscopic osmotic pressure “mo between the membrane phase m and the aqueous phase 0 can be deduced from (5.21), yielding the following equation "mo = ApO/Z + RT{c (2) a a + c(O)}[I 28r{coshW-l}dr/ I 20rdr] , (5.22) 0 0 where Wmo is defined by (5.4) in which p (m) is given as a _ a I anp (r) dr/ 6 andr . (5.23) p (m) 0 In the absence of initial hydrostatic pressure difference, i.e., when ApO = 0, then,(5.23) becomes "mo = RT {c(£)+c(0)} 6a an {coshy—l} dr/Tra2 . (5.24) For experimental convenience, in general, the salt con- centration at one side of the membrane is fixed while that at the other side of the membrane is varied. Thus, in practice we may rewrite (5.24) as "mo = c(£)RT(l+q) 6a 2r (cosh‘i’-l)dr/a2 (5.25) where q is given as q = c(0)/c(2) . (5.26) 89 The value of q can then be varied during the experiment. To evaluate the integral in (5.26), we make use of the conditions of Chapter IV. There follows, then, a a 2 I 2r (coshW-l) dr 5 0 r? dr (5.27) 0 where we have used coshw = 1 + 92/2 + 5% 44 + ... ~ 1 + % 92 (5.28) Introducing (4.50) and (4.52) into (5.27), we write 5a 6a :92 dr = (1/K2) I £72 66 , (5.29) 0 where K is the reciprocal of Debye length defined by (4.35), and hence (5.25) becomes 5 «an 5 {c (0) RT (1 + q)/€:} (1)3 642 as; . (5.30) Making use of the substitution: W = To + 91, where WC and 91 are defined by (4.65) and (4.96) respec- tively, we have Ea 2 ga 2 5a I 69 6: = I 59 as + 2 I 69 w 66 O O o 5 O l Ea 2 5 31) + 6 5W1 d6 . ( . For simplicity, denote Ea 2 L1 = I SW 66 (5.32) 0 O 90 ga L2 = a 5W0 W1 d5 (5.33) Ea 2 L3 = a £91 dE (5.34) The evaluation of (5.32) can easily be carried out. We introduce (4.65) into (5.32) and make use of (4.88). There follows 2 2 O (Ea) - Il (Ea)}/2 . (5.35) However, from (4.96) and (4.98) we notice that there is great difficulty in evaluating (5.33) and (5.34) in closed forms, and hence the evaluation of such integrals requires knowledge of the physical situation. Hence, (5.30) be- come S “mo 5 {c (0) RT (1 + q)/€:} [A2 a: (I: (6a) 2 - 11 (6a)}/2 + 2L2 + L3] . (5.36) The equation can, moreover, be used to determine the relative distribution of water between the membrane phase m and the aqueous phase 0, since by (5.6), i /x0 (0) = exp (- 8 62/81) (5.37) om mo 91 C. The Macroscopic Osmotic Pressure Across a Moder- ately Charged Membrane In this section, we derive a formula to represent the macroscopic osmotic pressure 0 across a moderately 20 charged membrane. Returning to (5.13), we integrate the equation over the length of the capillary, i.e., 0 i x i 8. There results Ap (r) = ApO + 2RT {cosh W-l} Ac (5.38) where Apo and Ac are defined by (5.18) and (5.19), re- spectively. The macroscopic osmotic pressure across a charged membrane can, then, be deduced from (5.38), yielding _ a 2 020 — Apo + 2RTAc 6 2r {coshW l} dr/a (5.39) In the absence of initial hydrostatic pressure difference, i.e., when Apo = 0, (5.39) becomes a 6,0 = 2RTAc I 2r (coshW—l) dr/a2 (5.40) 0 From (5.27), (5.29), (5.31), (5.35), and (5.26), there follows «,0 = 2 {c(4) RT (l-q)/€:} (A2 6: {IO (€)2 — I (ga)2}/2 + 2L 1 + L3] . (5.41) 2 92 This equation can be also used to determine the relative distribution of water at the interfaces with the two aqueous phases, by (5.7): xoi/xoo = exp (- n —O lovo/RT) (5.42) D. Osmotic Flow of Water in a Charged Membrane As has been known for years, knowledge of the macrosc0pic osmotic pressure is useful in the understand- ing of the mechanism of transport process of solutes and water across a charged membrane. In the following treat- ment we make use of (5.41) to study the osmotic flow phenomena of water across a charged membrane. As an example, we consider a charged membrane for which the fixed surface charge density is 10"3 C/m2 and the radius of the capillary is 17.6 X, The temperature is 25°C. The concentration of one solution is fixed at 2.0 M whereas the concentration of the salt solution varied is in the range of 0.02 to 2.0 M, where 5a > 5. Hence, the conditions of Chapter IV are valid throughout. Subject to the external conditions considered above, we now compute the macroscopic osmotic pressure n across the charged membrane against the concentration 20 varied. It is apparent that at very large 5a the effect of the concentration dependence of the dielectric 93 constant on potential becomes negligibly small and, hence, in computing 0 , in general, we may neglect the contribu- £0 tions due to L2 and L3 in the equation. To clarify this, we perform numerical computation on ”(o and Rio, where fig and n' are defined as o 80 ~ 2 2 2 2 4,0 = 2 {c (0) RT (l-q)/€a} [A Ea {IO (Ea) - I (a )21/2 + 2L 1 (5 43) l a 2 ' and 4' = c (2) RT (1 - q) A2 {I (a )2 - I (a )2} (5 44) lo 0 a l a ‘ Note also that 0i 0 zero—order solution in W and that in writing (5.43) the is the contribution due only to the contribution due to L3 is excluded since it is easily realizable to be immaterial. Some calculated results of n and Rio are listed in Table 5.1. It can be shown £0 .. | I from the results that the effect of “£0 020 020 in general, about 1% at very large Ea. Hence, at very on is, large 5a, equations (5.36) and (5.41) tend to be c (2) RT (1 + q) A2 {Io (Ea) 2 2 -11 (6a) }/2 (5.45) :1 ll mo 2c (2) RT (1 - q) A2 {10 (5a)2 2 1r - Il (Ea) } (5.46) 20 without introducing significant error. These equations are useful in studying and interpreting osmotic flow of 94 Table 5.l-—Some calculated values of Rio and ”'0 at vargous salt concentratigns. c (£) = 2.0 10‘ mole/m3. a = 17.6 A. I0] = 10’3 C/mz. t = 25°C. ° ‘3212763-3' .. “(cg/:03 3837' 0.02 5.819 0.291 0.2928 0 05 5.862 0.280 0.2818 0.80 6.851 0.105 0.1059 0.95 7.032 0.085 0.0854 1.55 7.714 0.027 0.0274 1.97 8.158 0.002 0.0015 water across various loose or porous charged membranes, when there is no initial hydrostatic pressure difference across the membrane. Making use of (5.46), 010 is plotted against the logarithm of the salt concentration varied in Fig. 5.1. On the basis of our calculation, it is found that 0 required decreases monot0nically with increasing con— £0 centration varied, and that7&0 is positive when 0 (0)/c (£) < 1. Hence, our results are in good agreement with the experimentally observed phenomena that the flow of water Inoves toward the more dilute solution. This is contrary 'to the flow of solutes which tends to move toward the more II...“ 4:“: .oomm 0 .ms\0aos moa x o. m n 151 0 s\0 3 n _6% .101 o .qoaumuuaqoqoo pawn mcu MO San. Homoa mcu mo coauoc m m m AME no mmasofl Gav Can we uoamula.m .on + 2: 0 ca hom.b mm.m . m . . . 4 . . . . q . . . . . . . 4 . . .mN.OI 95 mmm.o H n mm~.o ll (6 .hvh.m 96 concentrated solution in charged membranes. The latter phenomena is referred to as "anomalous osmosis," which we shall disucss in more detail in the next chapter. E. Conclusion and Discussion Considering the field dependence of the pressure, i.e., Bp/Bw # 0, for the first time we have obtained, for a v-component system, and an uni-univalent electrolyte system, the macrosc0pic pressure distribution equations which simultaneously take account of the concentration dependence of the pressure. As a result, the macroscopic formula for the osmotic pressure 0 across a moderately £0 charged membrane has been obtained for an uni-univalent electrolyte system and, accordingly, the relative distri— bution of water in two aqueous phases has been calculated. We have also deriVed, for an uni-univalent electrolyte system, the macroscopic formula for the osmotic pressure ”mo between the membrane phase m and the aqueous phase 0 and hence calculated the relative distribution of water between the phases. These equations are useful in the treatment of osmotic flow phenomena of water in a porous, moderately charged membrane at steady state, subject to various salt concentrations and geometrical conditions of the membrane in question. Consequently, the fluid- membrane system dealing with an uni-univalent electrolyte 97 solution can now be described more adequately, and experi- ments in the study of transport phenomena of water across a porous, moderately charged membrane can now be inter- preted more accurately. CHAPTER VI THE STEADY-STATE PHENOMENOLOGICAL THEORY OF OSMOSIS IN CHARGED SYSTEMS A. Introduction Osmotic flow phenomena of fluids across a charged continuous phase separating two aqueous solutions of dif- ferent concentrations have been known for years to be important and vital in the understanding of various mechanisms of transport process encountered in many areas of physical and life sciences. It has been realized experimentally (Grim, 1957) that the flow in a charged continuous phase, in contrast to the normal experience obtained with uncharged phases or with non- electrolyte solutions, occurs toward the more concentrated solution (anomalous positive osmosis), and its rate is roughly proportional to the concentration difference. Moreover, when the concentration of one solution is fixed and that of the other is varied, plots of the flow rate against the logarithm of the varied concentration often give an N-shaped curve. Since the early findings of Dutrochet (1835), various transport theories have been developed (gf. 98 99 Kobatake and Fujita, 1964; Gross and Osterle, 1967; Toyoshima, 1967; Fujita and Kobatake, 1968), in order to interpret the mechanism of osmotic flow in a charged membrane. Although most of the theories are satisfactory in many respects, they are all inadequate in one way or another—-e.g., the theory of Kobatake and coworkers con- tains many contradictory assumptions, while the theory of Osterle and coworkers is restricted to extremely dilute solutions. Moreover, the conditions of numerous experi- V'. H . .. _ ments so far reported have usually not been well-defined. As has been known for years, membranes--artificial or natural--vary so widely in structure and function that it is impossible to say that anything approaching a general membrane transport theory has been established. In this chapter, we make use of the principles of nonequilibrium thermodynamics and the equations of hydrodynamics without recourse to most of the restrictive simplifications re- quired by previous workers, and we derive a steady-state phenomenological theory that accounts for the osmotic flow of fluids across a charged continuous phase. It is hoped that the theory can be used better to describe and understand transport phenomena in charged membrane systems. In the following treatment, we consider the system which is composed of a moderately charged membrane for 3 2 which the fixed surface charge density is about 10- C/m 100 and which separates two aqueous uni—univalent electrolyte solutions of different concentrations at the same tempera- ture without the presence of initial hydrostatic pressure difference. A capillary model is used for the membrane. We make use of the macroscopic distribution equations of ions, pressure, and electrostatic potential in the trans- port equations derived in the previous chapters. The re— sulting steady-state differential equations are then solved to yield the average osmotic flow rate as a function of initial salt concentrations. Finally, the theory is com- pared with experimental observations. B. Transport Eguations As mentioned in Chapter II, the transport equations dealing with electrochemical systems are conveniently written with respect to the solvent-fixed (SF) frame of reference. In the following treatment we consider the fluid-membrane system in which the conditions of Chapter IV are all satisfied. Moreover, we hereafter confine the discussion to the system which separates two aqueous solutions containing the solvent molecules and a single uni-univalent electrolyte of the same kind. Positive ions and negative ions are denoted by l and 2, respec- tively. It is of course quite important for biological and other purposes to consider also solutions of many 101 components. In that context, our purpose in this chapter, and in this thesis, is to lay the foundation upon which a general theory can be built. Transport equations for multi-component systems are similar in form to those for a binary system, but they are so long and complicated that they could obscure our simple, central purpose. The diffusional fluxes of species 1 and 2 in the direction of the capillary axis relative to the SF refer- ence frame are, according to (2.78), (jl)O = - (02152 + QEZV§)(3p/8x) - 911 RT (alncl/ax) - QEZRT (alncz/ax) - (011218 + 0122 2F)(80/ax) (j2)o = - (021Vl + 922V 2)(Bp/ax) - Q; lRT(81ncl/8x) - 032RT(81nc2/8x) - (021218 + 0222 2F)(80/8x) , (6.1) where we also make the important assumption that the total electrostatic potential 0 is separable into a part 9 which depends only on x and a part E which depends only on r: I) (x,r) = 9 (x) + E (r) . (6.2) The second term, W (r), on the right-hand-side of (6.2) is what we have considered exclusively in Chapters II--V. A deeper analysis may reveal that the separation of (6.2) is not adequate [e.g., it may well be that w (xr) = 0 (x) + W (x,r)] to satisfy all the governing equations. wamutr.mrmi . 102 It should certainly be an excellent approximation physi- cally, however, since it is clearly experimentally pos- sible to vary independently the trans-membrane potential difference A¢ = ¢ (£) - ¢ (0) and the charge. Also in (6.1), we have taken in fa to be constant, and Cd is the concentration of species a in mole m-3, za (a = 1,2) is the charge valency of species a per mole, R is the gas constant, T is the uniform temperature, F is the Faraday constant, and the 9:8 (d,8 = 1,2) are the phenomenological coefficients in the SF reference frame. We see that for the problem at hand the solute ions are subject to three kinds of forces, namely, osmotic forces corresponding to concentration gradients, mechanical forces corresponding to pressure gradients and electrical forces corresponding to potential gradients. The solvent molecules, however, are subject to only osmotic forces and mechanical forces according to (5.13) in Chapter V. These "forces" are of course not independent--there are only two independent "forces," the chemical potential gradients of l and 2. The partial electric current and the diffusion flux of the salt along the capillary axis relative to the SF reference frame are, by (2.78) and (2.79), 2 (i ) = Z 2a F (jd)o (6.3) d=l 103 (6.4) By (2.65), the electric current and the diffusion flux of the salt along the capillary axis both relative to the capillary wall are, then, c u (6.5) 2 Z zd'F(ja)o + 2 F c u (6.6) d=l a=1 a a d=l a where 110 is the axial component of the reference velocity in SF frame of reference which, under the assumptions which give (2.81), satisfies the axial component of the Navier-Stokes equation in the form (ap/ax) + ‘z'F (dCD/dx) = nr-l[(d/dr)r(duO/dr)] , (6.7) where we have used (6.2) and where we assume that uO is independent of x. In (6.7), n is the isothermal shear viscosity of the fluid, taken as a constant, and E is de- fined by (2.12). Poisson's equation which relates the electrostatic potential to the excess surface charge density is, by (4.1) and (6.2), {1 [(d/dr) (rD(d$/dr))] = — “2p , (6-8) 104 where D is the dielectric tensor of the electrolyte solu- tion, taken as nonconstant. This equation was solved in Chapter IV, under the conditions listed there. For uni- univalent electrolytes, we have, by (4.99), I = (RT/F) A10 (5) [1 + ¢l (£)] , (6.9) where E Kr, A = 4noF/KDORTI (Ka), K = (817F2 E/DORT)l/2, l and $1, defined by (4.98), results from inclusion of the concentration dependence of the dielectric constant. We ‘v‘mnma have also, by (5.13), (Sp/8x) = (dpo/dx) + 2RT(coshW—l)(dc/dx) , (6.10) where W = FW/RT, where po (assumed a function of x only) is the pressure when $ = 0, and where c is assumed a function of x only-~thus, c is effectively, the average over any cross section of the capillary. Finally, for any uni-univalent electrolyte, by (4.31), ‘z' = - 2c sinh‘i’ , (6.11) C. General Formula for 110 In the following treatment, we combine all the differential equations, solve the resulting equations sub- ject to the appropriate boundary conditions and derive a steady-state phenomenological theory of osmosis in a 105 charged continuous phase for an uni-univalent electrolyte system. It has been realized experimentally (Sollner, 1945) that anomalous osmosis of fluid occur only if the continuous phase is in a charged state and is porous to some degree. It is evident that anomalous osmosis does not occur for semipermeable phases. Thus, we take a and Ka very large compared to unity. We have demonstrated in Chapter V through numerical computation that at very large Ea = Ka, (6.9) can, in prac- tice, be represented by the zeroth-order solution of the form E = (RT/F) AIO (£) (6.12) since, for large Ka, A is very small. Making use of (6.10) and (6.11), we rewrite (6.7) as nr‘l [(d/dr) (rduO/dr)] = F (x,r) (6.13) where F (x,r) = de/dx + 2RT (coshW-l) (dc/dx) - 2chinhW (d¢/dx) . (6.14) we solve (6.13) subject to (2.83), yielding the expression, a _ r nuo = I r l I F (x,r) rdrdr (6.15) r O ._ -‘TI'IS_ATTA f'i r W 106 Introduction of (6.14) into (6.15) u0 = + {(aZ-r2)/4n}(dpo/dx) + (RT/n) Ia r-1 Irwzrdrdr(dc/dx) . r 0 a -l r' — (2cF/n) I r 0 Wrdrdr (d¢/dx) r where we have used cosh? = 1 + wZ/z , sinh) = w Substitution of (6.12) into (6.1) yields u0 = + {(E: - 52)/4K2n}(dpo/dx) C + (AZRT/2K2n) 1 ag (Ii-Ii) d§(dc/dx) E 2 - (2AcF/K n) {10(Ea)-IO(€)} (d¢/dx) where we have made use of the substitutions and where we have used the recurrence formulas d (aIl)/da= 5:0 . d IO/da = II . a I 51: d6 = £2 (Ii—Ii)/2 0 (6.16) (6.17) (6.18) (6.19) (6.20) 107 D. Special Formula for 1.10 and its Consequences Although the general formula (6.18) could be used for any experimental situation which fulfills the require- ments previously listed in Chapter IV and in this chapter, the occurrence of the integral term on the right-hand-side would require numerical integration. In order to obtain formulas to which we may easily attach physical signifi- cance, we now specialize to the case in which the second term on the right-hand-side of (6.18) is negligible com- pared to the third term. The requirement we find, (6.22), is so easily met that we may claim with confidence that our "special" formulas are in fact those which apply to the great majority of experimental situations. Rather than deal directly with (6.18), we find a stronger condition by requiring, in (6.14), |2RT (coshW-l) (dc/dx)l : 0.01 |2chinhW(d¢/dx)| (6.21) By (6.17) and the definition of W (E FE/RT), (6.21) becomes |(d 2n c/dx)/(d¢/dx)| 3 0.02|'()7|‘l (6.22) In Chapter IV, we required [WI 3 0.245. For 298°K, F/RT = 38.93 V"1 and we therefore have [El < 0.0063 v , IEW—l 3_158.9 v‘1 (6.23) -lll I'll 108 So for the largest allowed value of |$|, we get |(d 2n c/dx)/(d®/dx)| 1.3-2 (6.24) Ordinarily, I$l is much smaller than 0.0063 V because it decreases strongly with the square root of average com- position and with increasing pore size. Noting that it is fin c which appears in the denominator of (6.22) and (6.24), it is clear that (6.22) will ordinarily be easily satisfied. Typical values of the trans-membrane potential A¢ are 60 to 100 mV (Woodbury, et_al., 1970) and therefore (6.24) would be satisfied for IA£n<3Ii 0.3. For smaller, more typical value of (0|, IAln cl may be even larger but still negligible. We therefore reduce (6.18) to 2 2 no = {(5a - a )/4K20} (dpo/dX) 2 - (ZACF/K 0) {IO (Ea)-Io(€)} (d¢/dx) (6.25) Equation (6.25) is generally valid for loose membranes with very large pores. Moreover, it is mathematically more tractable. The average values of the reference velocity, electric current density, and the diffusional flux of the electrolyte component over the cross-section of the capillary are [Mn fa‘a‘auil Hw'F‘E'JT 109 U - 1a 2 ( ) d a 2 d o - 0 nr uo r/(I) TTI‘ r _ a a I = I an (I) dr/I 20rdr 0 0 _ a a J = 0 an (J) dr/I 20rdr 0 Introducing (6.25) into (6.26), there follows U0 = A01 (de/dx) + A02 (d¢/dx) where the coefficients A and A are 01 02 2 2 A01 _ ga/8K n A 02 - (4AcF/€:Kzn) {6: 10(Ea)/2-€aIl(€a)} (6.26) (6.27) (6.28) (6.29) (6.30) where we have used (6.20). It is interesting to note that in the coefficient A02 of (d0/dx), usually referred to as the electroosmotic coefficient, is a function of the salt concentration. Previous existing theories for the anomalous osmosis (Schlogl, 1955; Kedom, 1961; Kobatake, 1958) have not taken this fact into account. As we shall see later, this effect is essential for the observed phenomenological behavior of the osmotic flow in charged membranes. We substitute (6.1) into (6.5) and (6.6) and hence have, for the general case, Wm~m.mzn:n:r 3mm) 110 2 o —o o —o 2 o o J = - E (levl + Qa2v2)(8p/3x) - E (901 + 0&2) d—l a—l 2 o 0 RT (dincaO/dx) ail (90121 + 9&222)F(d¢/dx) 2 + ail Cao exp (-zaW) uO (6.31) 2 o —o o *0 =.. F I 6:1 (Qalvl + QaZVZ) 2a (Sp/BX) 2 o o — ail (gal + Qaz) z£7RT (dincaO/dx) 2 o o 2 - ail (Qalzl + 0&222) zaF (d¢/dx) 2 + ail Cao exp (-zaW) zaFuO (6.32) where we have made use of the substitution, dflnca/dx = dincaO/dx , (6.33) since (6.33) is assumed independent of radical coordinate. For an uni-univalent electrolyte system, there result 2 J = — 2 (0:16? + 03263) (ap/ax) a=l 2 o o 2 o o - 2 (gal + QGZ)RT(d£nc/dx) - Z (Gal—0a2)F(d¢/dx) d-l a=1 + 2c cosh? 110 (6.34) ‘mmrmmfiqugwmmzr 111 2 _ _ o —o o -o _ 0+1 I — ail (Qalvl + Qazvz) ( 1) E’(3p/3x) 2 o 0 +1 - Z (0 1 + 0 2) (-l)a F RT (anC/dx) d=l a 2 o o d+1 2 - 2c F sinh? 110 (6.35) The practical equation for the axial pressure gradient (Sp/3x) can be deduced from (6.10), yielding ap/ax g de/dx + RT )2 (dc/dx) (6.36) where we have come used (6.17). Hence, (6.34) and (6.35) be- 2 : (0: 6° + 0° 6°) (de/dx) a l 1 1 a2 2 2 o —o o —o 2 :1(Qalvl + Qazvz) RT W (dc/dx) 2 o o 2 o o ail (gal + Qa2)RT(d2nc/dx) - a:l(Qal-Qa2)(d¢/dX) 26 (1 + 22/2) uO (6.37) W¥-fiflmvnmsrrmrz-ml ‘. 112 1 — - i (0° 6° + 0° ’°) (-1)°‘+l F (d /d ) ‘ = 61 1 62V2 po X 2 - z (0: 6° + 03263) (-1)""'1 F RT (2 (dc/dx) 2 - 2 (031 + 032) (-1)°+1 F RT (dlnc/dx) d=l - (0:1 0:2) (-1)O‘+1 F2 (d¢/dx) =1 + 2cFWub , (6.38) where we have used (6.17). Introduction of (6.37) and (6.38) into (6.27) and (6.28) and use of (6.12) and (6.19) yield + o —o + Q0 —0 a l 61V1 a2v2) (’1’ F (dpo/dX) o —o 0+1 2 2 01 1 + Qa2v2)(-1) {(ZA FRT/ga) a 2 I 51 d5} (dc/dx) 0 o 2 - 2 (0° + no 0+1 a1 02) F RT (dinc/dx) {-1) d+1 2 ° ° -1) F (d6/dx) 2 - g (9.1 - 9.2) < 2 5a + (4AcF/ga) I gIOuOdg (6.39) 0 Wfi“fi:flflr“?‘fiw1 . 113 2 — _ _ o —o —o J - ail (levl + QaZVZ) (de/dx) 2 o —o —o 2 2 5a 2 — ail (levl + 002v2)(2A RT/Ea) 0 EIO d5 (dc/dx) 2 o o 2 o o _ 0:1 (901 + QaZ)RT(d9.nc/dx) - a:1(Q°1-Q°2)F(d¢/dX) E F; E‘ + 2 a 2 2 h (4c/ga) I 5(1 + A 10/2) uO dg . (6.40) E O . E The integrals in (6.39) are readily evaluated, yielding E 4 5 a _ 2 2 _ 2 0 SI d5 — 5a {IO(€a) Il (5a) }/2 (6.41) and E where we have used (6. 2 (l/ZKzn) 4a I2 (Ea)(dpo/dX) (2AcF/K20) [gale (Ea)Il (ea) 52 a 2 (a ) - 11 (5a)2}/21 (d4/dx) (6.42) {10 a 20) and the recurrence formula 3 a 2 d: 4 Il (Ea) - 26a 12 (Ea) (6.43) Nevertheless, it is apparent that at very large Ea, the integral in (6.40) E a I 0 2 g{1 + A IO (6)2/2} no 65 (6.44) 114 tends to be 5 1a gu dg (6.45) 0 0 because A + O as Ka + w. This can be evaluated easily in closed form, yielding the following result, (aa)/2 + (Si/IGKZU) (de/dx) - (2AcF/Kzn){§: IO ,7 1‘1er gain. I - Ea I1 (53)} (d0/dx) (6.46) -..l‘l‘ Introduction of (6.41), (6.42) and (6.46) into (6.39) and (6" ‘1 (6.40) finally yields, in the general case, I = BOl (de/dx) + B02 (d¢/dx) + BO3 (dc/dx) + BO4 (dlnc/dx) (6.47) and J = ROl (de/dx) + R02 (d¢/dx) + Ro3 (dc/dx) + RO4 (anC/dx) , (6.48) where the coeff1c1ents BOi and Roi (1 = l,....4) are B — - g (0° -0 + 0° ‘0)(-1)OL+1 F + (2AcF/K20) I (E ) Ol ‘ 6:1 61V1 azvz 2 a B _ _ g (0° _ Qo )(_l)d+1 F2 02 0:1 dl 02 2 2 2 2 2 - F (8A 0 /€aK n) [Ea Io (Ea) 11 (Ea) 2 2 2 - 4a {10 (a ) - 11 (Ea) 1/21 a 115 2 o —o o —0 0+1 2 2 2 B03 = - 0:1 (Qalvl + Qazvz) (-l) (A FRT){IO(§a) -Il(€a) } 2 o 0 0+1 B04 = - 0,21 (901 + {202) (-1) FRT 2 o -o o —o 2 2 R01 = - 0:1 (Qalvl + Q02v2) + Ea C/4K n 2 o o 2 2 2 2 R02 = - 0:1 (901 - 9&2) - (8Ac F/K 05a){€a10(€a)/2-§a11(§a)} 2 o —o o —o 2 2 2 R03 = ' 6:1 (Qalvl + Q62V2) (A RT) {Io (ga) —Il(€a) } 2 o 0 R04 = - 0:1 (901 + 0&2) RT (6.49) Although the (dc/dx) term and the (d in c/dx) term could obviously be combined in (6.47) and (6.48), and elsewhere, no useful simplification is thereby obtained. Formally, the coefficient of(d.£n c/dx)ixx(6.47) would be (5 B03 + 304) and the coefficient of(d.£n C/dX)le(6.48) would be (5 R034-Ro4), where E is some average value of c. E. The Working Equation for Anomalous Osmosis We have so far obtained explicitly, making use of some eXplicit, justifiable simplifying assumptions, the equations for the averages of reference velocity, electric current density and diffusional flux of the electrolyte WHIP-immanmmrma , 116 component over the cross section of the capillary. From these equations we now derive a working equation which can be used to interpret the characteristic behavior of osmotic flow through a charged membrane. As mentioned before, we consider that initially there is no hydrostatic pressure difference between the two sides of the membrane, and assume that the system has attained a steady state. The steady-state assumption is justifiable since it is real- izable in most biological systems and physicalchemical Wmm1nnmr1‘m‘fi’) systems. Returning to (6.29), (6.47) and (6.48), at steady state, from the conservations of mass and electric current density and from the incompressibility of the fluid there follow d UO/dx = 0, df/dx = 0, dE/dx = 0 . (6.50) Furthermore, in the present system there is no applied (electric field across the membrane and, hence, from (6.60) we have Uo = constant, T = 0, 3 = constant. (6.51) 11ccording to the assumptions made above the apprOpriate boundary conditions for the problem are, then, c = c (0) at x = 0 and c = c (£) at x = 2 117 =(O)atx=0and4>=(£)atx=IL (6.52) Making use of the conditions given by (6.51) we now integrate (6.29), (6.47), and (6.48) over the length of the capillary, i.e., 0 i x i 2, subject to (6.52). However, by examination of (6.30) and (6.49) we imme- diately notice that the coeff1c1ents A02, B01, B02, R01 and R02 are concentration-dependent. Hence, we may write . ‘_ ‘Tg'i-mflfifpv _ R _ a; U0 — AOl (Ape/1) + 6 A02 (d¢/dx) dx/Q — constant (6.53) I = I2 E (dp /dx) dx/E + I2 B (d¢/dx)dx/£ 0 ol 0 o 02 + BO3 (Ac/2) + Bo4 (Ainc/E) = 0 (6.54) 3 — )2 R (dp /dx)dx/£ + 1" R (dCI>/dx)dx/IL 01 o o 02 + RO3 (Ac/R) + RO4 (Alnc/l) = constant (6.55) \nhere Apo, Ac, and Alnc are defined as ApO = po (x = i) - po (x = 0) Ac=C(£)-C(0) A£nc = inc (2) - inc (0) in {c (2)/c (0)} (6.56) FHthhermore, we have required in Chapter V that 118 P0 = (APO/R) X + pO (x = 0) 0 II (Ac/IL) x + C (0) 0 II (04/2) x + 4 (0) , (6.57) whence dpo/dx = ApO/£ dc/dx Ac/l ‘f’merww‘m 1.1:). d0/dx A0/2 (6.58) With the aid of the above equations, the evaluation of the integrals in equations (6.53)-(6.55) then can easily be performed. The resulting expressions are 2 _ 2 2 2 3 A02 (d¢/dx)dX/£ — - (ZAF/EaK n) c (£)(l+q){€aIO(€a)/2 - Ea Il (5a)} (04/2) (6.59) It B (d /d )d /2 " (A /2) [- g (00 VG + 52° VON-DOW]? 0 01 p0 X X _ po 0:1 01 l 02 2 - (AF/K20) 12 (6a)c(4)(1+q)1 (6.60) 1% B (66/6 )d /2 — (44/2) [- 2 (0° -0° )(-1)°‘+1F2 0 02 X X ‘ “=1 01 02 -(8A2F2/E:K2n){{c (9.)2 (1 + q + q2)/3} 2 2 2 [Ea IO (6a) 11 (Ea) - 6a {10(Ea) -Il(€a) }/2]}](6.61) 119 2 2 o -o o —o 0 R01 (dpo/dx)dx/£ = (Ape/2) [- 0:1 (0alvl + 0a2v2) - (ii/BKZU) c (2) (1 + q)] (6.62) 1% R (d¢/dx)dx/£ = (0¢/£) [- g (00 - 0O ) F 0 02 OF]. a]- 012 - (8AF/K2€:n){{c (IL)2 (1+q+q2)/3} 2 {Ea IO (Ea)/2 - 5a I1 (Ea)}}] (6.63) “ mac-0731::ms‘ . We introduce equations (6.59)-(6.63) into equations (6.53)-(6.55). From (6.52), there follow “F Uo = A02 (0¢/£) = constant (6.64) 0 (6.65) I (Afinc/fi) B (00/0) + BO 02 (Ac/2) + BO 3 4 J = R02 (00/2) + RO (Ac/2) + Ro4 (Alnc/i) constant (6.66) 3 where A02, 302’ and R02 are given by __ 22 2 _ A02 - - (ZAP/EaK n){c (£)(l+q)}[€a 10(Ea)/2 Ea11(Ea)] (6.67) 2 — o 0 0+1 2 B02 = - E (901 - Q02) (-1) F 0—1 - (SAZFz/EiKZn){{C (1)2 (1+q+q2)/3} [ I ( ) I ( )- 2(1 ( )2-1 ( )2}/2]}(6 68) 5a o E;a l E(a ga o 8‘:a 1 ga ' — 2 o o 2 2 2 2 R02 = - ail (0&1 - 0&2)F- (8AF/gaK n)HC(£) (1+q+q )/3} [5: 10 (ga)/2 - 5a I1 (ga)]} (6.69) 120 Our chief goal, however, is to obtain explicitly the expressions for U0 and 3 as functions of salt concen- trations. Hence, we combine (6.64) through (6.66) appro- priately and rearrange the resulting equations. There follow, then, UO 2 = (AOZBo3/B02) Ac - (AozBo4/B02)A2nc (6.70) g 6 a“ J 2 = {R03 - (R02303/Boz)} Ac f + {Ro4 — (ROZBO4/B02)} A2nc (6.71) which give (a) 2 and 3 2 as functions of salt concentra— tions. As we shall see later, these equations can be used to study and interpret osmotic flow phenomena of fluids in loose, moderately charged membranes. F. The Onsager Trans— port Coefficients Before we can proceed further to investigate our results numerically, it is essential to acquire consider- ably more knowledge of the physical and chemical proper- 'ties of the linear phenomenological coefficients, or Onsager Coefficients, 9:8. For simplicity, we follow the treatment of Miller (1966) since it is probably the best reference which deals with the determination of 121 I I 0 I O O ionic transport coeff1c1ents QaB’ for isothermal vector transport process in binary electrolyte systems based on the SF reference frame. Consider an isothermal system consisting of a neutral solvent, e.g., water, and a binary electrolyte j I which ionizes completely into the solution giving two ions. The independent flows for the system are those of the two ions since (30)o E 0 for the SF reference frame. Hence, the transport equations are IMmew-Pm‘wmnr + i + (2° :2 (6.73) where the 2a (a = 1,2) are the conjugate forces, as given explicitly in Chapter II. These equations completely describe the isothermal vector transport properties in a binary electrolyte solution provided the 9:8 are known as functions of the temperature T, the pressure p, and com- position. The 9° a8 because they arise from a fundamental thermodynamic theory are the fundamental transport coefficients as outlined in Chapter II. Any isothermal transport process in a binary system is completely characterized by equations (6.72) and (6.73) together with a knowledge of the Q: as functions of c, T, p. Moreover, the same 8 122 9:8, apply when the phenomena is two or three dimensional, e.g., where an electric field is perpendicular to the diffusion direction. Consequently, even though the con- centration dependence of the 9:8, is determined experi- mentally from the one-dimensional special cases, the re- sulting numbers can be applied to any process no matter how complex. The experimental Onsager transport co- m: wm‘r'“?! ( efficients for some uni-univalent electrolytes are listed in Table 6.1. “_lin‘fl ' I " . - ‘ I 1:! G. Results and Discussion In the following treatment we make use of equation (6.70) to study anomalous osmosis. We consider a porous, moderately charged membrane for which the fixed surface charge density '0' is 10.3 C/m2 and capillary radius 150 g, whence Ka >> 1 in the range of the salt concentrations used for this study. In particular, we assume a negatively charged membrane with 0 = - 10-3 C/mz, and choose the following experimental conditions: the salt used is NaCl, the temperature is 25°C, the isothermal shear viscosity of water, n at 25°C is 8.903 X 10-10 J sec cm-B, the electric permittivity of water, i.e., Do/4W, 10 1 -l at 25°C is 6.94 x 10. C V- m , and the partial molar volumes of Na+ and Cl- ions at infinite dilution are -l.4 and 18 cm3 mole-l, respectively (see also Table 4.1, 123 Table 6.1a--The Onsager transport coefficients for H O-NaCl at 25°C. 2 '6', 1012 x (221/6, 1012 x (232/5, 1012 x (232/6, mole/2 mole cm2/J sec mole cmz/J sec mole cm2/J sec 0.0000 5.381 0.000 8.201 0.0010 5.363 0.026 8.177 0.0005 5.341 0.058 8.146 0.0010 5.325 0.081 8.123 0.0050 5.263 0.170 8.036 0.0100 5.219 0.233 7.974 0.0500 5.065 0.440 7.742 0.1000 4.971 0.554 7.601 0.2000 4.851 0.682 7.435 0.5000 4.613 0.840 7.121 0.7000 4.484 0.882 6.950 1.0000 4.311 0.911 6.772 1.5000 4.053 0.923 6.370 2.0000 3.812 0.911 6.035 2.5000 3.581 0.884 5.708 3.0000 3.366 0.858 5.393 ~ ..f1;_m1" 'r l .b . | 124 Table 6.1b--The Onsager transport coefficients for E, 1012 x 021/5, 1012 x 0:2/5, 1012 x 052/6, mole/2 mole cmz/J sec mole cmZ/J sec mole cmZ/J sec - 0.0000 7.892 0.000 8.198 E 0.0001 7.872 0.026 8.176 § 0.0005 7.844 0.059 8.148 E 0.0010 7.826 0.086 8.129 1 0.0050 7.746 0.187 8.045 0.0100 7.694 0.256 7.991 0.0500 7.520 0.503 7.810 0.1000 7.430 0.647 7.715 0.2000 7.331 0.809 7.613 0.5000 7.193 1.038 7.478 0.7000 7.140 1.124 7.425 1.0000 7.077 1.214 7.365 1.5000 6.972 1.304 7.265 2.0000 6.866 1.362 7.160 2.5000 6.754 1.404 7.050 3.0000 6.634 1.440 6.929 125 Table 6.1c—-The Onsager transport coefficients for H20-LiCl at 25°C. 1 E, 1012 x 0‘1’1/‘5, 1012 x 0‘132/6, 1012 x (232/6, mole/2 mole cmz/J sec mole cmZ/J sec mole cmz/J sec 0.0000 4.153 0.000 8.197 0.0001 4.137 0.021 8.166 0.0005 4.115 0.051 8.136 0.0010 4.103 0.073 8.112 0.0050 4.046 0.162 8.020 0.0100 4.011 0.223 7.956 0.0500 3.870 0.417 7.718 0.1000 3.774 0.513 7.547 0.2000 3.624 0.614 7.290 0.5000 3.316 0.700 6.827 0.7000 3.159 0.714 6.593 1.0000 2.942 0.699 6.286 1.5000 2.662 0.677 5.877 2.0000 2.406 0.622 5.475 2.5000 2.156 0.548 5.073 3.0000 1.954 0.500 4.718 1W Wfi‘tfinmmmsm) . 126 Chapter IV). The experimental Onsager transport co- efficients for the NaCl-HZO system at 25°C are taken from Table 6.1. The ratio of the concentrations of two solu- tions is fixed. Plots of U0 (2) calculated from (6.80) at q = 0.5 and 0.25 against the logarithm of the salt concentration varied are shown in Fig. 6.1. It can be seen from Fig. 6.1 that bell-shaped curves are obtained .. WWfi‘rr—mr' for both cases, and the maximum flow rate at q = 0.5 occurs approximately at the salt concentration of 0.06667 mole 2-1 while that at q = 0.25 occurs approximately at the salt concentration of 0.08 mole 2-1. Moreover, to see that our equation is also useful for positively charged membranes, we plot Uo 2 calculated 3 C/m2 against from (6.70) at q = 0.5, taking 0 = + 10‘ the logarithm of the salt concentration varied. The re- sult is shown in Fig. 6.2. It is found that the values of U0 2 are all negative. This indicates that the direction of the osmotic flow of fluid is changed and the flow moves toward the more dilute solution (anomalous negative osmosis) as can be seen from Fig. 6.2. As a result, it is apparent that the direction of the osmostic flow of fluid in charged membranes changes as the sign of 0 changes. This fact has not been ex- plained by previous theories (sf. Kobatake, 1964). The failure of previous theories to account for this fact cm sec (1), 10’9 70+ 60‘ 50‘ 404 30‘ 20‘ 10‘ 127 II 1 1 1 1 —ll -10 -9 -8 -7 -6 2n c (0) + Fig. 6.1-—Plot of U0 2 (in cm2 per sec) as a function of the logarithm of the salt concentration, c (0). Curve I, q = 0.25. ‘Curve II, q = 0.5. The glectrolyte used is NaCl. 0 = -10-3 C/mz. a = 150 A. t = 25°C. U (2), cm sec 10 -1O -20 -30 -4O -12 Fig. 6.2-—Plot of U 2 (in cm2 per sec) as a function of the logarithmoof 0.5. 0 = 10 a = 150 X. 128 4 2n c (0) + -7 She salt concentration, c(0). C/mz. The electrolyte used is NaCl. t = 25°C. q: VMH“na.wr:—m:m 129 apparently arises from the use of restrictive simplifica- tions in their treatment of the theories, which we have discussed before. These restrictions have been excluded in our theoretical development. Hence, we are confident that our equation is much more complete and better in describing various membrane systems. Consequently, the experiments in study of osmotic flow phenomena in charged membranes can now be interpreted more accurately, in principle. Unfortunately, so little experimental infor- mwgwnr-tm'mflmw’ mation is available regarding the actual values of mem- brane charge and pore size that we cannot make a full comparison with eXperiment. Perhaps our equation can be used to determine charge and pore size in conjunction with fully—characterized experiment. The shapes of our curves are qualitatively the same as those found experi- mentally (Kobatake and Fujita, 1964; Tasaka, Kondo and Nagasawa, 1969) and we certainly predict the proper direction of flow. In the theoretical development of this chapter, we have made use of the assumptions that Ka >> 1, and that the gradient of 2n c is small enough compared to the trans-membrane potential that (6.22) is satisfied, and have adopted only the zeroth order solution for the potential. Although we have clearly made only a start at obtaining a complete theory of transport through 130 membranes, it is at last, a well-formulated, self— consistent start. Hopefully, we have laid the founda— tion sufficiently firmly that future investigators, including ourselves, will not need to reexamine the foundation, but may instead build upon it. I "’E _fl‘n I {‘5 . ,‘ t" 1, fi:W-T"‘r C a? BIBLIOGRAPHY I" .' .'I‘- . I“ - ,___ ‘I " 7 BIBLIOGRAPHY Abramowitz, M., and I. A. Stegun, Handbook of Mathe- matical Functions, National Bureau of Standards, Applied MathematiCs Series, 55, U.S. Government , Printing Office, Washington, D.C., 1964. E: Alexandrowicz, A., and A. Katchalsky, J. Polymer Sci. Al, i 3231 (1963). E Andrews, F. C., Ind. Eng. Chem. Fundamentals 6, 48 (1967). % Bartelt,J;Ih, Ph.D. TheSis,Michigan State University, ;4 1968. , and F. H. Horne, J. Chem. Phys. 51 (l), 210 (1969). , and F. H. Horne, Pure and Applied Chemistry, 22, 349 (1970). Bellman, R., Perturbation Techniques in Mathematics, Physics and Engineering (Holt, Rinehart, and Winston, Inc., New York, N.Y., 1964). Booth, F., J. Chem. Phys. 19 (4), 391 (1951). Chapman, T. W., The Transport Properties of Concentrated Electrolyte Solutions, Ph.D. Thesis, University of California, Berkeley (University Microfilms, Ann Arbor, Michigan, 1967). Coleman, B. D., and C. Truesdell, J. Chem. Phys. 33, 28 (1960). de Groot, S. R., and P. Mazur, Nonequilibrium Thermo- dynamics (North-Holland, Amsterdam, 1962). Defay, R., and P. Mazur, Bull. Soc. Chim. Belges, 63, 562 (1954). , and A. Sanfeld, J. Chim. Phys. 60, 634 (1963). 131 132 Derjaguin, B. V., Research in Surface Forces (Consultants Bureau, New York, N.Y., 1964, 1966). Devillez, C., Sanfeld, A., and A. Steinchen, J. Coll. Interf. Sci. 25, 295 (1967). Dutrochet, M., Ann. Chim. Phys. 69, 337 (1835). Fair, J. C., and J. F. Osterle, J. Chem. Phys. 53 (8), 3307 (1971). Fitts, D. D., Nonequilibrium Thermodynamics (McGraw-Hill, New York, N.Y., 1962. Fujita, H., and Y. Kobatake, J. Coll. Interf. Sci. 21, 609 (1968). Grim, E., and K. Sollner, J. Gen. Physiol. 40, 887 (1957). Gross, R. J., and J. F. Osterle, J. Chem. Phys. 49 (1), 228 (1968). Gyarmati, I., Nonequilibrium Thermodynamics (Springer— Verlag, New York. N.Y., 1770). Haase, R., Thermodynamics of Irreversible Processes (Addison-Wesley Publishing Co., Reading, Massa- chusetts, 1969). Hasted, J. B., Ritson, D. M., and C. H. Collie, J. Chem. Phys. 16 (l), 1 (1948). ' Horne, F. H., J. Chem. Phys. 45, 3069 (1966). , and T. G. Anderson, J. Chem. Phys. _3 (6), 2321 (1970). Hurwitz, H. D., Sanfeld, A., and A. Steinchen, Electrochim. Acta, 9, 929 (1964). Ingle, S. E., Ph.D. Thesis, Michigan State University, 1971. Irving, J., and N. Mullineux, Mathematics in Physics and Engineering (Academic Press, New York, N.Y.,71959). Katchalsky, A., and P. F. Curran, Nonequilibrium Thermo- dynamics in Biophysics (Harvard’Univ. Press, Cambridge, Massachusetts, 1967). 133 Kedem, 0., and A. Katchalsky, J. Gen. Physiol. 45, 143 (1961). _— Kirkwood, J., and B. Crawford, J. Phys. Chem. 56, 1048 (1952) . ‘— Kobatake, Y., and H. Fujita, Kolloid Z. 999, 58 (1964). , and H. Fujita, J. Chem. Phys. 99, 2212 (1964). , Bull. Tokyo Inst. Technol. Ser. B, p. 97 (1959). , J. Chem. Phys. 99, 442 (1958). Kotin, L., and M. Nagasawa, J. Chem. Phys. 99, 873 (1962). Lakshminarayanaiah, N., Transport Phenomena in Membranes (Academic Press, New York, N.Y., 1969). MacGillivray, A. D., and J. J. Winkleman, J. Chem. Phys. 99 (6), 2184 (1966). Manning, G. S., J. Chem. Phys. 99 (6), 2668 (1968). Mazur, P., and I. Prigogine, Mem. Acad. Roy. Belg. 99, l (1953). Miller, D. G., Chem. Rev. 99, 15 (1960). ', J. Phys. Chem. 19 (8), 2639 (1966). Mfiller, I., Arch. Rational Mech. Anal. 99, l (1968). Onsager, L., Phys. Rev. 91, 405 (1931); 38, 2265 (1931) Philip, J. R., and R. A. Wooding, J. Chem. Phys. 99 (2), 953 (1970). Pottel, R., Chemical Physics of Ionic Solutions, ed. B. E. Conway and R. G. Barradas (John Wiley & Sons, Inc., New York, N.Y., 1964). Prigogine, I., Mazur, P., and R. Defay, J. Chem. Phys. 99, 146 (1953). Sanfeld, A., and R. Defay, J. Chim. Phys. 63, 577 (1966). , Introduction to the Thermodynamics of Charged and Polarized Layers (John Wiley & Sons, Inc., New York, N.Y., 1968). 134 Schlogl, R., Z. Physik. Chem. N. F. 9, 73 (1955). Sollner, K., J. Phys. Chem. 99, 47, 171, 265 (1945). Sparnaay, M. J., Rec. Trav. Chim. 77, 872 (1958); 81, 395 (1962). __ '_— Staverman, A. J., Rec. Trav. Chim. 19, 344 (1951). , Rec. Trav. Chim. 29, 623 (1952). Tasaka, M., Kondo, Y., and M. Nagasawa, J. Phys. Chem. 19 (10), 3181 (1969). Toyoshima, Y., Kobatake, Y., and H. Fujita, Trans. Faraday Soc. 99, 2828 (1967). Truesdell, C., Rational Thermodynamics (McGraw-Hill, New York, N.Y., 1969). Verwey, E. J. W., and J. Th. G. Overbeek, Theory of the Stability of Lyophobic Colloids (Elsevier, Amster— dam, 1948). ‘ Watson, G. N., A Treatise on the Theory of Bessel Functions, 2d ed. (Cambridge Univ. Press, Cambridge, England, 1958). Woodbury, J. W., White, S. H., Mackey, M. C., Hardy, W. L., and D. B. Chang, Physical Chemistry, An Advanced Treatise, ed. H. Eyring, Vol. IX/Electrochemistry (Academic Press, New York, N.Y., 1970). A. i .. _. . r .31... NJ...» . «”3 a. Per-L. aft-...EHE a APPENDICES APPENDIX A A DERIVATION OF THE CHAPMAN-GOUY EQUATION FOR A CHARGED CYLINDER Consider a charged cylinder containing an aqueous, binary, uni—univalent electrolyte solution. The Chapman- Gouy equation can be derived from the usual Poisson— Boltzmann equation in the form B B (Do/r) (dr (dw/dr)/dr) = - 40F (c+ - c_) (A.l) where r is the radial coordinate, D0 is the isothermal dielectric constant of water, 0 is the electrostatic potential, F is the Faraday constant, and where c: (a = +,-) is the Boltzmann equation molar concentration of species a, B _ Ca = c exp (-zaF0/RT) , (A.2) where E is the average salt concentration, 2a is the valency of a, R is gas constant, and T is the absolute temperature. The boundary conditions are dw/dr = O at r = 0, and dw/dr = 4no/DO at r = a , (A.3) where o is the surface charge density fixed on the cylinder wall, and a is the radius of the cylinder. 135 136 We introduce (A.2) into (A.1) and make use of the following substitutions: Fw/RT = w r = E/K where K—1 is the Debye length K = /80F26/DORT There results, then, (l/E) (d€(dY/d€)/d€) = sinh? . where we have made use of the transformation We rewrite (A.7) as d2 Y/dgz + (1/5) (dY/dg) = sinh? Equation (A.9) reduces to d2 W/dgz = sinh? , for (1/5) (dY/dg) << d2 Y/dgz, in particular, for Ka >> (A.4) (A.5) (A.6) (A.7) (A.8) (A.9) (A.10) 1 and (dW/dg)o = 0. It is also assumed in this derivation that W vanishes at E = 0. The boundary conditions on W then become, WWI” WW .1 dW/dg 0 and W = 0 at E = 0 (A.11) dW/dg 4noF/DORTK at S = E , (A.12) a where Ea is the value of g at the wall, i.e., Ea = Ka, and where we have used (A.3) and (A.4). We now solve (A.10) subject to (A.11) and (A.12). Making use of the transformation dW/dg = p , (A.13) whence (129/dz:2 = p (dp/dY) , (A.14) there follows, from (A.10), p (dp/dW) = sinh? . (A.15) The solution to (A.15) is immediately obtained, yielding 2 p /2 = cosh? + b , (A.16) where b is the constant of integration which is readily evaluated, subject to (A.11), to give b = -1 . (A.17) Hence, (A.16) becomes p2/2 = coshY - 1 (A.18) .. Vfifi “m Tfi‘. 4 ( “14.1.71- ‘ 1'}! 138 The Chapman-Gouy equation relates the surface potential to the surface charge density fixed on the wall. Hence, we have, from (A.12), (A.13), and (A.18), l 2 _ _ f (4noF/DORTK) — cosh‘i’a 1 , (A.19) where Ya is the value of W at the wall. By rearrangement, (A.19) becomes 0 4 (zeRTE) (coshYa - 1)/2 (A.20) where e is the electric permittivity e = Do/40 . (A.21) From (A.20), there follows, then, 0 = 2 (268TE)1/2 sinh (Ya/2) , (A.22) where we have used sinh (Ya/2) = /(cosh‘i’a - 1)/2 . (A.23) Equation (A.22) is the Chapman-Gouy equation. . 1.: -D :' . v ~ I ‘3 Q . 7" ". r' J {r- . ' '. ... r ' i ‘ -.. ml 0 1 '1‘ ' ‘ ..1 "" 1! m . y.) 4 I 4 : 1‘ \ ‘ ... ( « . .1 - 1"“ I . , . . ) c) . ‘( . ... . l A. I. H! V 4 ~( 1 I I ‘ . " . ’ ‘A v " ' I . , . . ._. r . . H, APPENDIX B DERIVATION OF EQUATIONS (4.87) AND (4.88) In the following we derive equations (4.87) and (4.88). From the recurrence formula, d EIl/dg = 510 , (B.l) we have, in (4.87), E I 3 = 5 2 = E 2 o 610 dg 0 IO (:10) d6 3 Io dg Il . (8.2) By integration by parts, there results 3 _ 2 _ g 0 d6 — 51110 2 g gIlIOdIo (8.3) 15 61 0 Note, however, that dIO/dg = Il , (8.4) whence :5 61 I d1 = 15 61 1 (dI /dg)dg = I5 6121 dg (B 5) 0 1 o o 0 1 o o 0 1 o ' Substitution of (B.5) into (B.3) yields, by rearrangement, equation (4.87), E 2 _ 2 _ g 3 0 gIOIldg — 61011/2 0 gIodg/z (8.6) 139 140 To obtain equation (4.88), we write )5 Elida = 15 I:d(€2/2) (B.7) 0 0 By integration by parts, there results a 2 _22 _62 3 gIodg - a 10/2 I g IodIO (8.8) From (3.4), we have {'1'}. I.,!“ = . El .I '” lllllim. E 2 _ E 2 _ E a a 10910 - 3 6 IO (dIO/da) d6 — 5 (EIO)(€Il)d€ _ E _ 2 2 0 where we have used (B.l). Substitution of (B.9) into (B.8) yields equation (4.88), 2 o (g 41 d: = 62 (I: _ Ii)” (8.10) 0 (mm) 462 WM” 80